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

Tunable broadband hyperbolic light dispersion in metal diborides

Open Access Open Access

Abstract

The naturally hyperbolic materials that conquer the limitations of artificially structured hyperbolic metamaterials are promising candidates for the emerging devices based on light. However, the variety of natural hyperbolic materials and their hyperbolic frequency regime are presently limited. Here, on the basis of first-principles calculations, we demonstrated a family of natural hyperbolic materials, graphite-like metal diborides, with a broadband hyperbolic region from near-IR (∼2.5µm) to the ultraviolet regime (∼248 nm). The operating hyperbolic window and negative refraction can be effectively modulated by extracting electrons from the materials, offering a promising strategy for regulating the optical properties.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Hyperbolic materials refer to a medium in which the permittivity and permeability tensor elements are of opposite signs [1]. For the electric hyperbolic materials, we only consider the permittivity written as:

$$\hat{\varepsilon } = \left( {\begin{array}{ccc} {{\varepsilon_ \bot }}&0&0\\ 0&{{\varepsilon_ \bot }}&0\\ 0&0&{{\varepsilon_\parallel }} \end{array}} \right)$$
where the subscripts ⊥ and || indicate components perpendicular and parallel to the anisotropy axis, respectively. If we have either ${\varepsilon _ \bot } > 0\; $ and ${\varepsilon _\parallel } < 0$ (type I) or ${\varepsilon _ \bot } < 0$ and ${\varepsilon _\parallel } > 0$ (type II) [24], we will have a hyperbolic light dispersion between frequency (ω) and wave vectors (kx, ky, kz),
$$\frac{{k_x^2 + k_y^2}}{{{\varepsilon _ \bot }}} + \frac{{k_z^2}}{{{\varepsilon _{||}}}} = {\left( {\frac{\omega }{c}} \right)^2}$$
where c is the speed of light. The unique hyperbolic light dispersion leads to novel properties and potential applications, such as all-angle negative refraction [58], subwavelength imaging [912] in the far field, optical nanoscale cavity [13,14], steering of signal [15,16] and sensing [1719].

Hyperbolic light dispersion has been achieved in artificially engineered metamaterials composed of metallic nanostructures embedded in a dielectric medium [6,7,9,2022]. However, the performance is limited by the finite size of the metallic components. Compared with the metamaterials, natural hyperbolic materials have obvious advantages [2328]. They require no complicated nanofabrication processes and avoid internal interfaces associated with imperfections for the electrons to scatter off, which lower the energy loss. Moreover, the hyperbolic frequency window is greatly extended without the constraint of the length scale of metallic component of metamaterials. Recently, hyperbolic dispersion has been predicted in the crystal materials with layered structures, such as MgB2 [23], cuprate and ruthenate [23], graphite [29], the tetradymites Bi2Se3 and Bi2Te3 [30], transition metal dichalcogenides (TMDs) [25]. The strongly anisotropic atomic and electronic structures of these natural hyperbolic materials offer a useful guidance for the subsequent exploration. However, the rarity of the natural hyperbolic materials and the narrow frequency window limit their device applications. Few materials have been reported to have the frequency window ranging from near-infrared to the ultraviolet regime.

In this work, we report on the discovery of a class of natural hyperbolic materials, metal diborides with a broadband of hyperbolic light dispersion from near-infrared to ultraviolet regimes. More interestingly, the hyperbolic frequency window can be effectively tuned by extracting electrons from the materials. The all-angle negative refraction for transverse magnetic (TM) polarized incident waves was also verified in these natural hyperbolic materials. Considering the experimental observation of hyperbolic dispersion in MgB2 with a limited frequency window [31,32], our findings offer not only a number of natural hyperbolic materials superior to MgB2, but also a promising strategy to regulate the hyperbolic properties of the mediums.

2. Method

Our first-principles calculations were performed within the framework of density-functional theory (DFT) as implemented in the Vienna ab initio simulation package (VASP) [33] using the projector augmented-wave method [34]. The energy cut-off for the plane-wave expansion was set to 500 eV. During geometry optimization, the lattice constants and the atomic positions were fully relaxed until the atomic forces on the atoms were less than 0.01 eV/Å and the total energy change was less than 10−5 eV. The van der Waals (vdW) interaction was taken into account by using the DFT-D3 strategy [35]. The Brillouin zone (BZ) integrations were carried out by using Monkhorst-Pack k-point grids [36] of 13×13×10 for geometry optimization, while a denser k-mesh of 38×38×29 was adopted to ensure the convergence of the optical property calculations. The Fermi level regulating was achieved by extracting electrons from the lattice and adding a homogeneous background charge of opposite sign. For the exchange-correlation functional, the generalized gradient approximation (GGA) of Perdew-Burke-Ernzerhof (PBE) [37] was used to relax the structural parameters and calculate the electronic and optical properties, while the hybrid functional (HSE06) [38] was adopted to calculate the electronic properties and permittivities of CaB2 for comparison. The validity of the DFT calculations in reproducing the electronic structures of MgB2 and other related borides has been verified in many pervious literatures [3941]. Additionally, the hyperbolic frequency regime of MgB2 predicted by our DFT calculations agrees well the experimental data [23]. We also adopted different vdW methods, including vdW-DF [42], optB88-vdW [43], optB86b-vdW [44] and DFT-D2 [45] in our calculations. It was found that different vdW methods have negligible effect on the hyperbolic frequency regime with the average difference less than 0.1 eV. The numerical simulations of negative refraction are confirmed in a 6 μm × 2 μm CaB2 sample by the finite-element method (FEM) implemented in COMSOL MULTIPHYSICS [46,47].

3. Results and discussion

The metal diborides (MB2) considered in this work have layered structures with a point group of p6/mmm, which are similar to that of graphite, as shown in Fig. 1. The atoms in the metal layer form a planar triangular lattice, while the boron layer exhibits a honeycomb configuration. The adjacent layers show an AB stacking pattern with the interlayer distances of about 1.46-2.48 Å. MgB2, a representative of this kind of materials, has been synthesized and extensively investigated, revealing interesting properties such as covalent bonds driven metallic and high-temperature superconductivity [31,32,4850]. The other fourteen metal diborides considered in this work were constructed by replacing the Mg with group IIA alkaline earth metal atoms or group IIIB-VIIB transition metal atoms. In the following parts, we took CaB2 as an example to discuss the electronic and optical properties of the metal diboride materials.

 figure: Fig. 1.

Fig. 1. The crystal structure of metal diborides (MB2) contains graphite-like layers of B, which are intercalated by hexagonal close-packed layers of metal atoms. Fifteen types of metal atoms as listed in the insets were considered in our calculations.

Download Full Size | PDF

The equilibrium lattice constants of CaB2 are a = b = 3.193 Å and c = 4.046Å, compared to those of MgB2, a = b = 3.049 Å and c = 3.483 Å. The interlayer distance of CaB2 is longer than that of MgB2 by about 0.28 Å and thus leads to weaker interlayer coupling, due to the different ionic radii, rMg = 0.74 Å and rCa = 1.04 Å. The electronic band structure of CaB2 is plotted in Fig. 2(a). We can see the intrinsic metallic nature of CaB2 with few bands crossing the Fermi level. The orbital-resolved band lines show that the metallic properties of CaB2 are mainly dominated by the σ bands of the px,y orbitals of B and π bands of pz orbitals, analogous to the case of MgB2 [51]. The larger interlayer distance in CaB2 leads to less dispersive pz bands along the z-direction and enhances the anisotropy of the electronic structures. The remarkable anisotropy of electronic band properties of CaB2 will be responsible for the anisotropic dielectric properties as described below.

 figure: Fig. 2.

Fig. 2. (a) Orbital-resolved band structure of CaB2 along the high symmetry in the Brillouin zone. The energy of the Fermi level is set to zero. The contributions of px, py orbitals of B and pz orbitals of B are indicated by red and blue disks, respectively. (b) Real and imaginary part of the permittivities of CaB2 obtained at the PBE level. The blue shaded region shows the hyperbolic frequency window. The triangle symbols indicate the hyperbolic frequency regions obtained from HSE06 functional. Real and imaginary part of the permittivities of (c) SrB2 and (d) CrB2 obtained at the PBE level.

Download Full Size | PDF

Permittivity which was commonly adopted to represent the optical properties of crystalline solids is composed of a real part and an imaginary part ɛ(ω) = Re ɛ(ω) + iIm ɛ(ω), which is contributed by the intra-band and inter-band transition. The inter-band transition of the imaginary part Im ɛ(ω) can be calculated from the standard DFT calculations, while the real part Re ɛ(ω) can be determined from the Kramers-Konig relation [52]. The contribution of intra-band transition in metallic materials can be approximately described by the Drude model [53],

$$\textrm{Im[}\varepsilon _{\alpha \beta }^{\textrm{intra}}\textrm{]} = \frac{{\gamma \omega _{p,\alpha \beta }^2}}{{\omega ({\omega ^2} + {\gamma ^2})}}$$
$$\textrm{Re[}\varepsilon _{\alpha \beta }^{\textrm{intra}}\textrm{]} = 1\textrm{ - }\frac{{\omega _{p,\alpha \beta }^2}}{{{\omega ^2} + {\gamma ^2}}}$$
Here ωp is the plasma frequency and γ is a life-time broadening which is the reciprocal of the electron life time.

The two components of the permittivity of CaB2, ${{\varepsilon }_ \bot }({\omega } )$ and ${{\varepsilon }_\parallel }({\omega } )$ along the direction perpendicular and parallel to the anisotropy axis (z-axis) are plotted in Fig. 2(b). We can see two obvious type-II hyperbolic frequency regimes with ${{\varepsilon }_ \bot }({\omega } )$ < 0 and ${{\varepsilon }_\parallel }({\omega } )$ > 0 from 0.80 eV to 1.87 eV and from 2.12 eV to 3.26 eV. The hyperbolic frequency regimes cover the near-infrared, visible and ultraviolet regimes and thus are thus accessible in experiments. In order to verify the reliability of the PBE functional in predicting the hyperbolic frequency regimes, we calculated the permittivities of MgB2 crystal. The type-II hyperbolic frequency regime 2.34-3.54 eV given by our first-principles calculations is very close to the experimental data 2.60-3.72 eV obtained from the spectroscopic ellipsometry measurements at room temperature [23]. Additionally, we found that the anisotropy in plasma frequencies of CaB2, 5.27 eV (in-plane) and 2.78 eV (out-of-plane) is more prominent than those of MgB2, 7.27 eV (in-plane) and 6.85 eV (out-of-plane), which can be attributed to the strongly anisotropic atomic and electronic structures of CaB2. The stronger anisotropy of plasma frequencies of CaB2 leads to a wider hyperbolic frequency regime than MgB2. The permittivities of SrB2 and CrB2 were presented in Figs. 2(c) and 2(d) for comparison. SrB2 has a type-II hyperbolic frequency regime 0.99-2.82 eV. Although SrB2 is isoelectronic to CaB2, the larger ionic radius of Sr leads to more band overlap along the z-direction, which contributes to weaker anisotropy. For CrB2, the type-I hyperbolic frequency regime can be ascribed to the high metallicity along the z-direction. It also has a fairly wide hyperbolic window 2.25-4.70 eV, but suffers from high energy loss.

The hyperbolic frequency regimes of the fifteen metal diborides obtained by our first-principles calculations are plotted in Fig. 3. Hyperbolic frequency windows ranging from the near-infrared regime to the ultraviolet regime can be found in these layered materials. Beside MgB2, CaB2, SrB2, BaB2, VB2 and SrB2 exhibit broadband hyperbolic frequency windows, offering a class of candidates for achieving hyperbolic light dispersion and relevant device applications working in wide-range frequencies.

 figure: Fig. 3.

Fig. 3. The hyperbolic frequency regimes of the 15 metal diborides. The Type I and Type II hyperbolic frequency ranges are shown in green and blue bold lines, respectively.

Download Full Size | PDF

Another interesting scenario of the metal diborides is the tunability of the electronic properties by regulating the Fermi level of the materials. For example, it has been found that the superconducting properties of MgB2 and other metal diborides can be manipulated by electron or hole doping [51,5456]. This idea can be naturally extended to the optical properties of the metal diborides. From the band structure of CaB2 in Fig. 2(a), we can see that along the Г-A direction the Fermi level crosses the band only in the region near the Г point, while several bands cross the Fermi level along other directions. As the wave vector along the Г-A direction corresponds to the Bloch wavefunction propagating along the z-direction, the anisotropy of the electronic band structure will be greatly enhanced if the band across with the Fermi level is avoided by lowing the Fermi level. To mimic this effect, we extracted one electron in a unit cell of bulk CaB2, corresponding to a hole doping concentration of 2.86 × 1022 cm-3. The electronic band structure of the hole-doped CaB2 is plotted in Fig. 4(a). The Fermi level is pushed down and has no intersection with the electronic bands along the Г-A direction. The permittivities of the hole-doped CaB2 show a broad hyperbolic frequency regime, 0.5-3.96 eV (0.313 - 2.48 μm) which is greatly extended compared with the undoped CaB2. The energy loss determined by imaginary part of permittivity has been obviously suppressed, as shown in Fig. 4(b).

 figure: Fig. 4.

Fig. 4. (a) The electronic band structure of CaB2 within doping a hole. (b) The permittivities of the hole-doped CaB2.

Download Full Size | PDF

The hyperbolic frequency regimes of the fifteen metal diborides at the same hole-doping concentration (extracting one electron from a unit cell) are plotted in Fig. 5. These metal diborides exhibit different responses in permittivity to the hole-doping. The hyperbolic frequency windows of MgB2 and CrB2 move to the high energy regime, whereas those of the CaB2, SrB2, ScB2 and YB2 are greatly extended compared with the undoped materials and cover the whole visible light regime. The hyperbolic light dispersion in the visible light regime is highly demanding in light device applications.

 figure: Fig. 5.

Fig. 5. Calculated hyperbolic frequency regimes of 15 metal diborides with doping a hole. The Type I and Type II hyperbolic frequency ranges are shown in green and blue bold lines, respectively.

Download Full Size | PDF

To reveal the origins of the tunable hyperbolic frequency regimes, we plotted the real part of the permittivity $\textrm{Re}\varepsilon (\omega )$ of pristine and hole-doped CaB2 in Fig. 6(a). In the hyperbolic frequency window, CaB2 has a positive $\textrm{Re}{{\varepsilon }_\parallel }({\omega } )$ and a negative $\textrm{Re}{{\varepsilon }_ \bot }({\omega } )$, and the boundaries of the window are defined by $\textrm{Re}{{\varepsilon }_\parallel }({\omega } )= 0$ and ${\textrm{Re}}{{\varepsilon }_ \bot }({\omega } )= 0$. We can see that the hole-doping push the ${{\varepsilon }_{0\parallel }}({\omega } )$ and ${{\varepsilon }_{0 \bot }}({\omega } )$ to the low and high energy regions, respectively, expanding the hyperbolic frequency regions. In addition, the intra-band and inter-band transitions of electrons contribute differently to the permittivity. The intra-band transition modeled by the free electron theory determines the permittivity at low energy region, resulting in negative values of ${\textrm{Re}}{\varepsilon }({\omega } )$. The inter-band transition, however, dominates the permittivity at high energy region. The decomposed permittivities along the perpendicular and parallel directions of the pristine and hole-doped CaB2 are plotted in Figs. 6(b) and  6(c). We can see that the enlarged hyperbolic frequency window is mainly attributed to the changes of intra-band contributions, i.e., the redshift of $\textrm{Re}{{\varepsilon }_\parallel }^{intra}({\omega } )$ and blueshift of $\textrm{Re}{{\varepsilon }_ \bot }^{intra}({\omega } )$. In the Drude model of intra-band transitions, the plasma frequency ${\omega _{p,\alpha \beta }}$ which characterizes the collective oscillations of the valence electrons plays a decisive role. The magnitude plasma frequency depends on the conducting electron density and the electron velocity $\frac{{\partial {E_{n,k}}}}{{\partial k}}$, according to the formula [48,57]:

$$\omega _{p,\alpha \beta }^2 ={-} \frac{{4\pi {e^2}}}{V}\mathop \sum \nolimits_{n;k} 2f{^{\prime}_{nk}}\left( {{e_\alpha } \cdot \frac{{\partial {E_{n,k}}}}{{\partial k}}} \right)\left( {{e_\beta } \cdot \frac{{\partial {E_{n,k}}}}{{\partial k}}} \right)$$
The hole-doping pushes the Fermi level downwards, avoiding the intersection of Fermi level with the electronic bands along the Г-A direction. This lowers the density of the conducting electrons along the z-direction and consequently the out-of-plane plasma frequency. By contrast, the total density of states at the EF level increase from 0.96 states/eV to 1.61 states/eV after hole-doping. Therefore, the in-plane plasma frequency will be increased by hole-doping. This is consistent with our DFT calculations on the plasma frequencies. Our calculations show that the in-plane plasma frequency ${\omega _{p, \bot }}$ increases from 5.27 to 7.25 eV, while the out-of-plane plasma frequency ${\omega _{p,\parallel }}$ decreases from 2.78 to 1.48 eV. The inverse variation of ${\omega _{p, \bot }}$ and ${\omega _{p,\parallel }}$ are responsible for the blueshift of $\textrm{Re}{{\varepsilon }_ \bot }^{intra}({\omega } )$ and redshift of $\textrm{Re}{{\varepsilon }_\parallel }^{intra}({\omega } )$, respectively, leading to the wider hyperbolic frequency windows of the hole-doped CaB2.

 figure: Fig. 6.

Fig. 6. (a) The real parts of the permittivities Re ɛ of the undoped CaB2 (solid lines) and doped CaB2 (dashed lines). (b) The intra-band contributions of Re ɛ. (c) The intra-band contributions of Re ɛ. The red areas (I, II and III) indicate the expanded hyperbolic regions after hole-doping. Black and red lines indicate the permittivities along vertical and parallel directions, respectively.

Download Full Size | PDF

The inter-band transition dominates the imaginary part of permittivity and thus the energy losses of the materials. The contribution of the inter-band transition to the imaginary part of permittivity can be obtained from the momentum matrix elements between the occupied and unoccupied wavefunctions [5860]:

$$\textrm{Im}{\varepsilon _{\alpha \beta }}^{({inter} )}(\omega )= \frac{{4{\pi ^2}{e^2}}}{{\Omega }}\mathop {\lim }\nolimits_{q \to 0} \frac{1}{{{{|q |}^2}}}\mathop \sum \nolimits_{c,v,k} 2{w_k}\delta \times ({{\varepsilon_{ck}} - {\varepsilon_{vk}} - \omega } )\times \langle{u_{ck + {e_\alpha }q}}\textrm{|}{u_{vk}}\rangle \times {\bigg \langle} {u_{ck + {e_\beta }q}}\textrm{|}{u_{vk}}{\bigg \rangle}^\ast$$
Obviously, $\textrm{Im}{{\varepsilon }_{\alpha \beta }}^{({inter} )}({\omega } )$ is dependent on the electronic band structure, and in particular proportional to the joint density of states (JDOS):
$$\textrm{JDOS} = \frac{1}{{{N_k}}}\mathop \sum \nolimits_{c,v,k} \delta ({{\varepsilon_{ck}} - {\varepsilon_{vk}} - \omega } )$$
where the band indices c and v are restricted to the conduction and the valence band states. The in-plane and out-of-plane components of the $\textrm{Im}{{\varepsilon }_{\alpha \beta }}^{({inter} )}({\omega } )$ of pristine and hole-doped CaB2 are plotted in Fig. 7. Hole-doping suppresses the magnitudes of the imaginary permittivity and leads to the blue shifts of the two peaks. For the in-plane imaginary permittivity, the first peak at 0.49 eV corresponds to the transitions near ${\Gamma }$ point marked by blue arrows in this figure. Hole-doping pushes the Fermi downwards and a higher photon energy is needed for the direct inter-band transition from occupied states below the Fermi level to the unoccupied states at the same k point. This accounts for the blue shifting of the absorption peaks. The absorption peak at 2.18 eV is a consequence of the “parallel-bands” effect existing in the marked yellow shaded regions. The occupied and unoccupied bands approximately parallel to each other in this region, resulting in high JDOS and strong absorption. The yellow shaded region becomes narrower after hole-doping, suggesting that the density of excited electrons at this photon energy decreases and the relevant adsorption is suppressed. The magnitude of out-of-plane permittivity is much lower than that of the in-plane permittivity, due to the low out-of-plane JDOS which can be further reduced by hole-doping. The above mechanisms also hold for other metal diborides, offering a useful guidance for optimizing the optical properties of hyperbolic materials.

 figure: Fig. 7.

Fig. 7. Imaginary parts of permittivities of pristine and hole-doped CaB2 contributed by inter-band transitions along (a) in-plane direction and (b) out-of-plane directions, respectively. Diagrams of inter-band transitions of (c) intrinsic and (d) hole-doped CaB2. The arrows indicate the direction of the transition processes between bands. The yellow shaded area represents the ‘parallel-bands’ effect.

Download Full Size | PDF

Finally, we demonstrated the all-angle negative refraction in these hyperbolic materials. Considering a transverse magnetic (TM) polarized light incident from air into a type II hyperbolic material. Let the optic axis (z-direction) be parallel to the air/MB2 interface and incident plane be the x-z plane, as shown in Fig. 8(a). We plotted the equifrequency contour (EFC) of the light dispersion,

$$\frac{{k_x^2}}{{{\varepsilon _{||}}}} + \frac{{k_z^2}}{{{\varepsilon _ \bot }}} = {\left( {\frac{\omega }{c}} \right)^2}$$
in the hyperbolic region. The EFC is a circle in air, but in hyperbolic materials it turns into a hyperbola because of the opposite signs of permittivities ${\varepsilon _ \bot }(\omega )$ <0 and ${\varepsilon _\parallel }(\omega )$ >0. From the hyperbolic EFC, we can see each ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _i}$ corresponds to a solution of ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _r}$, as shown in Fig. 8(b). This indicates that the electromagnetic wave can propagate without any cutoff in the range of lattice length scale, rather than evanescent waves like in the free space. Therefore, hyperbolic materials can support propagation of waves with large wavevectors. The Poynting vector S representing the group velocity of the light and the directional energy flux is defined as [8]
$$\langle\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over S}\rangle = \frac{1}{2}\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over E} \times \mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over H} = \frac{{\hat{\varepsilon } \cdot \mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} }}{{2\omega {\varepsilon _0}{\varepsilon _ \bot }{\varepsilon _{||}}}}H_0^2$$

 figure: Fig. 8.

Fig. 8. (a) The orientation of MB2 and the scheme of negative refraction when a TM polarized light is incident from air to MB2. The direction perpendicular to the layers is taken as the optic axis (z-axis). (b) The equifrequency contour (EFC) of the light dispersion relation. The negative refraction happens at the interface between the air (black circle) and MB2 (red hyperbola). The shaded region describes the incident angles for which negative refraction may occur. The solutions indicated by the dashed arrows are physically incorrect. (c) Simulation results for 756 nm TM light at incident of 30°. The color map shows the distribution of the electric field, and Poynting vectors are marked by black arrows.

Download Full Size | PDF

In the isotropic medium, e.g. air, the Poynting vector ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over S} _i}$ is in the same direction as the wavevector ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _i}$. When the electromagnetic wave enters into a hyperbolic material, however, the Poynting vector of refractive waves ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over S} _r}\; $ and the wavevector ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _r}$ are on the opposite side of the interface normal, as marked in Fig. 8(b). This scenario is called negative refraction. From the EFC, we can see that there are two solutions ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _r}$ and ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _r}^{\prime}$ on both side of ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _z}$ satisfying the continuity of the tangential component of ${\vec {\textrm k}}$ but according to the causality principle that the energy must flow away from the interface, i.e., ${S_z} = {\hat{e}_z} \cdot \mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over S} > 0$ [1,8,6163], only the ${\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} \over k} _r}$ is physically correct. The refraction angle can be determined by the Poynting vector as [8,24]

$${\theta _r} = {\tan ^{ - 1}}\left( {\frac{{\sqrt {{\varepsilon_{||}}} \sin {\theta_i}}}{{{\varepsilon_ \bot }\sqrt {1 - \frac{{{{\sin }^2}{\theta_i}}}{{{\varepsilon_ \bot }}}} }}} \right)$$
The expression shows that for a TM-polarized electromagnetic wave propagating along the optic axis, ${\theta _r}$ and ${\theta _i}\; $ are always of opposite signs, which is referred to as all-angle negative refraction.

The negative refraction can be mimicked by numerically solving the Maxwell equations with a finite element method (FEM). In the FEM simulation, we simplify the CaB2 crystal as a homogenous slab and put the air/CaB2 interface at x = 0, as shown in Fig. 8(a). The excitation wavelength was set to be 756 nm, corresponding to the chosen photon energy of 1.64 eV, and the incident angle of the transverse Gaussian beam is set to be 30 degrees. The permittivity tensors of CaB2 at this photon energy evaluated from abovementioned strategy are ɛ(ω) = -4.90 + 7.34i and ɛ (ω) = 8.17 + 2.52i, respectively. The electric filed distribution obtained by the FEM simulation is plotted Fig. 8(c). The Poynting vectors signed by arrows confirms the negative refraction effect at the interface.

4. Conclusions

In summary, using first-principle calculations, we demonstrated a class of metal diborides as promising hyperbolic materials with broadband hyperbolic frequency regimes and tunable optical properties. The hyperbolic frequency windows of CaB2, SrB2, VS2 and CrB2 are wider than that of MgB2 achieved in experiments, due to the enhanced structural anisotropy. More interestingly, the hyperbolic frequency window can be tuned by extracting electrons from the materials, achieving a broadband hyperbolic window from near-infrared (∼248 nm) to ultraviolet (∼2.5 µm) regimes in the hold-doped CaB2. The all-angle negative refraction for transverse magnetic polarized incident light was also verified in these natural hyperbolic materials by solving the Maxwell equations with a finite-element method. Our findings offer not only a new class of natural hyperbolic materials with broadband hyperbolic frequency windows but also a promising strategy to regulate the hyperbolic properties of the mediums.

Funding

National Natural Science Foundation of China (11774201, 21433006, 21833004); Basic Research Project of Natural Science Foundation of Shandong Province (ZR2018ZB0751).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. D. R. Smith and D. Schurig, “Electromagnetic Wave Propagation in Media with Indefinite Permittivity and Permeability Tensors,” Phys. Rev. Lett. 90(7), 077405 (2003). [CrossRef]  

2. P. Shekhar, J. Atkinson, and Z. Jacob, “Hyperbolic metamaterials: fundamentals and applications,” Nano Convergence 1(1), 14 (2014). [CrossRef]  

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

4. O. Takayama and A. V. Lavrinenko, “Optics with hyperbolic materials [Invited],” J. Opt. Soc. Am. B 36(8), F38–F48 (2019). [CrossRef]  

5. D. R. Smith, D. Schurig, J. J. Mock, P. Kolinko, and P. Rye, “Partial focusing of radiation by a slab of indefinite media,” Appl. Phys. Lett. 84(13), 2244–2246 (2004). [CrossRef]  

6. 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). [CrossRef]  

7. J. Yao, Z. Liu, Y. Liu, Y. Wang, C. Sun, G. Bartal, A. M. Stacy, and X. Zhang, “Optical Negative Refraction in Bulk Metamaterials of Nanowires,” Science 321(5891), 930 (2008). [CrossRef]  

8. Y. Liu, G. Bartal, and X. Zhang, “All-angle negative refraction and imaging in a bulk medium made of metallic nanowires in the visible region,” Opt. Express 16(20), 15439–15448 (2008). [CrossRef]  

9. H. Lee, Z. Liu, Y. Xiong, C. Sun, and X. Zhang, “Development of optical hyperlens for imaging below the diffraction limit,” Opt. Express 15(24), 15886–15891 (2007). [CrossRef]  

10. Z. Liu, H. Lee, Y. Xiong, C. Sun, and X. Zhang, “Far-Field Optical Hyperlens Magnifying Sub-Diffraction-Limited Objects,” Science 315(5819), 1686 (2007). [CrossRef]  

11. J. Rho, Z. Ye, Y. Xiong, X. Yin, Z. Liu, H. Choi, G. Bartal, and X. Zhang, “Spherical hyperlens for two-dimensional sub-diffractional imaging at visible frequencies,” Nat. Commun. 1(1), 143 (2010). [CrossRef]  

12. W. T. Lu and S. Sridhar, “Superlens imaging theory for anisotropic nanostructured metamaterials with broadband all-angle negative refraction,” Phys. Rev. B 77(23), 233101 (2008). [CrossRef]  

13. J. Yao, X. Yang, X. Yin, G. Bartal, and X. Zhang, “Three-dimensional nanometer-scale optical cavities of indefinite medium,” Proc. Natl. Acad. Sci. U. S. A. 108(28), 11327–11331 (2011). [CrossRef]  

14. X. Yang, J. Yao, J. Rho, X. Yin, and X. Zhang, “Experimental realization of three-dimensional indefinite cavities at the nanoscale with anomalous scaling laws,” Nat. Photonics 6(7), 450–454 (2012). [CrossRef]  

15. P. V. Kapitanova, P. Ginzburg, F. J. Rodríguez-Fortuño, D. S. Filonov, P. M. Voroshilov, P. A. Belov, A. N. Poddubny, Y. S. Kivshar, G. A. Wurtz, and A. V. Zayats, “Photonic spin Hall effect in hyperbolic metamaterials for polarization-controlled routing of subwavelength modes,” Nat. Commun. 5(1), 3226 (2014). [CrossRef]  

16. A. A. High, R. C. Devlin, A. Dibos, M. Polking, D. S. Wild, J. Perczel, N. P. de Leon, M. D. Lukin, and H. Park, “Visible-frequency hyperbolic metasurface,” Nature 522(7555), 192–196 (2015). [CrossRef]  

17. A. V. Kabashin, P. Evans, S. Pastkovsky, W. Hendren, G. A. Wurtz, R. Atkinson, R. Pollard, V. A. Podolskiy, and A. V. Zayats, “Plasmonic nanorod metamaterials for biosensing,” Nat. Mater. 8(11), 867–871 (2009). [CrossRef]  

18. K. V. Sreekanth, Y. Alapan, M. ElKabbash, E. Ilker, M. Hinczewski, U. A. Gurkan, A. De Luca, and G. Strangi, “Extreme sensitivity biosensing platform based on hyperbolic metamaterials,” Nat. Mater. 15(6), 621–627 (2016). [CrossRef]  

19. E. Shkondin, T. Repän, M. E. Aryaee Panah, A. V. Lavrinenko, and O. Takayama, “High Aspect Ratio Plasmonic Nanotrench Structures with Large Active Surface Area for Label-Free Mid-Infrared Molecular Absorption Sensing,” ACS Appl. Nano Mater. 1(3), 1212–1218 (2018). [CrossRef]  

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

21. S. M. Prokes, O. J. Glembocki, J. E. Livenere, T. U. Tumkur, J. K. Kitur, G. Zhu, B. Wells, V. A. Podolskiy, and M. A. Noginov, “Hyperbolic and plasmonic properties of Silicon/Ag aligned nanowire arrays,” Opt. Express 21(12), 14962–14974 (2013). [CrossRef]  

22. C. R. Simovski, P. A. Belov, A. V. Atrashchenko, and Y. S. Kivshar, “Wire Metamaterials: Physics and Applications,” Adv. Mater. 24(31), 4229–4248 (2012). [CrossRef]  

23. J. Sun, N. M. Litchinitser, and J. Zhou, “Indefinite by Nature: From Ultraviolet to Terahertz,” ACS Photonics 1(4), 293–303 (2014). [CrossRef]  

24. S. Guan, S. Y. Huang, Y. Yao, and S. A. Yang, “Tunable hyperbolic dispersion and negative refraction in natural electride materials,” Phys. Rev. B 95(16), 165436 (2017). [CrossRef]  

25. M. N. Gjerding, R. Petersen, T. G. Pedersen, N. A. Mortensen, and K. S. Thygesen, “Layered van der Waals crystals with hyperbolic light dispersion,” Nat. Commun. 8(1), 320 (2017). [CrossRef]  

26. K. Korzeb, M. Gajc, and D. A. Pawlak, “Compendium of natural hyperbolic materials,” Opt. Express 23(20), 25406–25424 (2015). [CrossRef]  

27. W. Ma, P. Alonso-González, S. Li, A. Y. Nikitin, J. Yuan, J. Martín-Sánchez, J. Taboada-Gutiérrez, I. Amenabar, P. Li, S. Vélez, C. Tollan, Z. Dai, Y. Zhang, S. Sriram, K. Kalantar-Zadeh, S.-T. Lee, R. Hillenbrand, and Q. Bao, “In-plane anisotropic and ultra-low-loss polaritons in a natural van der Waals crystal,” Nature 562(7728), 557–562 (2018). [CrossRef]  

28. H. Gao, Z. Wang, X. Ma, X. Zhang, W. Li, and M. Zhao, “Hyperbolic dispersion and negative refraction in a metal-organic framework Cu-BHT,” Phys. Rev. Mater. 3(6), 065206 (2019). [CrossRef]  

29. J. Sun, J. Zhou, B. Li, and F. Kang, “Indefinite permittivity and negative refraction in natural material: Graphite,” Appl. Phys. Lett. 98(10), 101901 (2011). [CrossRef]  

30. M. Esslinger, R. Vogelgesang, N. Talebi, W. Khunsin, P. Gehring, S. de Zuani, B. Gompf, and K. Kern, “Tetradymites as Natural Hyperbolic Materials for the Near-Infrared to Visible,” ACS Photonics 1(12), 1285–1289 (2014). [CrossRef]  

31. K. D. Belashchenko, M.v. Schilfgaarde, and V. P. Antropov, “Coexistence of covalent and metallic bonding in the boron intercalation superconductor MgB2,” Phys. Rev. B 64(9), 092503 (2001). [CrossRef]  

32. J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani, and J. Akimitsu, “Superconductivity at 39 K in magnesium diboride,” Nature 410(6824), 63–64 (2001). [CrossRef]  

33. G. Kresse and J. Furthmüller, “Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set,” Phys. Rev. B 54(16), 11169–11186 (1996). [CrossRef]  

34. P. E. Blöchl, “Projector augmented-wave method,” Phys. Rev. B 50(24), 17953–17979 (1994). [CrossRef]  

35. S. Grimme, S. Ehrlich, and L. Goerigk, “Effect of the damping function in dispersion corrected density functional theory,” J. Comput. Chem. 32(7), 1456–1465 (2011). [CrossRef]  

36. H. J. Monkhorst and J. D. Pack, “Special points for Brillouin-zone integrations,” Phys. Rev. B 13(12), 5188–5192 (1976). [CrossRef]  

37. J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made Simple,” Phys. Rev. Lett. 77(18), 3865–3868 (1996). [CrossRef]  

38. J. Heyd, G. E. Scuseria, and M. Ernzerhof, “Hybrid functionals based on a screened Coulomb potential,” J. Chem. Phys. 118(18), 8207–8215 (2003). [CrossRef]  

39. K. Kunc, I. Loa, K. Syassen, R. K. Kremer, and K. Ahn, “MgB2 under pressure: phonon calculations, Raman spectroscopy, and optical reflectance,” J. Phys.: Condens. Matter 13(44), 9945–9962 (2001). [CrossRef]  

40. S. F. Matar, “Electronic structure and chemical bonding within MgB2 and related borides from first principles,” Z. Naturforsch., B: J. Chem. Sci. 63(6), 673–680 (2008). [CrossRef]  

41. I. M. Ndassa, M. Gilleßen, and B. P. T. Fokwa, “Electronic structure, chemical bonding and magnetism of the metal-rich borides MRh6B3 (M = Cr, Mn, Fe, Co, Ni) with Th7Fe3-type structure: A density functional theory study,” Solid State Sci. 17, 14–20 (2013). [CrossRef]  

42. M. Dion, H. Rydberg, E. Schröder, D. C. Langreth, and B. I. Lundqvist, “Van der Waals Density Functional for General Geometries,” Phys. Rev. Lett. 92(24), 246401 (2004). [CrossRef]  

43. J. Klimeš, D. R. Bowler, and A. Michaelides, “Chemical accuracy for the van der Waals density functional,” J. Phys.: Condens. Matter 22(2), 022201 (2010). [CrossRef]  

44. J. Klimeš, D. R. Bowler, and A. Michaelides, “Van der Waals density functionals applied to solids,” Phys. Rev. B 83(19), 195131 (2011). [CrossRef]  

45. S. Grimme, “Semiempirical GGA-type density functional constructed with a long-range dispersion correction,” J. Comput. Chem. 27(15), 1787–1799 (2006). [CrossRef]  

46. E. J. F. Dickinson, H. Ekström, and E. Fontes, “COMSOL Multiphysics®: Finite element software for electrochemical analysis. A mini-review,” Electrochem. Commun. 40, 71–74 (2014). [CrossRef]  

47. J. H. Coggon, “ELECTROMAGNETIC AND ELECTRICAL MODELING BY THE FINITE ELEMENT METHOD,” Geophysics 36(1), 132–155 (1971). [CrossRef]  

48. J. Kortus, I. I. Mazin, K. D. Belashchenko, V. P. Antropov, and L. L. Boyer, “Superconductivity of Metallic Boron in MgB2,” Phys. Rev. Lett. 86(20), 4656–4659 (2001). [CrossRef]  

49. V. Guritanu, A. B. Kuzmenko, D. van der Marel, S. M. Kazakov, N. D. Zhigadlo, and J. Karpinski, “Anisotropic optical conductivity and two colors of MgB2,” Phys. Rev. B 73(10), 104509 (2006). [CrossRef]  

50. J. M. An and W. E. Pickett, “Superconductivity of MgB2: Covalent Bonds Driven Metallic,” Phys. Rev. Lett. 86(19), 4366–4369 (2001). [CrossRef]  

51. D. Goto, T. Machi, Y. Zhao, N. Koshizuka, M. Murakami, and S. Arai, “Improvement of critical current density in MgB2 by Ti, Zr and Hf doping,” Phys. C 392-396, 272–275 (2003). [CrossRef]  

52. S. Guan, S. A. Yang, L. Zhu, J. Hu, and Y. Yao, “Electronic, Dielectric, and Plasmonic Properties of Two-Dimensional Electride Materials X2N (X = Ca, Sr): A First-Principles Study,” Sci. Rep. 5(1), 12285 (2015). [CrossRef]  

53. B. Mortazavi, M. Shahrokhi, M. Makaremi, and T. Rabczuk, “Anisotropic mechanical and optical response and negative Poisson’s ratio in Mo2C nanomembranes revealed by first-principles simulations,” Nanotechnology 28(11), 115705 (2017). [CrossRef]  

54. N. Sudesh, S. Kumar, C. Das, G. D. Bernhard, and Varma, “Effect of graphene oxide doping on superconducting properties of bulk MgB2,” Supercond. Sci. Technol. 26(9), 095008 (2013). [CrossRef]  

55. O. De la Peña-Seaman, R. de Coss, R. Heid, and K. P. Bohnen, “Electron-phonon coupling and two-band superconductivity of Al- and C-doped MgB2,” Phys. Rev. B 82(22), 224508 (2010). [CrossRef]  

56. J. E. Hirsch, “Hole superconductivity in MgB2: a high Tc cuprate without Cu,” Phys. Lett. A 282(6), 392–398 (2001). [CrossRef]  

57. S. Laref, J. Cao, A. Asaduzzaman, K. Runge, P. Deymier, R. W. Ziolkowski, M. Miyawaki, and K. Muralidharan, “Size-dependent permittivity and intrinsic optical anisotropy of nanometric gold thin films: a density functional theory study,” Opt. Express 21(10), 11827–11838 (2013). [CrossRef]  

58. M. Shahrokhi and C. Leonard, “Tuning the band gap and optical spectra of silicon-doped graphene: Many-body effects and excitonic states,” J. Alloys Compd. 693, 1185–1196 (2017). [CrossRef]  

59. M. Shahrokhi and C. Leonard, “Quasi-particle energies and optical excitations of wurtzite BeO and its nanosheet,” J. Alloys Compd. 682, 254–262 (2016). [CrossRef]  

60. M. Fox, “Optical Properties of Solids,” Am. J. Phys. 70(12), 1269–1270 (2002). [CrossRef]  

61. P. A. Belov, “Backward waves and negative refraction in uniaxial dielectrics with negative dielectric permittivity along the anisotropy axis,” Microw. Opt. Technol. Lett. 37(4), 259–263 (2003). [CrossRef]  

62. C. Luo, S. G. Johnson, J. D. Joannopoulos, and J. B. Pendry, “Negative refraction without negative index in metallic photonic crystals,” Opt. Express 11(7), 746–754 (2003). [CrossRef]  

63. C. Tserkezis, N. Stefanou, and N. Papanikolaou, “Extraordinary refractive properties of photonic crystals of metallic nanorods,” J. Opt. Soc. Am. B 27(12), 2620–2627 (2010). [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 (8)

Fig. 1.
Fig. 1. The crystal structure of metal diborides (MB2) contains graphite-like layers of B, which are intercalated by hexagonal close-packed layers of metal atoms. Fifteen types of metal atoms as listed in the insets were considered in our calculations.
Fig. 2.
Fig. 2. (a) Orbital-resolved band structure of CaB2 along the high symmetry in the Brillouin zone. The energy of the Fermi level is set to zero. The contributions of px, py orbitals of B and pz orbitals of B are indicated by red and blue disks, respectively. (b) Real and imaginary part of the permittivities of CaB2 obtained at the PBE level. The blue shaded region shows the hyperbolic frequency window. The triangle symbols indicate the hyperbolic frequency regions obtained from HSE06 functional. Real and imaginary part of the permittivities of (c) SrB2 and (d) CrB2 obtained at the PBE level.
Fig. 3.
Fig. 3. The hyperbolic frequency regimes of the 15 metal diborides. The Type I and Type II hyperbolic frequency ranges are shown in green and blue bold lines, respectively.
Fig. 4.
Fig. 4. (a) The electronic band structure of CaB2 within doping a hole. (b) The permittivities of the hole-doped CaB2.
Fig. 5.
Fig. 5. Calculated hyperbolic frequency regimes of 15 metal diborides with doping a hole. The Type I and Type II hyperbolic frequency ranges are shown in green and blue bold lines, respectively.
Fig. 6.
Fig. 6. (a) The real parts of the permittivities Re ɛ of the undoped CaB2 (solid lines) and doped CaB2 (dashed lines). (b) The intra-band contributions of Re ɛ. (c) The intra-band contributions of Re ɛ. The red areas (I, II and III) indicate the expanded hyperbolic regions after hole-doping. Black and red lines indicate the permittivities along vertical and parallel directions, respectively.
Fig. 7.
Fig. 7. Imaginary parts of permittivities of pristine and hole-doped CaB2 contributed by inter-band transitions along (a) in-plane direction and (b) out-of-plane directions, respectively. Diagrams of inter-band transitions of (c) intrinsic and (d) hole-doped CaB2. The arrows indicate the direction of the transition processes between bands. The yellow shaded area represents the ‘parallel-bands’ effect.
Fig. 8.
Fig. 8. (a) The orientation of MB2 and the scheme of negative refraction when a TM polarized light is incident from air to MB2. The direction perpendicular to the layers is taken as the optic axis (z-axis). (b) The equifrequency contour (EFC) of the light dispersion relation. The negative refraction happens at the interface between the air (black circle) and MB2 (red hyperbola). The shaded region describes the incident angles for which negative refraction may occur. The solutions indicated by the dashed arrows are physically incorrect. (c) Simulation results for 756 nm TM light at incident of 30°. The color map shows the distribution of the electric field, and Poynting vectors are marked by black arrows.

Equations (10)

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

ε ^ = ( ε 0 0 0 ε 0 0 0 ε )
k x 2 + k y 2 ε + k z 2 ε | | = ( ω c ) 2
Im[ ε α β intra ] = γ ω p , α β 2 ω ( ω 2 + γ 2 )
Re[ ε α β intra ] = 1  -  ω p , α β 2 ω 2 + γ 2
ω p , α β 2 = 4 π e 2 V n ; k 2 f n k ( e α E n , k k ) ( e β E n , k k )
Im ε α β ( i n t e r ) ( ω ) = 4 π 2 e 2 Ω lim q 0 1 | q | 2 c , v , k 2 w k δ × ( ε c k ε v k ω ) × u c k + e α q | u v k × u c k + e β q | u v k
JDOS = 1 N k c , v , k δ ( ε c k ε v k ω )
k x 2 ε | | + k z 2 ε = ( ω c ) 2
S = 1 2 E × H = ε ^ k 2 ω ε 0 ε ε | | H 0 2
θ r = tan 1 ( ε | | sin θ i ε 1 sin 2 θ i ε )
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.