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

Resonant absorption of electromagnetic waves in transition anisotropic media

Open Access Open Access

Abstract

We study the mode conversion and resonant absorption phenomena occurring in a slab of a stratified anisotropic medium, optical axes of which are tilted with respect to the direction of inhomogeneity, using the invariant imbedding theory of wave propagation. When the tilt angle is zero, mode conversion occurs if the longitudinal component of the permittivity tensor, which is the one in the direction of inhomogeneity in the non-tilted case, varies from positive to negative values within the medium, while the transverse component plays no role. When the tilt angle is nonzero, the wave transmission and absorption show an asymmetry under the sign change of the incident angle in a range of the tilt angle, while the reflection is always symmetric. We calculate the reflectance, the transmittance and the absorptance for several configurations of the permittivity tensor and find that resonant absorption is greatly enhanced when the medium from the incident surface to the resonance region is hyperbolic than when it is elliptic. For certain configurations, the transmittance and absorptance curves display sharp peaks at some incident angles determined by the tilt angle.

© 2017 Optical Society of America

1. Introduction

In recent years, there have been many studies on the strong absorption of electromagnetic (EM) waves occurring in media with near-zero effective dielectric permittivity ϵ, termed epsilon-near-zero media, and in transition metamaterials inside which ϵ varies continuously from positive to negative values, passing through a narrow resonance region where the real part of ϵ is identically zero [1–14]. In the presence of an infinitesimally small amount of damping in the resonance region, a large absorption of the EM wave energy occurs. This strong resonant absorption arises due to the mode conversion of incident p-polarized EM waves into longitudinal plasma oscillations at the resonance region [1, 5, 15–23]. In this paper, we study theoretically similar phenomena occurring in transition anisotropic media with unequal values of the transverse and longitudinal components of the permittivity tensor, using the invariant imbedding theory of wave propagation.

The characteristics of EM wave propagation in anisotropic media, such as uniaxial and biaxial media, have been studied extensively for a long time. More recently, hyperbolic media, where one of the principal values of the permittivity tensor has a different sign from the others and their dispersion surface is a hyperboloid, have attracted great interest from many researchers [24–26]. The strong absorption of EM waves in uniform anisotropic media has been studied recently by several authors, with an emphasis on the role of a hyperbolic dispersion relation [9–14]. In this paper, we consider an inhomogeneous slab of a stratified anisotropic medium, optical axes of which are tilted with respect to the direction of inhomogeneity. We are interested in the case where the tensor components in the longitudinal and/or transverse directions vary from positive to negative values within the medium and the associated mode conversion and resonant absorption phenomena. We examine the different roles of the transverse and longitudinal components on resonant absorption. We also investigate interesting asymmetric transmission and absorption phenomena occurring when the tilt angle is nonzero.

2. Model

We consider an anisotropic dielectric medium, the dielectric permittivity tensor of which is written as

ϵ=(ϵ1000ϵ2000ϵ3)
in the coordinate system (x′, y′, z′). We assume that the primed coordinate system is obtained by rotating the unprimed coordinate system (x, y, z) by an angle ϕ with respect to the y axis, as depicted in Fig. 1. Then the permittivity tensor ϵ in the unprimed coordinate system is related to ϵ′ by
ϵ=RTϵR,
where R is the rotation matrix given by
R=(cosϕ0sinϕ010sinϕ0cosϕ).
The explicit form for ϵ is
ϵ=(ϵ110ϵ130ϵ20ϵ130ϵ33),ϵ11=ϵ1cos2ϕ+ϵ3sin2ϕ,ϵ33=ϵ1sin2ϕ+ϵ3cos2ϕ,ϵ13=(ϵ3ϵ1)sinϕcosϕ.

 figure: Fig. 1

Fig. 1 Schematic view of an anisotropic medium slab of thickness L with the optical axis z′ tilted from the z axis by an angle ϕ. A p wave is incident from the right-hand side with an incident angle θ.

Download Full Size | PDF

The anisotropic medium is stratified along the z axis and all tensor components ϵ1, ϵ2 and ϵ3 are arbitrary functions of z in general. Plane monochromatic EM waves of frequency ω and vacuum wave number k0 (= ω/c) are assumed to propagate in the xz plane. Then the wave equations for s and p waves are completely decoupled. For p waves, the y component of the magnetic field and the x component of the electric field satisfy the differential equation

ddz(HyEx)=(iqϵ13ϵ33ik0ϵ1ϵ3ϵ33ik0(1q2k0 21ϵ33)iqϵ13ϵ33)(HyEx),
where q is the x component of the wave vector. For s waves, the y component of the electric field satisfies
d2Eydz2+(k0 2ϵ2q2)Ey=0.
We can transform Eq. (5) into a second-order differential equation for Hy by eliminating Ex, though the present form is more convenient for applying the invariant imbedding method. When ϕ is equal to zero, the equation for Hy takes a simple form given by
d2Hydz21ϵ1dϵ1dzdHydz+(k0 2ϵ1q2ϵ1ϵ3)Hy=0.

We assume that an inhomogeneous anisotropic medium of thickness L lies in 0 ≤ zL and the waves are incident from a uniform dielectric region (z > L) and transmitted to another uniform dielectric region (z < 0). The incident and transmitted regions are filled with ordinary isotropic dielectric media, where ϵ (=ϵi) is a scalar quantity. In this paper, we are interested only in the resonant absorption phenomenon associated with the mode conversion of transverse EM waves into longitudinal plasma oscillations. In nonmagnetic media, mode conversion can occur only for p waves. Therefore we need to consider only Eq. (5) and the configurations for ϵ1 and ϵ3. The tensor component ϵ2 is completely irrelevant.

In this work, we consider three different configurations for ϵ1 and ϵ3, which are depicted in Fig. 2. Mode conversion is found to occur at the positions where ϵ33 (=ϵ1 sin2 ϕ + ϵ3 cos2 ϕ) is zero and a singularity occurs in Eq. (5). For the purpose of numerical calculations, we introduce extremely small imaginary parts of ϵ1 and ϵ3 to avoid the singularity. Even in the limit where these imaginary parts go to zero, the absorption of the wave energy is finite due to mode conversion and converges to a constant. We have verified that the choice of 10−8 for the imaginary parts yields sufficiently convergent results for the absorptance. As we mentioned already, mode conversion does not occur for s waves. In the special case where the tilt angle ϕ is zero, ϵ33 is equal to ϵ3, and therefore only ϵ3 plays a role in causing resonant absorption and ϵ1 is irrelevant. The vanishing of ϵ1 alone does not cause resonant absorption.

 figure: Fig. 2

Fig. 2 Spatial configurations of ϵ1 and ϵ3 inside the slab of an anisotropic medium of thickness L considered in this paper. A p wave is assumed to be incident from the region where z > L. In numerical calculations, we introduce extremely small imaginary parts of ϵ1 and ϵ3.

Download Full Size | PDF

3. Invariant imbedding method

We consider a p wave of unit magnitude incident obliquely at an incident angle θ on the anisotropic medium. In the invariant imbedding method, we are mainly interested in the reflection and transmission coefficients, r = r (L) and t = t(L), defined by

Hy(z)={eip(Lz)+reip(zL),z>Lteipz,z<0,
where p(=ϵik0cosθ) is the negative z component of the wave vector in the incident and transmitted regions. Using the invariant imbedding method [27–31], we derive exact differential equations satisfied by r and t:
1ipdrdl=2ϵ1ϵ3ϵ33ϵir+12(sec2θϵ1ϵ3ϵ33ϵiϵiϵ33tan2θ)(1+r2),1ipdtdl=(ϵ1ϵ3ϵ33ϵi+ϵ13ϵ33tanθ)t+12(sec2θϵ1ϵ3ϵ33ϵiϵiϵ33tan2θ)(1+r)t,
starting from Eq. (5). These equations are integrated numerically from l = 0 to l = L using the initial conditions, r(0) = 0 and t(0) = 1.

The invariant imbedding method can also be used in calculating the field amplitude Hy(z; L) inside the inhomogeneous medium, which is considered as a function of both z and L and satisfies

1ipHyl=(ϵ1ϵ3ϵ33ϵi+ϵ13ϵ33tanθ)Hy+12(sec2θϵ1ϵ3ϵ33ϵiϵiϵ33tan2θ)(1+r)Hy.
For a given z (0 < z < L), the field amplitude is obtained by integrating this equation from l = z to l = L using the initial condition Hy(z; z) = 1 + r(z).

We notice that when ϕ is nonzero (hence ϵ13 ≠ 0), the equations for the field amplitude and the transmission coefficient are manifestly asymmetric under the sign change of θ due to the ϵ13 tan θ term, while the equation for the reflection coefficient is always symmetric. In the next section, we will show that this can lead to asymmetric transmittance and absorptance under the sign change of θ. Furthermore, we point out that all quantities are symmetric under the simultaneous sign changes of θ and ϕ.

The symmetry of the reflection coefficient under the sign change of θ is a direct consequence of the reciprocity principle, which can be applied to systems with time-reversal symmetry [32]. We note that the reflection coefficients for θ and −θ correspond to two time-reversed processes. In contrast to the reflection coefficient, the transmission coefficients for θ and −θ do not describe time-reversed processes and are generally unequal. On the other hand, the two transmission coefficients for waves incident from the right-hand side of the slab and from the left-hand side of the slab in precisely opposite directions correspond to two time-reversed processes and should be the same. Since the configuration II is obtained by flipping the left and right sides of the configuration I, this will lead to the identity of the transmittances at the same θ for the configurations I and II.

4. Numerical results

4.1. ϕ = 0 case

We first consider the case where the tilt angle ϕ is zero. Then Hy satisfies Eq. (7), which can be transformed further by introducing a new field ψ defined by

Hy=ϵ1ψ
into
ψ+[k02ϵ1q2ϵ1ϵ3+12ϵ1 ϵ134(ϵ1 )2ϵ1 2]ψ=0,
where a prime denotes a differentiation with respect to z. This equation can be written as
ψ+k2[1η(z)]ψ=0,
where k=ϵik0 and η is given by
η=1ϵ1ϵi+ϵ1ϵ3sin2θ121k2ϵ1 ϵ1+341k2(ϵ1 )2ϵ1 2.
Equation (13) has the same form as the Schrödinger equation and the function η plays the role of V(z)/E, where V(z) is the potential and E is the energy of an incident quantum particle. For given configurations of ϵ1 and ϵ3, it is possible to get some insights on the wave propagation by looking at the form of the function η(z).

For the configurations considered in this paper and when ϕ is zero, mode conversion occurs at z = 0.5L, where Re ϵ3 vanishes. In order to have strong resonant absorption via mode conversion, it is necessary for a large fraction of wave energy to arrive at z = 0.5L without substantial reflection by the medium in 0.5 < z/L < 1. Strong reflection can result from the presence of an evanescent region, where η > 1 and the wave number is complex. Therefore, if there exists an evanescent region of substantial size in 0.5 < z/L < 1, we expect mode conversion to be strongly suppressed.

In Fig. 3, we plot the reflectance R (= |r|2), the transmittance T (=|t|2) and the absorptance A (= 1 − RT), which can be equivalently called the mode conversion coefficient, for p waves incident on the slabs described by the configurations I, II and III shown in Fig. 2 versus incident angle, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−8 and L = 5λ. We find that the absorptance is fairly large in a wide range of the incident angle in Case I, while it is substantial only in the small θ region in Case II and in the large θ region in Case III. The maximum values of A are 0.994 at θ = 21.4° in Case I, 0.252 at θ = 6.8° in Case II and 0.822 at θ = 78.6° in Case III. We note that the absorptance in Case II is substantially smaller than those in the other cases.

 figure: Fig. 3

Fig. 3 (a) Absorptance, (b) reflectance and (c) transmittance of p waves incident from the region where z > L versus incident angle for the configurations I, II and III shown in Fig. 2, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−8 and L = 5λ.

Download Full Size | PDF

In Case I, the transverse and longitudinal tensor components, ϵ1 and ϵ3, have different signs in the region 0.5 < z/L < 1, where the slab can be considered as an inhomogeneous hyperbolic medium. More specifically, since ϵ3 is negative, it can be regarded as a type I hyperbolic medium, if ϵ2 = ϵ1. The function η is given by

ηI=sin2θ2(z/L)1.
Since ηI is negative in 0.5 < z/L < 1, no evanescent region occurs there for any incident angle, which results in a large absorption over a wide range of θ, except for θ very close to 0° or 90°. The reflectance increases monotonically from 0 to 1 as θ increases, which leads to a suppression of A for large θ. This suppression results from two factors. First, the mode conversion efficiency decreases as θ increases to 90°, which makes a larger fraction of the wave energy reach beyond the resonance region at z = 0.5L. Second, an evanescent region occurs in 0.5 cos2 θ < z/L < 0.5. The width of this region increases, and therefore the reflectance increases and the transmittance decreases, as θ increases. When θ is close to zero, the effective potential η is very small, since it is proportional to sin2 θ. This makes the slab almost transparent, with T ≈ 1 and A ≈ 0.

In Case II, ϵ1 and ϵ3 have the same signs in 0.5 < z/L < 1, where the slab is an elliptic medium. The function η is given by

ηII=ηI=sin2θ2(z/L)1.
An evanescent region appears in 0.5 < z/L < 0.5(1 + sin2 θ). Since the width of this region increases rapidly as θ increases, total reflection and zero absorption occur for reasonably large values of θ. When θ is close to zero, η is very small, which results in a large transmission and a small absorption, as in Case I. As we have already explained in Sec. 3, the transmittances in Case I and II are identically the same because of the reciprocity principle.

In Case III, ϵ1 and ϵ3 have different signs in 0.5 < z/L < 1, where the slab is a hyperbolic medium. Since ϵ1 is negative, it can be regarded as a type II hyperbolic medium, if ϵ2 = ϵ1. The function η is given by

ηIII=2sin2θ2(z/L)1.
In this case, the wave is evanescent everywhere except in 0.5 < z/L < 0.5(1 + sin2). In order to have nonzero absorption, the width of this region has to be sufficiently large. For θ smaller than about 50°, the slab behaves as a perfect reflector and shows zero absorption. Since the wave is evanescent in the wide region 0 < a/L < 0.5 for any incident angle, the transmittance is always zero.

Summarizing the discussions above, we conclude that mode conversion is much more enhanced in the case where the medium from the incident surface to the resonance region is hyperbolic than that where it is elliptic. A type I hyperbolic medium is more efficient in this regard than a type II hyperbolic medium.

In Fig. 4, we show two examples of the spatial distributions of the electric and magnetic fields inside the inhomogeneous slab when a p wave is incident from the region where z > L. When the tilt angle ϕ is zero, the electric field components are obtained using

Ez=qk0Hyϵ3,Ex=ik0ϵ1dHydz.
We consider the cases where a wave is incident on the slab I of Fig. 2 at θ = 20° and on the slab II of Fig. 2 at θ = 7°, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−3 and L = 5λ. In the former case, R (= 0.0034) is very small and A (= 0.993) is large, while, in the latter, R is 0.268 and A is 0.252. We observe clearly that mode conversion of transverse waves into longitudinal plasma oscillations occurs at z = 0.5L where Re ϵz = 0, as can be seen from the strong enhancement of Ez near z = 0.5L in Fig. 4(b). When the reflectance is not very small in Case II, the incident waves are primarily reflected from the boundary of the evanescent region located at z = 0.5(1 + sin2 θ)L. The wave incident from the right and propagating to the left and that reflected from the evanescent region interfere and the field distribution shows oscillatory behavior in the region 0.5(1 + sin2 θ)L < z < L, as can be clearly seen in the blue curves. If the reflectance is negligible, no oscillatory feature is observed as in the red curves.

 figure: Fig. 4

Fig. 4 Spatial distributions of (a) the magnetic field, (b) the z component of the electric field and (c) the x component of the electric field inside the inhomogeneous anisotropic medium when a p wave is incident from the region where z > L on the slab I of Fig. 2 at θ = 20° and on the slab II of Fig. 2 at θ = 7°, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−3 and L = 5λ.

Download Full Size | PDF

4.2. ϕ ≠ 0 case

Next, we consider the more general case where the tilt angle ϕ is nonzero. In Fig. 5, we show the absorptance, the reflectance and the transmittance versus incident angle for the configuration I in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and ϕ = 0°, 15°, 30°, 45°, 60° and 75°. As we have already mentioned in Sec. 3, all quantities are symmetric under the simultaneous sign changes of θ and ϕ. Therefore, we need to consider only the cases with ϕ > 0. We find that the absorptance is zero for all incident angles, if ϕ > 45°. This can be understood easily from the form of ϵ33 given by

ϵ33=12zLcos2ϕ.
The resonance point zR at which mode conversion occurs is determined by ϵ33 = 0, which gives
zRL=12cos2ϕ.
If ϕ > 45°, this condition cannot be satisfied at any point within the medium.

 figure: Fig. 5

Fig. 5 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration I shown in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45°, (e) 60° and (f) 75°.

Download Full Size | PDF

Another interesting feature is that both A and T are asymmetric under the sign change of θ when 0 < ϕ ≤ 45°, whereas T is symmetric and A is zero when ϕ > 45°. Since A is zero for all θ when ϕ > 45°, the symmetry of T has to follow from the law of energy conservation R +T + A = R +T = 1, considering that R is always symmetric. When ϕ ≤ 45°, the transmittance curve is sharply peaked at θ = −ϕ, with the peak value equal to 1. This corresponds to the situation where the wave is incident precisely along the optical axis z′. When ϕ > 45°, the transmittance curve has two peaks at θ = ±ϕ. This can be understood from the behavior of the coefficient of the (1 + r)2 and (1 + r)t terms in Eq. (9):

sec2θϵ1ϵ3ϵ33ϵiϵiϵ33tan2θ=2(z/L)(cos2θcos2ϕ)ϵ33cos2θ,
which vanishes at θ = ±ϕ. When θ = −ϕ, the inhomogeneous slab is effectively a transparent medium with T = 1 and R = A = 0 for all values of ϕ. In contrast, when θ = ϕ, the slab is transparent with T = 1 and R = A = 0 for ϕ > 45°, whereas it is a perfect absorber with A = 1 and R = T = 0 for 0 < ϕ ≤ 45°.

As ϕ increases from zero to 45°, the reflectance near θ = 0 increases monotonically. For ϕ > 45°, the slab is a perfect reflector at all incident angles except for those in a narrow region around θ = ±ϕ. When ϕ is equal to 90°, the permittivity configuration is the same as that for the ϕ = 0 case with ϵ3 = 1 and ϵ1 = 1 − 2(z/L). The effective potential η for this configuration is given by

η=1+(2zL1)cos2θ+3ζ2[2(z/L)1]2,
where ζ = kL = 10π. Since this function is always larger than 1 in 0.5 < z/L < 1, the wave is evanescent in that wide region, which causes the slab to behave as a total reflector at all incident angles.

In Fig. 6, we show the absorptance, the reflectance and the transmittance versus incident angle for the configuration II shown in Fig. 2, for the same parameter values as in Fig. 5. Similarly to Fig. 5, we find that the absorptance is zero for all incident angles, if ϕ > 45°. The function ϵ33 in the present case is given by

ϵ33=1+2(zL1)cos2ϕ.
The resonance point zR satisfies
zRL=112cos2ϕ.
If ϕ is greater than 45°, this condition cannot be satisfied at any point within the medium, and therefore A is always zero.. The curves for the transmittance, which show one peak at θ = −ϕ if ϕ ≤ 45° and two peaks at θ = ±ϕ if ϕ > 45°, are identical to those in Fig. 5 for all values of ϕ, which is a consequence of the reciprocity principle. Except when the incident angle is near ±ϕ, the slab behaves as a perfect reflector with R = 1. For very small values of ϕ, the absorptance curve shows two small peaks near θ = 0, but for ϕ larger than about 5°, it has one large peak at θ = ϕ with the peak value equal to 1.

 figure: Fig. 6

Fig. 6 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration II in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45°, (e) 60° and (f) 75°.

Download Full Size | PDF

In Fig. 7, we show the absorptance, the reflectance and the transmittance versus incident angle for the configuration III in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and ϕ = 0°, 15°, 30°, 45° and θ > 45°. Once again, we find that the absorptance is zero for all incident angles, if ϕ > 45°. ϵ33 in the present case is given by

ϵ33=2zLcos2ϕ1.
The equation for the resonance point zR is the same as Eq. (19) and cannot be satisfied at any point within the medium if ϕ > 45°, which leads to the vanishing of A. In the present case, the evanescent region is so wide that there is no transmission at any incident angle for all values of ϕ. Since R is always symmetric under the sign change of θ, the absorptance is also symmetric as shown in Fig. 7. When ϕ > 45°, both T and A are zero for all θ, and therefore the slab acts as an omnidirectional perfect reflector. As ϕ increases from zero to 45°, the broad absorption peaks become wider and move to θ = ±45°.

 figure: Fig. 7

Fig. 7 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration III in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45° and (e) ϕ > 45°. In (e), both T and A are zero.

Download Full Size | PDF

Next, we discuss the effects of the thickness of the slab, the mismatch between ϵ1 and the background ϵi and the loss parameter Im ϵ3 on the results presented above. In Fig. 8, we show how the absorptance, reflectance and transmittance curves change as the thickness of the inhomogeneous layer increases, for the configurations I, II and III, when Im ϵ3 = 10−8, ϵi = 1, ϕ = 30° and L = 0.5λ, λ, 5λ and 20λ. We notice a general trend that the peaks and dips become sharper as L/λ increases. Nevertheless, general features of mode conversion phenomena are observable for L as small as 0.5λ.

 figure: Fig. 8

Fig. 8 Absorptance, reflectance and transmittance versus incident angle for (a–c) Case I, (d–f) Case II and (g–i) Case III, when Im ϵ3 = 10−8, ϵi = 1, ϕ = 30° and L = 0.5λ (black), λ (red), 5λ (green) and 20λ (blue).

Download Full Size | PDF

In Fig. 9, we show how the mismatch between ϵi and ϵ1 influences the absorptance. The configurations are the same as I and II in Fig. 2, except that ϵ1 is generally not equal to 1. The expression for ϵ33 given in Eq. (4) depends on ϵ1. For ϕ = 30° chosen in Fig. 9, we can show that the region where Re ϵ33 = 0 can exist only when 0 < ϵ1 < 3 for both cases. Furthermore, by examining the left-hand side of Eq. (21) for ϵ1 ≠ 1, we can show that the absolute value of the incident angle at which A = 0 in Case I and A = 1 in Case II increases (decreases) as ϵ1 increases (decreases) from 1. Though the detailed shape of the absorptance curve depends on ϵ1, mode conversion and resonant absorption behaviors are qualitatively similar to the matched case.

 figure: Fig. 9

Fig. 9 Absorptance versus incident angle for the configurations similar to (a) I and (b) II in Fig. 2, when Im ϵ3 = 10−8, L = 5λ, ϕ = 30°, ϵi = 1 and ϵ1 = 1 (black), 2 (red), 2.9 (green) and 0.5 (blue).

Download Full Size | PDF

In Fig. 10, we consider the influence of the loss parameter, Im ϵ3, on mode conversion phenomena. Resonant absorption due to mode conversion is not due to damping, but rather to the conversion of transverse wave modes into longitudinal modes. Therefore the absorptance in this case converges to a finite value even in the limit where Im ϵ3 goes to zero. In realistic cases, however, there is always finite damping, which will enhance the overall absorption. In Fig. 10, we consider a thin slab of thickness L = λ for the configurations I, II and III in Fig. 2 and vary Im ϵ3. We notice that the results for Im ϵ3 = 0.001 are almost the same as those obtained when Im ϵ3 → 0+. The influence of the loss parameter appears to be stronger in Case II than in the other cases. We find that the value of 0.01 is sufficiently good and mode conversion phenomena are not dominated by damping if Im ϵ3 is smaller than about 0.01.

 figure: Fig. 10

Fig. 10 Absorptance versus incident angle for (a) Case I, (b) Case II and (c) Case III in Fig. 2, when L = λ, ϕ = 30°, ϵi = 1 and Im ϵ3 = 10−8 (black), 0.001 (red), 0.01 (green) and 0.1 (blue).

Download Full Size | PDF

Finally, we comment on the experimental feasibility of the results presented in this paper. Even though we have chosen rather specific types of linear configurations for the purpose of simplicity, their detailed shapes are not essential for the occurrence of mode conversion and resonant absorption phenomena. What is essential is that there has to a region where the real part of ϵ33 in the tilted case and ϵ3 in the non-tilted case vanishes. Transition anisotropic media can be fabricated using a scheme similar to that for hyperbolic media such as metal-dielectric multilayers or metallic wire arrays [24, 25]. In case of metal-dielectric multilayers, one can use an effective medium description such that

ϵ3=fϵm+(1f)ϵd,1ϵ1=fϵm+1fϵd,
where ϵm and ϵd are the dielectric permittivities of the metal and dielectric layers and f is the metal fraction. In Fig. 11, we show the absorptance versus incident angle curve for the configurations similar to I and II in Fig. 2, when L = λ, ϕ = 0° and ϵi = 1. The configurations correspond to graded metal-dielectric multilayers made of Silver with ϵm = −15.24 + 0.4i at λ = 587.6 nm and MgF2 with ϵd = 1.9. An effective medium description where the metal fraction f varies linearly as f = 0.1 + 0.2(z/L) in (a) and as f = 0.3 − 0.2 (z/L) in (b) has been used. In the experimental fabrication of these systems, we suggest to fabricate a graded multilayer with 20 periods, where the thickness of each period is 40 nm and the metal fraction is varied from 0.1 to 0.29 with a step of 0.01. We notice that the basic features of mode conversion phenomena are clearly observed in Fig. 11. In Fig. 11(a), the obtained result is almost the same as in the hypothetical case where the metallic damping is ignored.

 figure: Fig. 11

Fig. 11 Absorptance versus incident angle for the configurations similar to (a) I and (b) II in Fig. 2, when L = λ, ϕ = 0° and ϵi = 1. The configurations correspond to graded metal-dielectric multilayers made of Silver with ϵm = −15.24 + 0.4i at λ = 587.6 nm and MgF2 with ϵd = 1.9. An effective medium description where the metal fraction f varies linearly as f = 0.1 + 0.2 (z/L) in (a) and as f = 0.3 − 0.2 (z/L) in (b) has been used. The results for Im ϵm = 0.4 are compared with those obtained in the hypothetical cases with Im ϵm = 10−8.

Download Full Size | PDF

5. Conclusion

In this paper, we have studied theoretically the mode conversion and resonant absorption phenomena occurring in an inhomogeneous anisotropic medium. We have considered the more general asymmetric case, where the optical axes of the medium are tilted with respect to the direction of stratification. When the tilt angle is zero, we have found that mode conversion occurs if the longitudinal component of the permittivity tensor varies from positive to negative values within the medium, while the transverse component is irrelevant in the mode conversion process. When the tilt angle is nonzero, we have found that the curves for the transmittance and the absorptance can be asymmetric under the sign change of the incident angle, while the reflectance curve is always symmetric. We have calculated the reflectance, the transmittance and the absorptance for three configurations of the permittivity tensor, using the invariant imbedding method, and found that resonant absorption is much more enhanced in the case where the medium from the incident surface to the resonance region is hyperbolic than that where it is elliptic. We expect that the asymmetric transmission and absorption phenomena discussed in this paper will find useful applications in photonic devices.

Funding

National Research Foundation of Korea Grant funded by the Korean Government (NRF-2015R1A2A2A01003494).

References and links

1. K. Kim, D.-H. Lee, and H. Lim, “Resonant absorption and mode conversion in a transition layer between positive-index and negative-index media,” Opt. Express 16(22), 18505–18513 (2008). [CrossRef]   [PubMed]  

2. I. Mozjerin, E. A. Gibson, E. P. Furlani, I. R. Gabitov, and N. M. Litchinitser, “Electromagnetic enhancement in lossy optical transition metamaterials,” Opt. Lett. 35(19), 3240–3242 (2010). [CrossRef]   [PubMed]  

3. E. A. Gibson, M. Pennybacker, A. I. Maimistov, I. R. Gabitov, and N. M. Litchinitser, “Resonant absorption in transition metamaterials: parametric study,” J. Opt. 13(2), 024013 (2011). [CrossRef]  

4. J. Sun, X. Liu, J. Zhou, Z. Kudyshev, and N. M. Litchinitser, “Experimental demonstration of anomalous field enhancement in all-dielectric transition magnetic metamaterials,” Sci. Rep. 5, 16154 (2015). [CrossRef]   [PubMed]  

5. S. Kim and K. Kim, “Resonant absorption and amplification of circularly-polarized waves in inhomogeneous chiral media,” Opt. Express 24(2), 1794–1803 (2016). [CrossRef]   [PubMed]  

6. Y. Jin, S. Xiao, N. A. Mortensen, and S. He, “Arbitrarily thin metamaterial structure for perfect absorption and giant magnification,” Opt. Express 19(12), 11114–11119 (2011). [CrossRef]   [PubMed]  

7. J. Yoon, M. Zhou, M. A. Badsha, T. Y. Kim, Y. C. Jun, and C. K. Hwangbo, “Broadband epsilon-near-zero perfect absorption in the near-infrared,” Sci. Rep. 5, 12788 (2015). [CrossRef]   [PubMed]  

8. M. Lobet, B. Majerus, L. Henrard, and P. Lambin, “Perfect electromagnetic absorption using graphene and epsilon-near-zero metamaterials,” Phys. Rev. B 93(23), 235424 (2016). [CrossRef]  

9. S. Feng and K. Halterman, “Coherent perfect absorption in epsilon-near-zero metamaterials,” Phys. Rev. B 86(16), 165103 (2012). [CrossRef]  

10. I. S. Nefedov, C. A. Valagiannopoulos, S. M. Hashemi, and E. I. Nefedov, “Total absorption in asymmetric hyperbolic media,” Sci. Rep. 3, 2662 (2013). [CrossRef]   [PubMed]  

11. S. Zhong, Y. Ma, and S. He, “Perfect absorption in ultrathin anisotropic -near-zero metamaterials,” Appl. Phys. Lett. 105(2), 023504 (2014). [CrossRef]  

12. K. Halterman and J. M. Elson, “Near-perfect absorption in epsilon-near-zero structures with hyperbolic dispersion,” Opt. Express 22(6), 7337–7348 (2014). [CrossRef]   [PubMed]  

13. J. Linder and K. Halterman, “Graphene-based extremely wideangle tunable metamaterial absorber,” Sci. Rep. 6, 31225 (2016). [CrossRef]  

14. K. V. Sreekanth, M. ElKabbash, Y. Alapan, A. R. Rashed, U. A. Gurkan, and G. Strangi, “A multiband perfect absorber based on hyperbolic metamaterials,” Sci. Rep. 6, 26272 (2016). [CrossRef]   [PubMed]  

15. D. G. Swanson, Theory of Mode Conversion and Tunneling in Inhomogeneous Plasmas (Wiley, 1998).

16. D. E. Hinkel-Lipsker, B. D. Fried, and G. J. Morales, “Analytic expressions for mode conversion in a plasma with a linear density profile,” Phys. Fluids B 4(3), 559–575 (1992). [CrossRef]  

17. E. Mjølhus, “On linear conversion in a magnetized plasma,” Radio Sci. 25(6), 1321–1339 (1990). [CrossRef]  

18. K. Kim and D.-H. Lee, “Invariant imbedding theory of mode conversion in inhomogeneous plasmas. I. Exact calculation of the mode conversion coefficient in cold, unmagnetized plasmas,” Phys. Plasmas 12(6), 062101 (2005). [CrossRef]  

19. K. Kim and D.-H. Lee, “Invariant imbedding theory of mode conversion in inhomogeneous plasmas. II. Mode conversion in cold, magnetized plasmas with perpendicular inhomogeneity,” Phys. Plasmas 13(4), 042103 (2006). [CrossRef]  

20. D. J. Yu, K. Kim, and D.-H. Lee, “Resonant enhancement of mode conversion in unmagnetized plasmas due to a periodic density modulation superimposed on a linear electron density profile,” Phys. Plasmas 17(10), 102110 (2010). [CrossRef]  

21. D. J. Yu, K. Kim, and D.-H. Lee, “Temperature dependence of mode conversion in warm, unmagnetized plasmas with a linear density profile,” Phys. Plasmas 20(6), 062109 (2013). [CrossRef]  

22. D. J. Yu and K. Kim, “Effects of a random spatial variation of the plasma density on the mode conversion in cold, unmagnetized, and stratified plasmas,” Phys. Plasmas 20(12), 122104 (2013). [CrossRef]  

23. D. J. Yu and K. Kim, “Broadband wide-angle absorption enhancement due to mode conversion in cold unmagnetized plasmas with periodic density variations,” Phys. Plasmas 23(3), 032112 (2016). [CrossRef]  

24. A. Poddubny, I. Iorsh, P. Belov, and Y. Kivshar, “Hyperbolic metamaterials,” Nat. Photonics 7(12), 948–957 (2013). [CrossRef]  

25. L. Ferrari, C. Wu, D. Lepage, X. Zhang, and Z. Liu, “Hyperbolic metamaterials and their applications,” Prog. Quantum Electron. 40, 1–40 (2015). [CrossRef]  

26. G. Pawlik, K. Tarnowski, W. Walasik, A. C. Mitus, and I. C. Khoo, “Liquid crystal hyperbolic metamaterial for wide-angle negative-positive refraction and reflection,” Opt. Lett. 39(7), 1744–1747 (2014). [CrossRef]   [PubMed]  

27. V. I. Klyatskin, “The imbedding method in statistical boundary-value wave problems,” Prog. Opt. 33, 1–127 (1994). [CrossRef]  

28. K. Kim, H. Lim, and D.-H. Lee, “Invariant imbedding equations for electromagnetic waves in stratified magnetic media: Applications to one-dimensional photonic crystals,” J. Korean Phys. Soc. 39(6), L956–L960 (2001).

29. K. Kim, D.-H. Lee, and H. Lim, “Theory of the propagation of coupled waves in arbitrarily inhomogeneous stratified media,” Europhys. Lett. 69(2), 207–213 (2005). [CrossRef]  

30. K. Kim, D. K. Phung, F. Rotermund, and H. Lim, “Propagation of electromagnetic waves in stratified media with nonlinearity in both dielectric and magnetic responses,” Opt. Express 16(2), 1150–1164 (2008). [CrossRef]   [PubMed]  

31. S. Kim and K. Kim, “Invariant imbedding theory of wave propagation in arbitrarily inhomogeneous stratified bi-isotropic media,” J. Opt. 18(6), 065605 (2016). [CrossRef]  

32. R. J. Potton, “Reciprocity in optics,” Rep. Prog. Phys. 67(5), 717–754 (2004). [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 (11)

Fig. 1
Fig. 1 Schematic view of an anisotropic medium slab of thickness L with the optical axis z′ tilted from the z axis by an angle ϕ. A p wave is incident from the right-hand side with an incident angle θ.
Fig. 2
Fig. 2 Spatial configurations of ϵ1 and ϵ3 inside the slab of an anisotropic medium of thickness L considered in this paper. A p wave is assumed to be incident from the region where z > L. In numerical calculations, we introduce extremely small imaginary parts of ϵ1 and ϵ3.
Fig. 3
Fig. 3 (a) Absorptance, (b) reflectance and (c) transmittance of p waves incident from the region where z > L versus incident angle for the configurations I, II and III shown in Fig. 2, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−8 and L = 5λ.
Fig. 4
Fig. 4 Spatial distributions of (a) the magnetic field, (b) the z component of the electric field and (c) the x component of the electric field inside the inhomogeneous anisotropic medium when a p wave is incident from the region where z > L on the slab I of Fig. 2 at θ = 20° and on the slab II of Fig. 2 at θ = 7°, when ϕ = 0, ϵi = 1, Im ϵ3 = 10−3 and L = 5λ.
Fig. 5
Fig. 5 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration I shown in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45°, (e) 60° and (f) 75°.
Fig. 6
Fig. 6 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration II in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45°, (e) 60° and (f) 75°.
Fig. 7
Fig. 7 Absorptance (red), reflectance (blue) and transmittance (black) of p waves incident from the region where z > L versus incident angle for the configuration III in Fig. 2, when ϵi = 1, Im ϵ3 = 10−8, L = 5λ and (a) ϕ = 0°, (b) 15°, (c) 30°, (d) 45° and (e) ϕ > 45°. In (e), both T and A are zero.
Fig. 8
Fig. 8 Absorptance, reflectance and transmittance versus incident angle for (a–c) Case I, (d–f) Case II and (g–i) Case III, when Im ϵ3 = 10−8, ϵi = 1, ϕ = 30° and L = 0.5λ (black), λ (red), 5λ (green) and 20λ (blue).
Fig. 9
Fig. 9 Absorptance versus incident angle for the configurations similar to (a) I and (b) II in Fig. 2, when Im ϵ3 = 10−8, L = 5λ, ϕ = 30°, ϵi = 1 and ϵ1 = 1 (black), 2 (red), 2.9 (green) and 0.5 (blue).
Fig. 10
Fig. 10 Absorptance versus incident angle for (a) Case I, (b) Case II and (c) Case III in Fig. 2, when L = λ, ϕ = 30°, ϵi = 1 and Im ϵ3 = 10−8 (black), 0.001 (red), 0.01 (green) and 0.1 (blue).
Fig. 11
Fig. 11 Absorptance versus incident angle for the configurations similar to (a) I and (b) II in Fig. 2, when L = λ, ϕ = 0° and ϵi = 1. The configurations correspond to graded metal-dielectric multilayers made of Silver with ϵm = −15.24 + 0.4i at λ = 587.6 nm and MgF2 with ϵd = 1.9. An effective medium description where the metal fraction f varies linearly as f = 0.1 + 0.2 (z/L) in (a) and as f = 0.3 − 0.2 (z/L) in (b) has been used. The results for Im ϵm = 0.4 are compared with those obtained in the hypothetical cases with Im ϵm = 10−8.

Equations (26)

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

ϵ = ( ϵ 1 0 0 0 ϵ 2 0 0 0 ϵ 3 )
ϵ = R T ϵ R ,
R = ( cos ϕ 0 sin ϕ 0 1 0 sin ϕ 0 cos ϕ ) .
ϵ = ( ϵ 11 0 ϵ 13 0 ϵ 2 0 ϵ 13 0 ϵ 33 ) , ϵ 11 = ϵ 1 cos 2 ϕ + ϵ 3 sin 2 ϕ , ϵ 33 = ϵ 1 sin 2 ϕ + ϵ 3 cos 2 ϕ , ϵ 13 = ( ϵ 3 ϵ 1 ) sin ϕ cos ϕ .
d d z ( H y E x ) = ( i q ϵ 13 ϵ 33 i k 0 ϵ 1 ϵ 3 ϵ 33 i k 0 ( 1 q 2 k 0   2 1 ϵ 33 ) i q ϵ 13 ϵ 33 ) ( H y E x ) ,
d 2 E y d z 2 + ( k 0   2 ϵ 2 q 2 ) E y = 0 .
d 2 H y d z 2 1 ϵ 1 d ϵ 1 d z d H y d z + ( k 0   2 ϵ 1 q 2 ϵ 1 ϵ 3 ) H y = 0 .
H y ( z ) = { e i p ( L z ) + r e i p ( z L ) , z > L t e i p z , z < 0 ,
1 i p d r d l = 2 ϵ 1 ϵ 3 ϵ 33 ϵ i r + 1 2 ( sec 2 θ ϵ 1 ϵ 3 ϵ 33 ϵ i ϵ i ϵ 33 tan 2 θ ) ( 1 + r 2 ) , 1 i p d t d l = ( ϵ 1 ϵ 3 ϵ 33 ϵ i + ϵ 13 ϵ 33 tan θ ) t + 1 2 ( sec 2 θ ϵ 1 ϵ 3 ϵ 33 ϵ i ϵ i ϵ 33 tan 2 θ ) ( 1 + r ) t ,
1 i p H y l = ( ϵ 1 ϵ 3 ϵ 33 ϵ i + ϵ 13 ϵ 33 tan θ ) H y + 1 2 ( sec 2 θ ϵ 1 ϵ 3 ϵ 33 ϵ i ϵ i ϵ 33 tan 2 θ ) ( 1 + r ) H y .
H y = ϵ 1 ψ
ψ + [ k 0 2 ϵ 1 q 2 ϵ 1 ϵ 3 + 1 2 ϵ 1   ϵ 1 3 4 ( ϵ 1   ) 2 ϵ 1   2 ] ψ = 0 ,
ψ + k 2 [ 1 η ( z ) ] ψ = 0 ,
η = 1 ϵ 1 ϵ i + ϵ 1 ϵ 3 sin 2 θ 1 2 1 k 2 ϵ 1   ϵ 1 + 3 4 1 k 2 ( ϵ 1   ) 2 ϵ 1   2 .
η I = sin 2 θ 2 ( z / L ) 1 .
η II = η I = sin 2 θ 2 ( z / L ) 1 .
η III = 2 sin 2 θ 2 ( z / L ) 1 .
E z = q k 0 H y ϵ 3 , E x = i k 0 ϵ 1 d H y d z .
ϵ 33 = 1 2 z L cos 2 ϕ .
z R L = 1 2 cos 2 ϕ .
sec 2 θ ϵ 1 ϵ 3 ϵ 33 ϵ i ϵ i ϵ 33 tan 2 θ = 2 ( z / L ) ( cos 2 θ cos 2 ϕ ) ϵ 33 cos 2 θ ,
η = 1 + ( 2 z L 1 ) cos 2 θ + 3 ζ 2 [ 2 ( z / L ) 1 ] 2 ,
ϵ 33 = 1 + 2 ( z L 1 ) cos 2 ϕ .
z R L = 1 1 2 cos 2 ϕ .
ϵ 33 = 2 z L cos 2 ϕ 1 .
ϵ 3 = f ϵ m + ( 1 f ) ϵ d , 1 ϵ 1 = f ϵ m + 1 f ϵ d ,
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.