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

Extraordinary high- and low-momentum lossless plasmonic modes in one-dimensional metamaterials

Open Access Open Access

Abstract

The implementation of lossless plasmonic modes with very high or very low wave vectors, which are associated with extraordinary large or small nonlocal effective permittivity, could bring a breakthrough in the development of high-performance plasmonic devices. We consider the possibility of obtaining volume plasmon-polariton modes with an extraordinary large or negligibly small longitudinal momentum and nearly zero loss in a 1d lattice consisting of alternating metal and gain semiconductor layers. Solving the dispersion equation, we show that under the condition of the full loss compensation, two traveling plasmonic TM modes can exist in the lattice. One of the modes is capable to provide extraordinary high momentum value, while the other can provide both extraordinary high and low (nearly zero) momentum. On the other hand, only one traveling TE mode can occur, which is capable of providing extraordinary low momentum. The use of such modes opens up new ways to design optical properties and to control light on the nanometer scale.

© 2017 Optical Society of America

1. Introduction

High-momentum (high-k) modes are particularly attractive for a range of technologically important applications, such as plasmonic sensors [1], superresolution imaging techniques [2–4], and others [5], where the low phase velocity and/or our ability to manipulate light propagation at a deep subwavelength scale is of critical importance [6–9]. Low-momentum (low-k) modes, which are characterized by the superluminal phase velocity, are also of interest for many applications, such as metatronics applications [10] or the electromagnetic field concentrators, wavefront converters, and cloaking [11]. In addition, a strong electric field, because of the boundary condition for the normal component of the displacement, gives rise to strong optical nonlinearities which, in turn, can result in a number of interesting effects [12,13]. In the case of macroscopically homogeneous media, for which a homogenization procedure can be applied, high-k modes can be described in terms of a very high permittivity, and low-k modes - in terms of a very low permittivity (epsilon-near-zero, or ENZ).

Low loss and hence long-range plasmonic modes are of relevance for a number of integrated optoelectronic devices. The imaginary part of the mode wave vector k close to zero is highly desirable. Because the energy dissipation is unavoidable in realistic plasmonic structures, the use of gain media like dyes or semiconductor quantum structures, which allow to compensate intrinsic loss of metals and semiconductor plasmonic materials, has received acceptance in recent years. Such metamaterials (MMs) exhibit effects similar to electromagnetically induced transparency [14–17].

Yet, in spite of a number of theoretical and experimental studies addressing the issue of loss compensation in various plasmonic structures (see, e.g., [18–25]), there are many challenging opportunities and problems to be solved. Even in the simplest (1d) case the problems of combining full loss compensation with the high-k and low-k regimes remain unconsidered. Volume plasmon modes with a very high momentum and corresponding giant permittivity can be achieved in metal/dye-doped dielectric lattices, but such an approach involves the use of very thin dielectric layers [9]. On the other hand, ENZ regime in the visible range has been realized for metal/dye-doped polymer nanolaminates, but the achieved imaginary part of the longitudinal component of the effective permittivity tensor could not be smaller than −0.04 [21]. The use of dyes is limited for such applications because of their bleaching, low damage threshold, and difficulties associated with current-injection pumping. Due to the limitations on the gain coefficient g for dyes, as g ∝ Imϵ/n, the negative imaginary part of their permittivity cannot be very high, that makes the problem of compensation of the plasmonic material loss difficult. Meanwhile, it can reach about −0.1 – −0.3 for semiconductor structures [23]. Among those, in the telecommunication range, transparent conducting oxides (TCO) and especially InGaAsP are worth to be noticed [26].

The aim of this paper is to examine the feasibility of the full loss compensation for high-k and low-k plasmonic modes for 1d metal/semiconductor MMs. In Sec. 2 we first describe possible solutions of the dispersion equation imposing the condition of the full loss compensation. Then, in Sec. 3 and Sec. 4 we analyze the high-k and low-k solutions, respectively. Sec. 5 contains our conclusions.

2. Loss compensation in layered structures in terms of Rytov’s equation

A schematic of the structure (1d infinite bi-layer MM) is shown in Fig. 1. While there exist a number of homogenization techniques for such MMs [27], we adopt a rigorous treatment of electromagnetic wave propagation in 1d lattices using Rytov’s dispersion equation [28]. Although this equation has been considered in many papers (see, e.g., [29–32]), finding its solutions is not a trivial task.

 figure: Fig. 1

Fig. 1 The unit cell of the structure under consideration.

Download Full Size | PDF

In the case of in-plane propagation (k = 0), as shown in Fig. 1, only one diagonal component of the effective permittivity tensor describes the nonlocal response [24]:

ϵeff(k)ϵeff(k)=k2/k02,
where k0 = ω/c and kk is the in-plane eigenwave number which should satisfy Rytov’s dispersion equation [28]
F=cos(k1d1)cos(k2d2)γsin(k1d1)sin(k2d2)1=0,
with γ=12(ϵ1k2ϵ2k1+ϵ2k1ϵ1k2) and γ=12(k2k1+k1k2) for the TM- and TE-polarization, respectively, k1,2=k02ϵ1,2k2, and d1,2 the layers’ thickness. As was shown, the dispersion equation (2) can be split into two equations, namely
k1ϵ1tank1d12+k2ϵ2tank2d22=0,
and
k1ϵ1tank2d22+k2ϵ2tank1d12=0,
for the TM case, and
k1tank1d12+k2tank2d22=0,
and
k1tank2d22+k2tank1d12=0,
for the TE case [28,29]. Eqs. (3), (5), and (4), (6) describe respectively the so-called even and odd plasmonic modes, for which the transverse quasimicroscopic electric and magnetic fields, E and H, are even and odd functions with respect to the middle of the layers.

In the following, we impose the condition of Im ϵeff = 0, that corresponds to the case of the full loss compensation with either purely real or purely imaginary k. Because Eq. (2) contains many parameters, it is difficult to analyze the entire parameter space. To simplify our consideration, in the subsequent development we assume that the permittivity of the plasmonic component is a constant and fix it at ϵ1 = −122.19 + i3.115, that corresponds to silver at λ = 1.5 µm [25]. To extrapolate our results to other spectral ranges, we note important scaling properties of Eq. (2). Let ϵ1=rϵ1 and k0=sk0 with real r and s. Then, as is easy to see from Eqs. (1) and (2), the following relation holds for ϵeff:

ϵeff(k0d1,k0d2,ϵ1,ϵ2)=rs2ϵeff(rsk0d1,rsk0d2,ϵ1,ϵ2/r).

For a gain medium, we take ϵ2=12+iϵ2, that approximatelly corresponds to InGaAsP multiple quantum wells at the same wavelength, neglecting the free carrier dependence of Reϵ2ϵ2 and assuming that Imϵ2ϵ2 is a negative parameter which depends on the gain coefficient g as ϵ2=gn2/k0, where n2=Reϵ2. For the parameters chosen, the conditions ϵ1|ϵ1| and ϵ2<|Reϵ1| are met. We note that recently light emission from Ag/InGaAsP multiple quantum well structures in the near-IR has been demonstrated experimentally [33,34].

Assuming that k0d ≪ 1 (weak nonlocality), Eq. (2) can be expanded into a power series that yields an approximate analytical solution for the effective permittivity with nonlocal corrections [35]. In the general case, the solutions of Eq. (2) can be represented as a family of curves in the 3d space of (ϵeff;ϵ2;f), where f is the filling factor of the gain medium defined as f = d2/d with d = d1 + d2.

As a first step, we consider the behavior of the function F(f, ϵeff) at fixed ϵ2. In Figs. 2 and 3, we show the function |F(f, ϵeff)| in the logarithmic scale at ϵ2=0.2 and d = 100 nm for the TM and TE polarizations, respectively.

 figure: Fig. 2

Fig. 2 The function |F(f, ϵeff)| (logarithmic scale) at Im(ϵ2)=−0.2 and d = 100 nm for the TM polarization.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 The function |F(f, ϵeff)| (logarithmic scale) at Im(ϵ2)=−0.2 and d = 100 nm for the TE polarization.

Download Full Size | PDF

The solutions of Eq. (2), which correspond to the condition of ln |F| → −∞, lie within fractures which form a system of intersecting branches. Although ϵ1, ϵ2, and d are fixed in these calculations, such a behavior of the function F(f, ϵeff) is typical. For a finite range of values of ϵeff, if it includes its negative values, the number of solution branches depends on the lattice period d.

While an infinite number of solution branches can exist, only two of them, which correspond to positive values of ϵeff, are of interest for us. They are shown by the arrows in Figs. 2 and 3 as A and B. Branches A and B correspond to even and odd modes, respectively. In what follows, we define ϵeff(A) and ϵeff(B) as ϵeff(A)=(kA/ko)2 and ϵeff(B)=(kB/ko)2, where kA and kB are the solutions of Eq. (2) for branches A and B, respectively.

For the TM polarization, branch A gives the traveling-wave solutions of Eq. (3), because for it k is always real and hence ϵeff is positive. Indeed, this branch starts at ϵeff = ϵ2, when f = 1 and Imϵ2(0)=0, and then monotonically ϵeff → ∞ when f approaches zero [9] (here and below the superscript (0) denotes the value of the permittivity of the gain medium which satisfies the dispersion equation (2) under the condition of Imϵeff = 0).

The numerical solutions of Eq. (3) (branch A) in the form of the dependences ϵeff (f) and corresponding Imϵ2(0)(f) at different values of d, including the local approximation (the quasistatic limit d → 0), are presented in Fig. 4. In the local approximation, the effective permittivity takes for this mode the well-known form

ϵeff=[(1f)/ϵ1+f/ϵ2]1.

 figure: Fig. 4

Fig. 4 The dependence of ϵeff(A) (green curves) and corresponding ϵ2(0) (blue curves) vs f at different values of d for the even TM mode.

Download Full Size | PDF

As can be seen, nonlocality reduces the effective permittivity. At the same time, it can result in increasing as well as decreasing the gain coefficient needed to achieve the transparency condition (Imϵeff = 0) as compared to the case of local approximation. It is interesting to note that in the local approximation, there is a cut-off at f=fc=ϵ2/(ϵ2ϵ1) such that ϵeff → +∞ if ffc+ and ϵeff → −∞ if ffc. However, no cut-off takes place when nonlocality is taken into account (see Appendix).

Branch B arises as the solutions of Eq. (4). The numerical solutions of this equation are presented in Fig. 5, where we do not show the region of large negative values of ϵeff(B). This branch includes both the traveling (ϵeff(B)>0) and evanescent (ϵeff(B)<0) mode solutions. For this branch, ϵeff(B)= in the limit of f → 1, then it reduces, intersects the line of ϵeff(B)=0, and becomes negative as f becomes smaller. One can show that the solutions for this branch can exist only when f lies within an interval of [fmin, 1]. As d → 0, this interval shrinks, and in the quasistatic limit the solutions disappear. This means that branch B cannot be described within the framework of the local approximation. Indeed, in the local approximation, Eq. (8) cannot provide ϵeff = 0, except for the trivial case when ϵ1 = 0 or ϵ2 = 0. At the same time, taking ϵeff = 0 in Eq. (4), one has the equation

ϵ1tan(k0d1ϵ1/2)+ϵ2tan(k0d2ϵ2)=0,
which, as will be seen below, has nontrivial solutions.

 figure: Fig. 5

Fig. 5 The dependence of ϵeff(B) (green curves) and corresponding ϵ2(0) (blue curves) vs f at different values of d for the odd TM mode.

Download Full Size | PDF

For the TE-polarization, there is only one branch (A), which can give the traveling wave solution. As is seen from Fig. 3, this branch includes also evanescent mode solution, which can occur at smaller values of the parameter f.

3. High-momentum modes

As can be seen from Fig. 5, when dealing with branch B, to obtain ϵeff higher than 300, Imϵ2(0) must be about −0.3 or lower, and this value only slightly depends on the lattice period d. This requires very high gain coefficient, that can involve technical difficulties. At the same time, for branch A, ϵeff ≈ 300 can be achieved at ϵ20.25 (see Fig. 4). Thus, the use of branch A is more preferable than the use of branch B, if we want to reach the very high permittivity at reasonable values of the gain coefficient.

High-k modes (branch A), which correspond to the extraordinary high effective permittivity ϵeff, were addressed in [9]. Here, we consider this issue in more detail. First, we elucidate the relationship between ϵeff(A), Imϵ2, and the background permittivity, ϵ2, at fixed d and different values of d2. In Fig. 6, we show the dependences of Imϵ2(0)(ϵ2) and ϵeff(A)(ϵ2) at d = 90 nm for several values of d2. The presented results evidence that a decrease in the background permittivity ϵ2 leads to lowering both the transparency threshold Imϵ2(0), that can be considered as a useful effect, and the effective permittivity ϵeff, that is detrimental from the practical point of view. In a similar fashion, a decrease in the thickness d2, which corresponds to reducing f, leads to lowering both ϵ2(0) and ϵeff(A).

 figure: Fig. 6

Fig. 6 The dependence of ϵeff(A) (green curves) and corresponding Imϵ2(0) (blue curves) vs Reϵ2 for the even TM mode at d = 90 nm.

Download Full Size | PDF

Within the framework of our approach, as large as desired in-plane momentum can be obtained. In fact, however, it is bounded above by gain saturation, as well as due to the intrinsic nonlocality of the dielectric responses and crystallinity of the MM constituents [9].

Finally, it should be noted that Eq. (2) at high ϵeff can be approximately solved with respect to f. For this case, its closed-form analytical solutions for both branches, A and B, are given in Appendix.

4. Low-momentum modes

Low-momentum plasmonic modes can occur for the TM polarization (branch B) as well as for the TE polarization (branch A). However, in the limit of ϵeff → 0, Eqs. (4) and (5) (which describe the corresponding modes) coincide, yielding Eq. (9).

To determine the conditions for the existence of zero-momentum modes, we numerically solve Eq. (9) for different values of the background permittivity ϵ2. The results are shown in Fig. 7. They evidence that a decrease in Reϵ2 lowers the transparency threshold Imϵ2(0), while a decrease in the lattice period d elevates it. Such a low transparency threshold as Imϵ2(0)0.117 can be achieved at ϵ2=12 and f ≈ 0.8026, d = 200 nm (k0d = 0.8378), that is feasible with today’s technology. It is noteworthy that as the lattice period d becomes large, the value of Imϵ2(0) only slightly depends on the background permittivity ϵ2.

 figure: Fig. 7

Fig. 7 The dependence Imϵ2(0)(k0d) (blue curves) and corresponding f(k0d) (green curves) at which ϵeff = 0 for different values of Reϵ2.

Download Full Size | PDF

Although Eq. (9) is valid for both the TM and TE polarizations, the behavior of the solutions of Eq. (2) in the vicinity of the point of ϵeff = 0 is different. In particular, contrasting the TE polarization case, the TM case soulions of Eq. (2) are highly sensitive with respect to small changes in the material and geometrical parameters. For example, at d = 100 nm, one obtains ϵeff(A,B)=0 as ϵ2=0.2526 and f = 0.8952. Meanwhile, ϵeff(B)=1 as ϵ2=0.2562, f = 0.8965 (the TM polarization) and ϵeff(A)=1 as ϵ2=0.2358 f = 0.905 (the TE polarization).

5. Discussion

The existence of two solution branches of Rytov’s dispersion equation for the TM polarization does not come as a surprise. In photonic crystals, it can manifest itself in the splitting of the incidence beam due to the appearence of an additional wave, as was shown for the metallic nanorod MMs [36,37]. However, the additional wave is characterized by a non-vanishing component of the quasi-microscopic electric field in the longitudinal direction E, which has a symmetrical distribution, in contrast to the antisymmetrical distribution of its transverse component [28]. This means that the odd TM mode cannot be excited by a normally incident transverse plane wave. In other words, only one (even) mode can be excited at normal light incidence. The coexistence of the two modes and mode mixing become possible for obliquely incident light only. This problem calls for further consideration and is out of the scope of the present work.

Since the odd TM polarized mode does not couple to the normally incident transverse wave, this makes it impossible to get ϵeff = 0 in this regime. For example, ϵeff(B)=0 at d = 200 nm, when ϵ2(0)12i0.1167.and f ≈ 0.8016. At the same time, for such parameters our calculations give ϵeff(A)14.84i0.12. This yields the effective refractive index neff=Reϵeff(A)3.85, as the zero-momentum TM polarized mode gives no contribution to both the effective permittivity and effective refractive index. To check this, we performed RCWA simulation [38] to calculate the normal transmission and reflection by a slab of the thickness h consisting of altenating metal and dielectric layers with the above parameters (not shown here). The results showed that both the transmission and reflection are oscillating functions of h with the oscillation period of about 390 nm. This is approximately equal to a half wavelength in a homogeneous slab with n = 3.85, as 1500 nm/3.85 ≈ 389.6 nm.

As was noted in Sec. 3, the proper choice of the background permittivity Reϵ2 can be considered as a trade-off between the transparency threshold Imϵ2(0) and the mode momentum (the effective permittivity ϵeff). For a fixed gain, however, higher ϵeff can be obtained at smaller Reϵ2 and f values, as is seen in Fig. 6. A decrease in Reϵ2, in turn, can be accomplished by tuning the composition of ternary or quaternary semiconductor compounds (see, e.g., [39]).

6. Closing remarks

In this paper, we provided a guideline for the practical implementation of extraordinary high- and low-momentum lossless plasmonic modes dealing with the simplest 1d bi-layer MM. In particular, for the TM polarization we showed that Rytov’s dispersion equation gives rise to two traveling mode solution branches. One of the branches is an ordinary solution which yields the well-known result for the effective permittivity in the quasistatic limit (the local approximation). In contrast, the other branch cannot be described within the local approximation framework. It represents an extraordinary (additional) wave, which results from strong optical nonlocality and cannot be excited at normal incidence. Furthermore, we demonstrated that both solution branches can describe plasmonic modes with infinitely high in-plane momentum, but the ordinary mode can be generated at a smaller gain coefficient providing lower transparency threshold. A decrease in the semiconductor background permittivity allows one to lower the transparency threshold, but it also lowers the nonlocal effective permittivity.

For the TE polarization, there is only one traveling mode, which can be excited at normal incidence and is capable to provide the low in-plane momentum. In the limit of zero momentum, this solution coincides with that for the TM polarization.

It is well known that the background permittivity sufficiently depends on the compound composition. Thus, one can conclude that tuning the composition of ternary and quaternary compounds, such as, e.g., InGaAs, InGaN, InAlN, AlGaN, InGaAsP, InGaAsAl and others, used for fabricating semiconductor quantum structures, could greatly improve the efficiency of loss compensation and facilitate the design of 1d MMs, supporting plasmonic modes with extraordinary high- and low momentum.

Finally, we would like to make a few remarks on the limitations of our study. Dealing with nonlocal response of the 1d MM, we did not take into account the nonlocal optical response of constituent phases. In some cases, especially for metals, this intrinsic nonlocality cannot be neglected [40]. Furthermore, when dealing with the ENZ regime, giant enhancement of the local electric field can occur that, in turn, can result in various nonlinear effects. All these issues remain open for further consideration.

Appendix: Approximate solutions of Rytov’s equation for high-momentum modes

Using the identities

sinz=2tanz/21+tan2z/2andcosz=1tan2z/21+tan2z/2,

Equation (2) can be transformed into

tanz1(tanz1+k2ϵ1k1ϵ2)+tanz2(tanz2+k1ϵ2k2ϵ1)=0,
where z1=12k0d1ϵ1ϵeff, z2=12k0d2ϵ2ϵeff. In the limit of ϵeff → ∞, k1/k2→−1, and limϵeff(tanz2)=i for branch A, limϵeff(tanz1)=i for branch B,

where i=1. Substituting this into Eq. (10), one has

tan2z2itanz2(ϵ1ϵ2+ϵ2ϵ1)10,
tan2z1+itanz1(ϵ1ϵ2+ϵ2ϵ1)10,
for branches A and B, respectively. The solutions of these equations are
f2k0dRearctan[(b2+4b)/2]ϵ2ϵeff
for branch A and
f1+2k0dRearctan[(b2+4+b)/2]ϵ1ϵeff
for branch B, where we have introduced the notation
b=i(ϵ1ϵ2+ϵ2ϵ1).

As can be seen, if ϵeff → ∞, then f → 0 (1) for branch A (B).

Acknowledgments

Work supported in part by the Ministry of Science and Technology (MOST) of Republic of China under contract no. MOST 103-2221-E-001-011-MY3.

A.P. acknowledges partial support from AF SFFP Program.

References and links

1. J. Homola, “Surface plasmon resonance sensors for detection of chemical and biological species,” Chem. Rev. 108, 462–493 (2008). [CrossRef]   [PubMed]  

2. D. de Ceglia, M.A. Vincenti, M.G. Cappeddu, M. Centini, N. Akozbek, A. D’Orazio, J.W. Haus, M.J. Bloemer, and M. Scalora, “Tailoring metallodielectric structures for superresolution and superguiding applications in the visible and near-ir ranges,” Phys. Rev. A 77, 033848 (2008). [CrossRef]  

3. V. Intaraprasonk, Z. Yu, and S. Fan, “Image transfer with subwavelength resolution to metal-dielectric interface,” J. Opt. Soc. Am. B 28, 1335–1338 (2011). [CrossRef]  

4. J. Benedicto, E. Centeno, R. Polles, and A. Moreau, “Ultimate resolution of indefinite metamaterial flat lenses,” Phys. Rev. B 88, 245138 (2013). [CrossRef]  

5. D. Wei, C. Harris, C.C. Bomberger, J. Zhang, J. Zide, and S. Law, “Single-material semiconductor hyperbolic metamaterials,” Opt. Express 24, 8735–8745 (2016). [CrossRef]   [PubMed]  

6. Y. Guo, W. Newman, C.L. Lortes, and Z. Jacob, “Application of hyperbolic metamaterial substrates,” Adv. OptoElectron. 2012, 1–9 (2012). [CrossRef]  

7. Z. Jacob, I.I. Smolyaninov, and E.E. Narimanov, “Broadband Purcell effect: Radiactive decay engineering with metamaterials,” Appl. Phys. Lett. 100, 181105 (2012). [CrossRef]  

8. S.V. Zhukovsky, A.A. Orlov, V.E. Babicheva, A.V. Lavrinenko, and J.E. Sipe, “Photonic-band-gap engineering for volume plasmon-polaritons in multiscale multilayer hyperbolic metamaterials,” Phys. Rev. A 90, 013801 (2014). [CrossRef]  

9. A.V. Goncharenko and V.U. Nazarov, “Giant nonlocal lossless permittivity at optical frequencies,” Opt. Express 23, 20439–20449 (2015). [CrossRef]   [PubMed]  

10. R. Liu, C.M. Roberts, Y. Zhong, V.A. Podolskiy, and D. Wasserman, “Epsilon-near-zero photonic wires,” ACS Photon. 3, 1045–1052 (2016). [CrossRef]  

11. S. Hrabar, I. Krois, I. Bonic, and A. Kiricenko, “Ultra-broadband simultaneous superluminal phase and group velocities in non-Foster epsilon-near-zero matamaterial,” Appl. Phys. Lett. 102, 054108 (2013). [CrossRef]  

12. A. Ciattoni, C. Rizza, and E. Palange, “Extreme nonlinear electrodynamics in metamaterials with very small linear dielectric permittivity,” Phys. Rev. A 81, 043839 (2010). [CrossRef]  

13. A. Marini and F.J. Garcia de Abajo, “Self-organization of frozen light in near-zero-index media with cubic nonlinearity,” Scient. Rep. 6, 20088 (2016). [CrossRef]  

14. S. Zhang, D.A. Genov, Y. Wang, M. Liu, and X. Zhang, “Plasmon-induced transparency in metamaterials,” Phys. Rev. Lett. 101, 047401 (2008). [CrossRef]   [PubMed]  

15. S.Y. Chiam, R. Singh, C. Rockstuhl, F. Lederer, W. Zhang, and A.A. Bettiol, “Analogue of electromagnetically induced transparency in a terahertz metamaterial,” Phys. Rev. B 80, 153103 (2009). [CrossRef]  

16. J. Zhang, S. Xiao, C. Jeppesen, A. Kristensen, and N.A. Mortensen, “Electromagnetically induced transparency in metamaterials at near-infrared frequency,” Opt. Express 18, 17187–17192 (2010). [CrossRef]   [PubMed]  

17. Y. Yang, I.I. Kravchenko, D.P. Briggs, and J. Valentine, “All-dielectric metasurface analogue of electromagnetically induced transparency,” Nature Commun. 5, 5753 (2014). [CrossRef]  

18. M.P. Nezhad, K. Tetz, and Y. Fainman, “Gain assisted propagation of surface plasmon polaritons on planar metallic waveguides,” Opt. Express 12, 4072–4079 (2004). [CrossRef]   [PubMed]  

19. M.A. Noginov, G. Zhu, M. Bahoura, J. Adegoke, C.E. Small, B.A. Ritzo, V.P. Drachev, and V.M. Shalaev, “Enhancement of surface plasmons in an Ag aggregate by optical gain in a dielectric medium,” Opt. Lett. 31, 3022–3024 (2006). [CrossRef]   [PubMed]  

20. D.A. Genov, M. Ambati, and X. Zhang, “Surface plasmon amplification in planar metal films,” IEEE J. Quant. Electron. 43, 1104–1108 (2007). [CrossRef]  

21. C. Rizza, A. Di Falco, and A. Ciattoni, “Gain assisted nanocomposite with near zero permittivity modulus at visible frequencies,” Appl. Phys. Lett. 99, 221107 (2011). [CrossRef]  

22. C. Argyropoulos, N.M. Estakhri, F. Monticone, and A. Alu, “Negative refraction, gain and nonlinear effects in hyperbolic metamaterials,” Opt. Express 21, 15037–15047 (2013). [CrossRef]   [PubMed]  

23. R.S. Savelev, I.A. Shadrivov, P.A. Belov, N.N. Rosanov, S.V. Fedorov, A.A. Sukhorukov, and Y.S. Kivshar, “Loss compensation in metal-dielectric layered metamaterials,” Phys. Rev. B 87, 115139 (2013). [CrossRef]  

24. X. Ni, S. Ishii, M.D. Thoreson, V.M. Shalaev, S. Han, S. Lee, and A.V. Kildishev, “Loss-compensated and active hyperbolic metamaterials,” Opt. Express 19, 25242–25254 (2011). [CrossRef]  

25. J.S.T. Smalley, F. Vallini, S. Shahin, B. Kante, and Y. Fainman, “Gain-enhanced high-k transmission through metal-semiconductor hyperbolic metamaterials,” Opt. Mater. Express 5, 2300–2312 (2015). [CrossRef]  

26. J.S.T. Smalley, F. Vallini, B. Kante, S. Shahin, and Y. Fainman, “Modal amplification in active waveguides with hyperbolic dispersion at telecommunication frequencies,” Opt. Express 22, 21088–21105 (2014). [CrossRef]   [PubMed]  

27. V. Popov, A.V. Lavrinenko, and A. Novitsky, “Operator approach to effective medium theory to overcome a breakdown of Maxwell Garnett approximation,” Phys. Rev. B 94, 085428 (2016). [CrossRef]  

28. S.M. Rytov, “Electromagnetic properties of a finely stratified medium,” Sov. Phys. JETP 2, 466–475 (1956).

29. B. Sturman, E. Podivilov, and M. Gorkunov, “Eigenmodes for metal-dielectric light-transmitting nanostructures,” Phys. Rev. B 76, 125104 (2007). [CrossRef]  

30. X.B. Kang, W. Tan, and Z.G. Wang, “Validity of effective medium theory for metal-dielectric lamellar gratings,” Opt. Commun. 284, 4237–4242 (2011). [CrossRef]  

31. A. Orlov, I. Iorsh, P. Belov, and Y. Kivshar, “Complex band structure of nanostructured metal-dielectric metamaterials,” Opt. Express 21, 1593–1598 (2013). [CrossRef]   [PubMed]  

32. A.A. Krokhin, J. Arriaga, L.N. Gumen, and V.P. Drachev, “High-frequency homogenization for layered hyperbolic metamaterials,” Phys. Rev. B 93, 075418 (2016). [CrossRef]  

33. J.S.T. Smalley, F. Vallini, S. Montoya, L. Ferrari, C.T. Riley, S. Shahin, B. Kante, E. Fullerton, Z. Liu, and Y. Fainman, “Light-emitting hyperbolic metasurfaces at telecom frequencies,” in Advanced Photonics 2016 (IPR, NOMA, SPPCom, SOF), OSA Tech. Digest (OSA, 2016), paper NoM3C.3. [CrossRef]  

34. J. S. T. Smalley, F. Vallini, S. Montoya, L. Ferrari, S. Shahin, C. T. Riley, B. KantÃl’, E. E. Fullerton, Z. Liu, and Y. Fainman, “Luminescent hyperbolic metasurfaces,” Nature Commun. 7, 13793 (2017). [CrossRef]  

35. J. Elser, V.A. Podolskiy, I. Salakhutdinov, and I. Avrutsky, “Nonlocal effects in effective-medium response of nanolayered metamaterials,” Appl. Phys. Rev. 90, 191109 (2007).

36. R.J. Pollard, A. Murphy, W.R. Hendren, P.R. Evans, R. Atkinson, G.A. Wurtz, A.V. Zayats, and V.A. Podolskiy, “Optical nonlocalities and additional waves in epsilon-near-zero metamaterials,” Phys. Rev. Lett. 102, 127405 (2009). [CrossRef]   [PubMed]  

37. A.A. Orlov, P.M. Voroshilov, P.A. Belov, and Y. Kivshar, “Engineered optical nonlocality in nanostructured metamaterials,” Phys. Rev. B 85, 045424 (2011). [CrossRef]  

38. M. G. Moharam, E. B. Grann, D. A. Pommet, and T. K. Gaylord, “Formulation for stable and efficient implementation of the rigorous coupled-wave analysis of binary gratings,” J. Opt. Soc. Am. A 12, 1068–1076 (1995). [CrossRef]  

39. S.P. Lepkowski, J.A. Majewski, and G. Jurczak, “Nonlinear elesticity in II-N compounds: Ab initio calculations,” Phys. Rev. B 72, 245201 (2005). [CrossRef]  

40. S. Raza, S.I. Bozhevolnyi, M. Wubs, and N.A. Mortensen, “Nonlocal optical response in metallic nanostructures,” J. Phys.: Condens. Matter 27, 183204 (2015).

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 (7)

Fig. 1
Fig. 1 The unit cell of the structure under consideration.
Fig. 2
Fig. 2 The function |F(f, ϵeff)| (logarithmic scale) at Im(ϵ2)=−0.2 and d = 100 nm for the TM polarization.
Fig. 3
Fig. 3 The function |F(f, ϵeff)| (logarithmic scale) at Im(ϵ2)=−0.2 and d = 100 nm for the TE polarization.
Fig. 4
Fig. 4 The dependence of ϵ e f f ( A ) (green curves) and corresponding ϵ 2 ( 0 ) (blue curves) vs f at different values of d for the even TM mode.
Fig. 5
Fig. 5 The dependence of ϵ e f f ( B ) (green curves) and corresponding ϵ 2 ( 0 ) (blue curves) vs f at different values of d for the odd TM mode.
Fig. 6
Fig. 6 The dependence of ϵ e f f ( A ) (green curves) and corresponding Im ϵ 2 ( 0 ) (blue curves) vs Reϵ2 for the even TM mode at d = 90 nm.
Fig. 7
Fig. 7 The dependence Im ϵ 2 ( 0 ) ( k 0 d ) (blue curves) and corresponding f(k0d) (green curves) at which ϵeff = 0 for different values of Reϵ2.

Equations (16)

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

ϵ e f f ( k ) ϵ e f f ( k ) = k 2 / k 0 2 ,
F = cos ( k 1 d 1 ) cos ( k 2 d 2 ) γ sin ( k 1 d 1 ) sin ( k 2 d 2 ) 1 = 0 ,
k 1 ϵ 1 tan k 1 d 1 2 + k 2 ϵ 2 tan k 2 d 2 2 = 0 ,
k 1 ϵ 1 tan k 2 d 2 2 + k 2 ϵ 2 tan k 1 d 1 2 = 0 ,
k 1 tan k 1 d 1 2 + k 2 tan k 2 d 2 2 = 0 ,
k 1 tan k 2 d 2 2 + k 2 tan k 1 d 1 2 = 0 ,
ϵ e f f ( k 0 d 1 , k 0 d 2 , ϵ 1 , ϵ 2 ) = r s 2 ϵ e f f ( r s k 0 d 1 , r s k 0 d 2 , ϵ 1 , ϵ 2 / r ) .
ϵ e f f = [ ( 1 f ) / ϵ 1 + f / ϵ 2 ] 1 .
ϵ 1 tan ( k 0 d 1 ϵ 1 / 2 ) + ϵ 2 tan ( k 0 d 2 ϵ 2 ) = 0 ,
sin z = 2 tan z / 2 1 + tan 2 z / 2 and cos z = 1 tan 2 z / 2 1 + tan 2 z / 2 ,
tan z 1 ( tan z 1 + k 2 ϵ 1 k 1 ϵ 2 ) + tan z 2 ( tan z 2 + k 1 ϵ 2 k 2 ϵ 1 ) = 0 ,
tan 2 z 2 i tan z 2 ( ϵ 1 ϵ 2 + ϵ 2 ϵ 1 ) 1 0 ,
tan 2 z 1 + i tan z 1 ( ϵ 1 ϵ 2 + ϵ 2 ϵ 1 ) 1 0 ,
f 2 k 0 d Re arctan [ ( b 2 + 4 b ) / 2 ] ϵ 2 ϵ e f f
f 1 + 2 k 0 d Re arctan [ ( b 2 + 4 + b ) / 2 ] ϵ 1 ϵ e f f
b = i ( ϵ 1 ϵ 2 + ϵ 2 ϵ 1 ) .
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.