You are on page 1of 6

Demonstration of hyperbolic metamaterials at

telecommunication wavelength using Ga-doped


ZnO
Sascha Kalusniak,* Laura Orphal, and Sergey Sadofev
Department of Physics, Photonics Group, Humboldt-Universität zu Berlin, Newtonstraße 15, D-12489 Berlin,
Germany
*
kalus@physik.hu-berlin.de

Abstract: Hyperbolic metamaterials (HMMs) have attracted much attention


because they allow for broadband enhancement of spontaneous emission
and imaging below the diffraction limit. However, HMMs with traditional
metals as metallic component are not suitable for applications in the
infrared spectral range. Using Ga-doped ZnO, we demonstrate monolithic
HMMs operating at infrared wavelengths. We identify the material's
hyperbolic character by various optical measurements in combination with
theoretical calculations. In particular, negative refraction of the
extraordinary wave and propagation of light with wave vector values
exceeding that of free-space are demonstrated in the entire
telecommunication window. These findings reveal a considerable potential
for creating novel functional elements at telecommunication wavelengths.
©2015 Optical Society of America
OCIS codes: (160.3918) Metamaterials; (250.5403) Plasmonics; (160.1190) Anisotropic optical
materials.

References and links


1. V. M. Shalaev, “Optical negative-index metamaterials,” Nat. Photonics 1(1), 41–48 (2007).
2. C. M. Soukoulis and M. Wegener, “Past achievements and future challenges in the development of three-
dimensional photonic metamaterials,” Nat. Photonics 5, 523–530 (2011).
3. A. Poddubny, I. Iorsh, P. Belov, and Y. Kivshar, “Hyperbolic metamaterials,” Nat. Photonics 7(12), 948–957
(2013).
4. P. Shekhar, J. Atkinson, and Z. Jacob, “Hyperbolic metamaterials: fundamentals and applications,” Nano
Convergence 1(1), 14 (2014).
5. L. Ferrari, C. Wu, D. Lepage, X. Zhang, and Z. Liu, “Hyperbolic metamaterials and their applications,” Prog.
Quantum Electron. 40, 1–40 (2015).
6. D. Lu and Z. Liu, “Hyperlenses and metalenses for far-field super-resolution imaging,” Nat. Commun. 3, 1205
(2012).
7. C. L. Cortes, W. Newman, S. Molesky, and Z. Jacob, “Quantum nanophotonics using hyperbolic metamaterials,”
J. Opt. 14(6), 063001 (2012).
8. A. Boltasseva and H. A. Atwater, “Materials science. Low-loss plasmonic metamaterials,” Science 331(6015),
290–291 (2011).
9. A. J. Hoffman, L. Alekseyev, S. S. Howard, K. J. Franz, D. Wasserman, V. A. Podolskiy, E. E. Narimanov, D. L.
Sivco, and C. Gmachl, “Negative refraction in semiconductor metamaterials,” Nat. Mater. 6(12), 946–950
(2007).
10. G. V. Naik, J. Liu, A. V. Kildishev, V. M. Shalaev, and A. Boltasseva, “Demonstration of Al:ZnO as a
plasmonic component for near-infrared metamaterials,” Proc. Natl. Acad. Sci. U.S.A. 109(23), 8834–8838
(2012).
11. S. Sadofev, S. Kalusniak, P. Schäfer, H. Kirmse, and F. Henneberger, “Free-electron concentration and polarity
inversion domains in plasmonic (Zn,Ga)O,” Phys. Status Solidi B 252(3), 607–611 (2015).
12. S. Sadofev, S. Kalusniak, P. Schäfer, and F. Henneberger, “Molecular beam epitaxy of n-Zn(Mg)O as low-
damping plasmonic material at telecommunication wavelengths,” Appl. Phys. Lett. 102(18), 181905 (2013).
13. L. D. Landau, E. M. Lifschitz, and L. P. Pitaevskii, Electrodynamics of Continuous Media (Elsevier, 2000).
14. G. V. Naik, A. Boltasseva, “A comparative study of semiconductor-based plasmonic metamaterials,”
Metamaterials (Amst.) 5(1), 1–7 (2011).
15. Note that the Brewster angle for reflection at the interface between air and ZnO with a background dielectric
function of εb = 3.7 is αB = 63°.
16. D. D. Engelsen, “Ellipsometry of anisotropic films,” J. Opt. Soc. Am. 61(11), 1460–1466 (1971).

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32555
17. F. Intravaia and K. Busch, “Fluorescence in nonlocal dissipative periodic structures,” Phys. Rev. A 91(5),
053836 (2015).
18. A. S. Kuznetsov, S. Sadofev, P. Schäfer, S. Kalusniak, and F. Henneberger, “Single crystalline Er2O3:sapphire
films as potentially high-gain amplifiers at telecommunication wavelength,” Appl. Phys. Lett. 105(19), 191111
(2014).

1. Introduction
Metamaterials have been a subject of extensive research during the past years because these
artificial substances offer optical properties otherwise not available in nature. They typically
consist of periodically arranged discrete metallic building blocks which simultaneously
generate negative permittivity and permeability in a certain spectral range [1,2]. More
recently, hyperbolic metamaterials (HMMs) went into focus [3–5]. In their simplest version,
they are formed by stacks of alternating, subwavelength-thin metallic and dielectric layers and
can then be regarded as an effective uniaxial crystal. When adjusting material parameters
properly, the permittivity parallel (ε∥) and perpendicular (ε⊥) to the crystal axis have opposite
sign and the normally elliptic isofrequency curves of the extraordinary wave are transformed
into a hyperboloid. Consequently, HMMs provide access to wave vector states far exceeding
that of free-space and allow for propagation of otherwise evanescent waves. These high-k
states are the core of many potential applications like subwavelength imaging [6] or
broadband enhancement of spontaneous emission [7]. However, because of their large plasma
frequency the suitability of conventional metals is restricted to the ultraviolet and visible
spectral range and for operation at longer wavelengths alternative materials are required.
Quite recently, heavily doped semiconductors with a metallic-like dielectric function in
the infrared spectral range have received much attention [8]. Tunable free-electron
concentration, small optical losses and the possibility of epitaxial growth make them
promising candidates for fabrication of HMMs for infrared wavelengths. HMMs based on
semiconductors have been demonstrated at about 9 μm in an InGaAs/AlInAs superlattice [9]
and in the near-infrared spectral range at about 1.9 μm using ZnAlO/ZnO [10]. However,
operation at telecommunication wavelength requires even higher free-carrier concentrations
and well defined growth conditions are needed to avoid doping-induced degradation of the
host lattice. In this context, we recently developed a growth regime allowing for generation of
such high free-carrier concentrations in Ga-doped ZnO without significant deterioration of
crystalline quality [11]. On this basis, we demonstrate realization of ZnGaO-based HMMs
operating at wavelengths covering the entire telecommunication band. We use angle resolved
reflectivity and transmission as well as attenuated total reflection measurements in
combination with theoretical calculations to identify the hyperbolic character. In particular,
we demonstrate negative refraction as well as propagation of light with wave vector values
exceeding that of free space.
2. Experimental
The samples are grown by plasma-assisted molecular beam epitaxy (MBE) in a DCA 450
MBE system equipped with standard solid-source effusion cells (Zn and Ga) and an Addon
radio frequency plasma source providing active O. Growth is performed on a-plane sapphire
at a substrate temperature of 300 °C under strong excess of Zn. The samples consist of
alternating ZnO and ZnGaO layers. The individual layer thicknesses are defined by growth
time with corresponding growth rates calibrated on reference samples. This growth procedure
results in single crystalline c-axis oriented wurtzite films with structural perfection
comparable to ZnGaO epilayers grown under identical conditions [12]. The doping profile
across the ZnO/ZnGaO interface is assumed to be a step function based on previous X-ray
diffraction studies [11]. The root mean square (RMS) surface roughness of multilayer stacks
with a total thickness beyond 1.7 µm is about 2-3 nm, as measured over a scan area of 10x10
µm2 by atomic force microscopy. However, at extremely high Ga-doping levels additional
bumps appear, covering approximately 15% of the surface area. These features likely
originate from local polarity inversion within the ZnO host [11] and result in an effective
increase of the RMS value up to some 10 nm. Secondary ion mass spectroscopy, Hall

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32556
measurements as well as reflectance and transmission measurements are performed on
reference samples to control doping level and free-electron concentration within accuracy of a
few percent. Optical spectra are recorded with a resolution of 2 cm−1 by a BRUKER IFS66v/S
vacuum Fourier transform spectrometer. The light from a W or SiC source is focused to a spot
with 300 μm in diameter on the sample surface using a parabolic off-axis mirror and the
reflected/transmitted signal is detected by a Ge-diode or a MCT-detector. A polished ZnSe
hemisphere is used to perform attenuated total reflection measurements in Otto configuration.
Close contact between sample surface and basis of the hemisphere is achieved by mounting
the sample on a high pressure spring.
3. Results
As a first step, we discuss the performance of the HMM by presenting the calculated
components of the permittivity tensor as well as the dependence of the out-of-plane wave
vector component kz on the angle of incidence α. As justified previously [12], the permittivity
of ZnGaO is described by Drude’s dielectric function
ωp2
ε ZnGaO (ω ) = ε b − (1)
ω (ω +iΓ )
with εb = 3.7 being the background dielectric function of ZnO, ωp = (e2n/meε0)1/2 the plasma
frequency defined by the free-carrier concentration n, me the effective mass of the electrons
and Γ the mean electronic damping rate. Plasma frequency and damping rate of ZnGaO where
deduced from reflection and transmission measurements on reference films by fitting the
obtained spectra manually within a transfer matrix calculus [12]. The permittivity components
are calculated in the effective medium approximation
ε ZnGaO ε ZnO
ε = (2)
ρε ZnO + (1 − ρ )ε ZnGaO
ε ⊥ = ρε ZnGaO + (1 − ρ )ε ZnO (3)

with the fill factor ρ = dZnGaO(dZnGaO + dZnO)−1 and the dielectric functions of the individual
layers εZnGaO and εZnO = εb. Using these components, the out-of-plane wave vector component
kz of the extraordinary wave is calculated for an uniaxial crystal [13]. We focus here on a
sample consisting of 20 pairs of ZnO/ZnGaO with an individual layer thickness of dZnO = 45
nm and dZnGaO = 40 nm, as schematically sketched in Fig. 1(a). The doping level of the
ZnGaO-layer is adjusted to obtain a plasma frequency of ħωp = 1.88 eV, which centers the
spectral region where Re[ε∥] < 0 and Re[ε⊥] > 0 (Type I HMM) is attained to about 1.55 μm,
as shown in Fig. 1(b). This region is spectrally broad (more than 200 meV) and covers the
entire telecommunication band. Here, the real part of kz of the extraordinary wave increases
with α defining the in-plane wave vector kx = ω/c sin α and the isofrequency curves have
hyperbolic shape [Fig. 1(c)]. When λ approaches λc = 1.29 μm (0.96 eV), i.e. the wavelength
where Re[ε∥] changes sign from negative to positive, a sharp crossover occurs and kz
decreases with α eventually leading to metallic-like reflection. For the current plasmonic
damping, this response only appears for very large in-plane wave vectors and cannot be easily
tested using free-space reflectance measurements. As shown in Fig. 1(d), the imaginary part
of kz increases with α in the same spectral range giving rise for absorption. The figure of merit
(FoM) can be calculated from these plots by FoM = Re[kz]/Im[kz] [9]. A maximum FoM value
of 15 is attained for the present sample, which is slightly larger than that of comparable
HMMs based on ZnAlO operating even at longer wavelengths [10]. On the contrary, HMMs

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32557
Fig. 1. Performance of a ZnO/ZnGaO HMM. (a) Schematics of the sample and orientation of
the dielectric functions. (b) Calculated real (solid curves) and imaginary (dashed curves) part
of the dielectric function parallel (red) and perpendicular (blue) to the crystal axis. The real
part of the dielectric function of the sole ZnGaO is plotted for reference (black, dotted). The
spectral range where Re[ε∥] < 0 and Re[ε⊥] > 0 is attained is 0.72 eV – 0.96 eV. (c) Real part of
the out-of-plane wave vector component of the extraordinary wave for various angles of
incidence. The angle of incidence is increased in steps of 20° from α = 0° to α = 80°. Inset:
Isofrequency curve at ħω = 0.92 eV. (d) Same as (c) but imaginary part. All calculations are
performed in the effective medium approximation with Drude-parameters ħωp = 1.88 eV and Γ
= 112 meV of the ZnGaO layers.

based on noble metals cannot exhibit Type I hyperbolic behavior at telecommunication


wavelength [4] and their FoM in this spectral range is orders of magnitudes smaller [14].
Figure 2 summarizes optical spectra and respective theoretical calculations. For
transverse-magnetic (TM) polarized light (extraordinary wave), a pronounced dip in
transmission close to λc emerges when α is increased [Fig. 2(a)], directly reflecting the
dependence of the imaginary part of kz on α. This dip is absent for transverse-electric (TE)
polarized light (not shown). Zoomed plots of the TM polarized reflectivity are depicted in Fig.
2(b). At an angle of approximately 39°, a well-defined minimum in reflectivity starts to
appear close to 0.85 eV disappearing again when α is further increased. In agreement with the
calculated components of the dielectric function, the numerator of the Fresnel coefficient for
reflectance at the Air/HMM interface vanishes here and the Brewster condition is fulfilled
[15]. The residual signal level in the experiment is a result of a limited degree of polarization
of the incoming light. In TE polarization, this peculiarity is not observed and the reflectivity
signal continuously increases with α (not shown). The optical spectra can be well reproduced
describing the HMM as a slab of a uniaxial crystal [16] with the permittivity components
shown in Fig. 1(b). Figures 2(d) and 2(e) exemplarily show results of calculated transmission
and reflection spectra, respectively. States right of the vaccum light line cannot be addressed
in standard free-space reflectance measurements. In order to test the optical response of the

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32558
Fig. 2. Optical properties of ZnO/ZnGaO HMMs. (a) Transmission spectra recorded for
various angles of incidence. (b) Reflectance spectra recorded for various angles of incidence.
(c) Reflectivity excited at large in-plane wave vectors using a polished ZnSe hemisphere. The
in-plane wave vector is set by the angle of incidence through kx = (ω/c) (εZnSe)1/2 sin α. An
incidence angle of α = 25° corresponds to the vacuum light line. Inset: Experimental
configuration. The residual air gap between sample surface and basis of the hemisphere is less
than 80 nm as confirmed by Transfer-Matrix calculations. The angle of incidence is denoted on
the curves. (d) Calculated transmission spectra. (e) Calculated reflectance spectra. (f)
Transmission spectra of a ZnO/ZnGaO metamaterial with 25 layer pairs and Drude-parameters
ħωp = 1.52 eV and Γ = 98 meV of the ZnGaO layers.

metamaterial sample for larger kx values, we also performed reflectance measurements


exciting the sample through a ZnSe hemisphere, as schematically sketched in Fig. 2(c). In this
geometry, the in-plane wave vector of the incoming light exceeds that accessible in free-space
reflectance for α > 25° and evanescent waves are excited along the surface plane. Fully
consistent with the plots shown in Fig. 1(c), a reflection band starts to emerge at about 1 eV
when the in-plane wave vector is further increased. In the spectral range with hyperbolic
behavior, however, a pronounced dip in reflectance remains. We have confirmed by transfer
matrix simulations that the dip in reflectance is directly correlated with transmission through
the 1.7 μm-thick sample, demonstrating propagation of light in the HMM with large wave
vector values. An attractive feature of heavily doped semiconductors is precise adjustment of
ωp through control of the doping level which allows for tunability of the spectral region with
hyperbolic behavior. To demonstrate this, we have also fabricated a HMM sample with a
plasma frequency of ħωp = 1.52 eV of the ZnGaO layer which shifts the spectral range with
Re[ε∥] < 0 and Re[ε⊥] > 0 to 0.56 eV – 0.75 eV. This HMM exhibits similar optical properties
at the respective wavelength which agree also very well with theoretical calculations. In Fig.
2(f), transmission spectra of this sample are exemplarily shown.
A distinct property of Type I HMMs is negative refraction of the extraordinary wave. As
schematically sketched in Fig. 3, the different displacement of a transmitted beam can be
exploited to distinguish between negative and positive refraction. For this purpose, the
spectrum of the transmitted beam partly blocked by a razor blade is recorded. Depending on
the exact configuration (Set 1 or 2) negatively or positively refracted frequency components

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32559
shift away from the blade or behind the blade resulting in an increased or decreased signal in
the spectrum. The spectrum of the unblocked beam, however, contains all these frequency
components irrespective of their direction of refraction and can serve as reference spectrum.
The ratio of these two spectra (partly blocked to full beam) then directly reveals the direction
of refraction. As can be seen in the right panel of Fig. 3, negative refraction occurs in the
entire spectral range with hyperbolic behavior. On the high energy side, i.e. the spectral
region with 0 < Re[ε∥] < 1, an enhanced positive refraction is observed which eventually turns
into the standard refraction at the ZnO/Air-interface when Re[ε∥] approaches the background
dielectric function of ZnO. For TE polarized light, negative refraction is not observed.

Fig. 3. Negative refraction of the extraordinary wave in a ZnO/ZnGaO HMM. Left: Schematics
of the experimental configuration. Right panel: Intensity ratio of the partially blocked
transmitted beam and the unblocked transmitted beam for the two different configurations and
for TE polarized light in configuration 2 (light gray line). The transmission is shown for
reference, too (black, dashed line). The angle of incidence is 25°. For details see text.

5. Conclusion
In conclusion, we have demonstrated HMMs operating at telecommunication wavelengths
using heavily doped ZnGaO as plasmonic component. While operation at longer wavelengths
can be easily achieved by decreasing the doping level, extension towards shorter wavelengths
is currently hampered by deterioration of the plasmonic features when the doping level of
ZnGaO is further increased. These HMMs can now be combined with a selected emitter
material like Er2O3 which is grown on top or even embedded in the layer stack. In this way,
the effect of the hyperbolic high-k states on the spontaneous emission can be studied and
engineered. In particular, modifications of the spontaneous emission enhancement by
nonlocal properties of the metallic component might be identified [17]. Moreover, by
replacing the dielectric layer with this material, gain assisted reduction or even compensation
of the optical losses might be possible [18]. Another attractive application feasible is imaging
below the diffraction limit with telecommunication light.
Acknowledgments
The authors acknowledge financial support by the Deutsche Forschungsgemeinschaft in the
frame of SFB 951 (HIOS). We thank P. Schäfer for X-Ray measurements, K. Komander and
T. Meisel for AFM measurements as well as S. Peters from SENTECH Instruments for
ellipsometric measurements. We also thank O. Benson for helpful discussion.

#250299 Received 16 Sep 2015; revised 13 Nov 2015; accepted 16 Nov 2015; published 9 Dec 2015
© 2015 OSA 14 Dec 2015 | Vol. 23, No. 25 | DOI:10.1364/OE.23.032555 | OPTICS EXPRESS 32560

You might also like