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

Reconfigurable superdirective beamshaping using a PTX-synthesis metasurface

Open Access Open Access

Abstract

Parity-time-reciprocal scaling (PTX)-symmetry has been recently proposed to tailor the resonance linewidth and gain threshold of non-Hermitian systems with new exhilarating applications, such as coherent perfect absorber-laser (CPAL) and exceptional point (EP)-based devices. Here, we put forward a nearly-lossless, low-index metachannel formed by PTX-symmetric metasurfaces operating at the CPAL point, supporting the undamped weakly-guided fast wave (leaky mode) and thus achieving ultradirective leaky-wave radiation. Moreover, this structure allows for a reconfigurable and tunable radiation angle as well as beamwidth determined by the reciprocally scaled gain-loss parameter. We envision that the proposed PTX-symmetric metasurfaces will shed light on the design of antennas and emitters with ultrahigh directionality, as well as emerging applications enabled by extreme material properties, such as epsilon-near-zero (ENZ) and beyond.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In the past decade, there has been increasing research interest in non-Hermitian wave systems, which can exhibit the exotic spectral degeneracy (i.e., exceptional point [18]) and the self-dual singularity that allows co-existence of the coherent perfect absorber (CPA) and the laser (i.e., CPAL point [913]). These isolated singular points, at which Taylor series expansion fail to converge, are missed in generic situations and may require special engineering of a quantum [14], optical [113], acoustic [15], elastic [16,17], and electronic [18,19] systems. Perhaps, one of the most interesting fields emerging in non-Hermitian physics is parity-time (PT)-symmetry, since it provides an experimentally-accessible platform to study phase transitions and EPs in the eigenspectrum of non-Hermitian Hamiltonians [43]. In this context, many potential applications have been proposed, including unidirectional reflectionless propagation [4,20], negative refraction [21], and sub-diffraction focusing [22], unidirectional invisibility [23], non-reciprocal devices [8], high-contrast optical amplitude modulators enabled by the CPAL effect, and high-performance sensors based on the EP and CPAL singularities [2427]. The CPA-laser self-dual device is of particular interest. In general, a laser oscillator emits coherent outgoing radiations, while its time-reversed counterpart, CPA, is a dark medium absorbing incoming radiation from all directions. Nonetheless, a PT–symmetric CPAL device can act as a CPA and a laser, which are switchable at a given wavelength, for which the two eigenvalues of its scattering matrix approach zero and infinity, corresponding respectively to the CPA mode and the laser mode. The CPAL action has been theoretically proposed [9,10] and later experimentally demonstrated using optical waveguides and coupled resonators [11,12]. Furthermore, we have recently proposed a simple, low-profile CPAL devices based on PT-symmetric metasurfaces and have experimentally demonstrated this concept in the radio-frequency region by using a transmission line-based equivalent circuit [28,29]. In this context, a new generalized PT-symmetry, termed PTX-symmetry, has been proposed to reduce the gain threshold and tailor the resonance linewidth of the PT-symmetric system without affecting its eigenvalues, since the two system share the same eigenvalues, but different eigenstates and physical structures [2,30,31]. This enables breaking the balance between gain and loss coefficients in PT-symmetric non-Hermitian systems, thus increasing the design flexibility.

In the paper, we will present a tunable and ultradirective antenna or beamshaper based on the PTX-symmetric metasurface operating at the CPAL point. This device consists of a pair of active and passive metasurfaces with surface resistances of $- Z/\kappa _{}^{}$ and $\kappa Z$, respectively [see Fig. 1(a)], where Z is the characteristic impedance of the dielectrics of thickness d between the two metasurfaces, and $\kappa$ is the dimensionless scaling factor. The spatially distributed gain and loss, sourced respectively from the active metasurface with negative surface resistance and the passive metasurface with positive surface resistance, form the PTX-symmetric optical system [1]. In contrary to typical PT-symmetric systems that require balanced gain and loss (i.e., $\kappa = 1$), the PTX-symmetric systems comprise reciprocally scaled and flipped-sign surface resistances and host media, given by:

$${{Z}_{s}}({y}) = (\kappa Z)\delta ({y} + {d}/2) + ( - Z/\kappa )\,\delta ({y} - {d}/2)$$
$${Z}({y}) = \left\{ \begin{array}{l} \textrm{Z}/\kappa \quad \textrm{if}\;{y} > {d}/2\\ \textrm{Z}\quad \textrm{if}\; - {d}/2 < {y} < {d}/2\\ \kappa \textrm{Z}\quad \textrm{if}\;{y} < - {d}/2 \end{array} \right.,$$
where $\delta$ is the Kronecker delta function. Such PTX-symmetric metasurfaces display similar eigenvalue evolutions and phase transitions as those observed in the PT-symmetric counterpart, as shown in Fig. 2. Moreover, it is possible that at the CPAL point the loss parameter is greater than the gain parameter, showing similarities to the abnormal effect of “loss-induced revival of lasing” [8,22]. While scattering from PTX-symmetric metasurfaces has been studied in Ref. [24,29,31], propagation characteristics of the PTX-symmetric metasurface waveguide or metachannel sketched in Fig. 1(b), as well as the possible leaky radiation, are yet to be studied. Here, we will theoretically and numerically study its eigenmodes and the associated effective material properties, as well as new physical phenomena and potential applications underlying them. As detailed in the following, we will show that this metachannel can exhibit extreme dielectric properties, such as epsilon-near-zero (ENZ) observed in a dispersive lossy medium [6,7] or a waveguide around the cutoff [32,33], and, moreover, such a property can be independent of the value of $\kappa$. More interestingly, the propagation constant of the guided transverse electric (TE) mode can be continuously tuned from zero to the wavenumber of the background by changing the dimensionless gain-loss parameter $\gamma _{}^{}$ denoted in Fig. 1(a). In regard to the effective medium theory [32,33], the effective permittivity of the PTX-synthetic metachannel can vary from ENZ to that of the background medium, i.e., $0 \le \textrm{Re} [{\varepsilon _{{eff}}}] \le {\varepsilon _1},{\varepsilon _2}$. However, different from traditional ENZ and low-index media [6,7,32,33] supporting non-radiative fast-wave propagation mode. The proposed non-Hermitian open structure can have a fast-wave leaky mode, of which the ${\mathop{\rm Im}\nolimits} [{\varepsilon _{{eff}}}]$ associated with the attenuation rate can be vanishingly small. Such properties are intriguing and counterintuitive, since the leaky mode is typically lossy due to the radiation loss finite power damping rate (i.e., a combination of scattering and material losses), setting up a physical bound for the effective aperture size of leaky-wave waveguide or antenna. The “undamped” fast-wave leaky mode obtained in the PTX-symmetric metasurfaces may therefore be exploited to build a leaky-wave antenna with an ultralarge radiating aperture size and thus a superdirective radiation pattern. Moreover, the beam angle can be reconfigured between broadside and end-fire by varying surface resistances of metasurfaces. If the active metasurface is backed with a suitable low-index medium, satisfying the PTX-symmetry condition, ultradirective and angle-selective radiation/emission can be achieved in any dielectric background medium.

 figure: Fig. 1.

Fig. 1. (a) Schematics of PTX-symmetric metasurfaces composed of an active metasurface (-R2) and a passive metasurface (R1). The PTX-symmetric metasurfaces can be regarded as a metachannel having a longitudinal propagation constant $\beta $, varied between zero and the wavenumber of the background medium 1, resulting in an effective permittivity, $0 \le {\varepsilon _{{eff}}} \le \varepsilon .$ (b) Transmission-line network model for scattering of plane waves by PTX-symmetric metasurfaces, and (c) radiation from a line source inside PTX-symmetric metasurface.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Evolution of two eigenvalues of the scattering matrix of the PTX-symmetric metasurfaces shown in Fig. 1(b).

Download Full Size | PDF

2. Concept of the PTX-symmetry CPAL system and generation of a superdirectional beam

Figure 1(b) illustrates the PTX-symmetric metasurfaces, of which the surface resistance of the passive metasurfaces and the characteristic impedance of its host substrate are scaled by the dimensionless scaling factor $\kappa$, while those of the active sheet and its substrate are scaled by $1/\kappa .$ We first consider scattering of the TE-polarized plane wave from The PTX-symmetric metasurfaces [Fig. 1(b)], which can be modeled using the two-port transmission-line network (TLN) shown in Fig. 1(b). In the TLN model, the tangential wavenumber and characteristic impedance of the i-th medium are: ${{k}_{{i},{y}}} = \sqrt {{k}_{i}^2 - {{({{k}_1}\sin \alpha )}^2}}$ and ${{Z}_{i}} = {\eta _{i}}{{k}_{i}}/{{k}_{{i},{y}}},$ where $\alpha$ is angle of incidence and ${\eta _{i}} = \sqrt {{\mu _{i}}/{\varepsilon _{i}}}$ is the intrinsic impedance of the material. Active and passive metasurfaces are separated by a dielectric slab with impedance $\eta = \sqrt {\mu /\varepsilon }$ and wavenumber ${k} = \omega \sqrt {\mu \varepsilon }$. To satisfy the PTX-symmetry, the required surface impedances of metasurfaces and the permittivity of each layer are given by:

$${\varepsilon _1} = \varepsilon \frac{{1 + {{\cot }^2}\alpha }}{{1 + {\kappa ^2}{{\cot }^2}\alpha {\kern 1pt} }}$$
$${\varepsilon _2} = \varepsilon \frac{{1 + {\kappa ^4}{{\cot }^2}\alpha {\kern 1pt} }}{{1 + {\kappa ^2}{{\cot }^2}\alpha {\kern 1pt} }}$$
$${{R}_1} = \kappa \gamma \sqrt {1 + {{\tan }^2}\alpha /{\kappa ^2}{\kern 1pt} } \eta$$
$${{R}_2} ={-} (1/\kappa )\gamma \sqrt {1 + {{\tan }^2}\alpha /{\kappa ^2}{\kern 1pt} } \eta$$
where $\gamma$ is a dimensionless gain-loss parameter, $\varepsilon$ and $\mu$ are respectively the permittivity and permeability of the dielectric slab sandwiched by the two metasurfaces. The outgoing scattered waves and the incoming waves can be related by the scattering matrix [28], which can be obtained by matching boundary conditions and considering the field discontinuities across the metasurface. Explicit expressions of scattering parameters can be found in Ref. [24]. It is important to note that the scattering matrix unveils that when electrical length between the two metasurfaces ${x = }{{k}_{y}}{d = }\pi /2$, with n being a positive integer, all scattering coefficients become independent of $\kappa$ and identical to those in the PT-symmetric metasurfaces (i.e., $\kappa$= 1). This also enforces another necessary condition for the PTX-symmetry:
$${d } = \frac{{\sqrt {1 + {{\tan }^2}\alpha /{\kappa ^2}{\kern 1pt} } }}{4}{\lambda _0}$$
where ${\lambda _0}$ is the operating wavelength.

Figure 2 presents the evolution of the two eigenvalues ${\lambda _ \pm }$, given in [24], of scattering matrix as a function of $\gamma _{}^{}$; here $\alpha $ and $\kappa $ can be any arbitrary value provided that the condition in Eq. (1) is valid. From Fig. 2, we find that there exists an exceptional point dividing the system into the exact symmetry phase with unimodular eigenvalues and the broken symmetry phases with non-unimodular ones. In the broken symmetry phase, a self-dual singularity, CPAL point, appears at $\gamma = 1/\sqrt 2$, where the eigenvalues approach zero and infinity.

Let’s now consider PTX-symmetric metasurfaces in Fig. 1 as a parallel-plate waveguiding structure operating at the CPAL point. The seemingly unrelated scattering properties at the CPAL point may provide a useful guideline for manipulation of the effective medium property of this waveguiding structure. In general, the transverse-resonance method, considering the TLN model of the transverse cross section of the waveguide, is commonly used in analyzing the complex propagation constant of a waveguide. The same technique can be applied to study the PTX waveguide. Further, in this case, the transverse TLN model [Fig. 1(c)] is similar to that used in the scattering problem [Fig. 1(b)] [25,28]; however, each transmission line section has a transverse propagation constant ${{k}_{{i,y}}} = \sqrt {{k}_{i}^2 - {\beta ^2}}$ and a characteristic impedance ${{Z}_{i}} = \eta {{k}_{i}}/{{k}_{{i,y}}}$ (TE mode), where $\beta$ is the longitudinal propagation constant. We note that once the transverse resonance relation is satisfied, at any point along the y-axis, the sum of the input impedances seen looking to either side is zero, namely ${Z}_{{in}}^{( + )} + {Z}_{{in}}^{( - )} = 0,$ where ${Z}_{{in}}^{( + )}$ and ${Z}_{{in}}^{( - )}$ are input impedances seen looking to $+ \hat{\boldsymbol{y}}$ and $- \hat{\boldsymbol{y}}$ at any point along the resonant line, $- {d}/2 \le {y} \le {d}/2$ [28]. The dispersion equation derived based on the transverse resonance relation is given by:

$${\left( {\frac{1}{{{{R}_2}}}{ + }\frac{{\sqrt {{k}_2^2 - {\beta^2}} }}{{\omega \mu }}} \right)^{ - 1}} + \frac{{1 + {j}\left( {\frac{{\omega \mu }}{{{{R}_1}\sqrt {{k}_{}^2 - {\beta^2}} }}{ + }\frac{{\sqrt {{k}_1^2 - {\beta^2}} }}{{\sqrt {{k}_{}^2 - {\beta^2}} }}} \right)\tan \left( {\sqrt {{k}_{}^2 - {\beta^2}} d} \right)}}{{\frac{1}{{{{R}_1}}}{ + }\frac{{\sqrt {{k}_1^2 - {\beta ^2}} }}{{\omega \mu }} + {j}\frac{{\sqrt {{k}_{}^2 - {\beta ^2}} }}{{\omega \mu }}\tan \left( {\sqrt {{k}_{}^2 - {\beta^2}} d} \right)}} = 0$$

When CPAL conditions (Eq. (2) with $\gamma = 1/\sqrt 2$ and Eq. (3)) are met, solving Eq. (1) leads to a propagation constant $\beta = {{k}_1}\sin \alpha$. It is worthwhile mentioning that the laser mode also exists when ${Z}_{{in}}^{( + )} + {Z}_{{in}}^{( - )} = 0$ everywhere along the y direction, for which transmission and reflection coefficients are infinite. To better illustrate this idea, let’s consider the interface between the active metasurface and the host medium as an example. At the CPAL point, ${Z}_{{in}}^{( - )}|{_{{y} ={-} {d}/2}} = \eta \sqrt {{{x}^2} + {{\tan }^2}\alpha } /{{x}^2}$ and ${Z}_{{in}}^{( + )}|{_{{y} ={-} {d}/2}} ={-} \eta \sqrt {{{x}^2} + {{\tan }^2}\alpha } /{{x}^2},$ such that the reflection coefficient for oblique TE, ${{r}^ + }{ } = ({ Z}_{{in}}^{( + )} - {Z}_{{in}}^{( - )})/({Z}_{{in}}^{( + )} + {Z}_{{in}}^{( - )})$, is infinity. The CPA mode, which is of less interest in antenna and emitter applications, is achieved with the same setup, but requiring a different initial phase offset between two incident waves [17]. As a result, the CPAL point that achieves the lasing/amplification effect in the scattering event, also implies the validity of the transverse resonance relation in the waveguiding scenario. In other words, if the CPAL condition is satisfied for an TE-polarized plane wave incident at angle $\alpha ,$ the longitudinal propagation constant of the guided TE mode in the PTX waveguide is given by: ${\beta _{}} = {{k}_1}\sin \alpha .$ Further, the fast wave exists (i.e., $0 \le \beta \le {k}$), provided that the conditions in Eq. (2) are satisfied, leading to ${{k}_1}\sin \alpha = {k/}\sqrt {1 + {\kappa ^2}{{\cot }^2}\alpha } \le {k}$.

Figure 3 shows the dispersion diagram for the PTX waveguide with its geometric and material profiles described by Eqs. (1) and (2), and $\gamma = 1/\sqrt 2$ (i.e., the system is operated around the CPAL point). It can be evidently seen that at the CPAL frequency f0, $\beta = 0$, k/2, and $\sqrt 3 k/2$ can be obtained by setting (a)$\alpha = 0$, (b) $\alpha = \pi /6$, and (c) $\alpha = \pi /3$. The results presented in Fig. 3 validate the proposed concept, showing that the fast-wave propagation mode with $\beta < {{k}_1}$ can exist at the CPAL frequency, which infers a low effective permittivity ${\varepsilon _{{eff}}}/{\varepsilon _1}{ } = { }\sin^{2}\alpha .$ We should emphasize that when compared with traditional ENZ medium made of metamaterials or Drude-dispersion semiconductors/metals, the proposed low-index metachannel may not only ease manufacturing complexity, but also address the long-standing high loss issue at the narrow ENZ band, so as to facilitate the practice of ENZ-enabled applications, including leaky-wave antennas [7,34], superluminal effect, energy squeezing, and enhanced nonlinear wave mixing [6,7,32,33]. We should also point out from Fig. 3 that radiation (i.e., leaky mode) may occurs in this unbounded fast-wave medium, and the beam angle with respect to the broadside direction is simply $\alpha = {\sin ^{ - 1}}(\beta /{{k}_1})$ in the medium 1 and $\alpha ^{\prime} = {\sin ^{ - 1}}(\beta /{{k}_2}) = {\sin ^{ - 1}}({{k}_1}\sin \alpha /{{k}_2})$ in the medium 2 [Fig. 1(b)]. Ideally, the nearly lossless leaky mode obtained here (see Fig. 3 for ${\mathop{\rm Im}\nolimits} [\beta ]$ at the CPAL point) can lead to a huge radiating aperture and thus an ultrahigh directivity of the radiation pattern. Further, the beam angle can be tuned over a wide range from broadside to end-fire, depending on the metasurface impedance and the dielectric constant of the host medium.

 figure: Fig. 3.

Fig. 3. Dispersion relations for the PTX-symmetric metachannel when the electrical length between two metasurface is x = 2 (black), x = 1 (red), and x = 0.5 (green) in Fig. 1 under the CPAL condition (Eqs. (2) and (3) with $\gamma = 1/\sqrt 2$) with (a)$\alpha = 0$ (b)$\alpha = \pi /6$ and (c) $\alpha = \pi /3$; here, the solid and dashed lines represents the real and imaginary parts, respectively.

Download Full Size | PDF

According to the Lorentz reciprocity theorem [35,36], if a current density J1 placed at point r1 can produce an electric field E1 at point r2, their product remains constant even when the position of source and observer are switched. In the same vein, one can assume that the tangential electric field Es on PTX-symmetric metasurfaces is induced by an incident plane wave sustained by the current density JFF placed in the volume VFF. The CPAL point implies that a single source (Js) of arbitrary input amplitude can produce a huge Es. In fact, with the same structure, the CPA effect requires placing two sources with appropriate phase offsets on each side of the PTX-symmetric metasurfaces. Now, we apply reciprocity considerations to evaluate the radiated field EFF(x,y) produced by the equivalent surface current densities ${\mathbf{J}_{s}}({x})\delta ({y} \pm {d}/2)$ on metasurfaces enclosed by the volume VMTS. The reciprocity formula, $\int\limits_{{{V}_{{MTS}}}} {{\mathbf{J}_{s}}} \cdot {\mathbf{E}_{s}}{dV} = \int\limits_{{{V}_{{FF}}}} {{\mathbf{J}_{{FF}}}} \cdot {\mathbf{E}_{{FF}}}{dV}$, states that when the PTX metasurfaces operated at the CPAL point is excited by a source placed within it or in its proximity, strong and angle-selective radiated fields can be produced in the far (Fraunhofer) field. We also analyze radiation from an electric line source ($\overline {J} = \boldsymbol{\hat{z} }{{I}_0}\delta ({x})\delta ({y})\;\;[\rm{A}/{\rm{m}^2}]$) placed in the middle of a PTX-symmetric metasurfaces, as shown in Fig. 1(c). The radiation problem can be described by the transverse-equivalent network in Fig. 1(c). The electric field in the background produced by a unit amplitude electric line source can be represented as an inverse Fourier transform [37],

$${{E}_z}({x},{y}) = \frac{1}{{2\mathrm{\pi }}}\int\limits_{ - \infty }^{ + \infty } {\mathrm{\tilde{E}}_z^{}} ({{k}_{{i,x}}}){{e}^{ - {j}({{k}_{{i,x}}}{x} + {{k}_{{i,y}}}{y})}}{d}{{k}_{{i,x}}}$$
where ${\mathop{\rm Im}\nolimits} [{{k}_{y}}] \le 0$, necessary for satisfying the radiation condition at infinity. From the TLN model in Fig. 1(c), the spectral electric field on the interface between the active metasurface and the background medium can be derived as:
$${\tilde{E}}_z^{( + )}({{k}_{x}}) = {{I}_{g}}{Z}{{e}^{{j}\frac{{\beta {d}}}{2}}}\frac{{({{e}^{{j}\beta {d}}} + {\Gamma _1})(1 + {\Gamma _2})}}{{2({{e}^{2{j}\beta {d}}} - \Gamma _1^{}\Gamma _2^{})}}$$
where
$${\Gamma _{i}} = \frac{{({{Z}_{i}}{/{/} }{{R}_{i}}) - {Z}}}{{({{Z}_{i}}{/{/} }{{R}_{i}}) + {Z}}}$$
where the characteristic impedance, Z, for the TE polarization is ${Z} = ({{k}/{{k}_{y}}} )\eta ,$ and the vertical wavenumber ky depends on the spherical angle $\theta$ as ${{k}_{y}} = \sqrt {{k}_1^2 - {{({k}_1^{}\sin \theta )}^2}} = {{k}_1}\cos \theta$. Similarly, the spectral electric field on the interface between the passive metasurface interface and the background medium can be expressed as:
$${\tilde{E}}_z^{( - )}({{k}_{x}}) = {{I}_{g}}{Z}{{e}^{{j}\frac{{\beta {d}}}{2}}}\frac{{(1 + {\Gamma _1})({{e}^{{j}\beta {d}}} + {\Gamma _2})}}{{2({{e}^{2{j}\beta {d}}} - \Gamma _1^{}\Gamma _2^{})}}$$

The TLN model can be used to determine fields radiated by a line source by means of application of the Lorentz reciprocity theorem. Subsequently, the far-zone electric field ($\rho \gg {d}$) can be obtained through an asymptotic evaluation of Eq. (5) [3438]. The result for the upper half-plane can be written as:

$${E}_z^{({\pm} )}(\rho ,\theta ) = {E}_{z}^{{ff},({\pm} )}(\theta )\frac{{{{e}^{ {\mp} {j}{{k}_{i}}\rho }}}}{{\sqrt \rho }}$$
where $\theta$ is the angle measured from broadside, and ${\pm}$ in the parentheses represents the upper and lower half-planes. The normalized far-field patterns are:
$${E}_{z}^{{ff},({ + })}(\theta ) = \cos \theta \sqrt {\frac{{{j}{{k}_2}}}{{2\pi }}} {\tilde{E}}_{z}^{( + )}({{k}_2}\sin \theta )$$
$${E}_{z}^{{ff},( - )}(\theta ) = \cos \theta \sqrt {\frac{{{j}{{k}_1}}}{{2\pi }}} {\tilde{E}}_{z}^{( - )}({{k}_1}\sin \theta )$$

The radiated power density in the upper and lower half-planes are given by:

$${{P}^{({\pm} )}}(\theta ) = \frac{{|{E}_{z}^{{ff},({\pm} )}(\theta ){|^2}}}{{2\eta }}$$

Figures 4(a)-(c) show the theoretical and simulated results of far-field radiation patterns for the PTX-symmetric metasurfaces with $\kappa = 0.5$, 1, and 2, respectively. The PTX-symmetric metasurfaces operated near the CPAL point is excited by a line source [Fig. 1(c)]; here, the operating frequency ${f} = {{f}_0} - \delta {f}$ where ${{f}_0}$ is the CPAL frequency and $\delta {f} = { }{10^{ - 1}}{{f}_0}.$ Figs. 4(d)-(f) are the corresponding simulated contours of electric field distributions with $\kappa = 1$. It can be clearly seen from Fig. 4 that analytical (lines) and numerical (dots) results are in excellent agreement, and that radiation from the line source can be reshaped into a directive beam towards the desired direction, regardless of the value of $\kappa .$ The beam angle can be continuously tuned from broadside towards end-fire direction ($\theta = {0^o},{30^o},{45^o}$ and ${60^o}$) by varying the metasurface impedances (Eq. (2) with $\gamma = 1/\sqrt 2$ and Eq. (3)). The radiation pattern is somewhat asymmetric, due to the nature of unidirectional scattering in the generalized PT systems [19] and the reciprocal scaling operation. The beam angles on both sides are $\alpha$ (medium 1) and $\alpha ^{\prime}$ (medium 2); $\alpha = \alpha ^{\prime}$ when $\kappa = 1$, which makes the system degenerate into the PT-symmetric one. We note that at a fixed beam angle, the beamwidth increases with increasing the reciprocal scaling factor $\kappa$. From the dispersion diagram in Fig. 3, we find that the CPAL point is in the fast-wave region (leaky mode), with an abnormal zero attenuation constant, i.e., ${\mathop{\rm Im}\nolimits} [\beta ]\sim 0$, owning to the contactless gain-loss interaction. As a result, the proposed leaky-wave structure can have an ultralarge effective radiating aperture, which in turn leads to ultrahigh directivity. The simulated results also show that when $\alpha = 0,$ the electric fields inside the waveguiding structure have a nearly constant phase distribution, namely ENZ-like characteristics with $\textrm{Re} [\beta ] \approx 0$ can be obtained. Also, the amplitude contours reveals that this fast-wave propagation mode is nearly undamped, i.e., ${\mathop{\rm Im}\nolimits} [\beta ]\sim 0$.

 figure: Fig. 4.

Fig. 4. Far-field radiation pattern and electric field distribution for PTX metasurfaces excited by a line source, which is operated at the CPAL condition with (a)$\alpha = 0$ (b)$\alpha = \pi /6$ and (c) $\alpha = \pi /3$.

Download Full Size | PDF

Figure 5(a) shows radiation patterns of the CPAL-locked PTX-symmetric metasurfaces fed by a line source, with $\alpha = 0$ and $\kappa$ being varied from 1/5 to 5. It can be seen that the beam angle is locked to the broadside direction, while the beamwidth is tuned by $\kappa $. Here, we note that despite the permittivity ${\varepsilon _1}$ of the dielectric background could differ, the directive pencil-beam can be formed in the desired direction by properly choosing the value of $\kappa$. Figure 5(b) is similar to Fig. 5(a), but for beam angle $\alpha = {45^o}$. Figure 5(c)-(e) shows the corresponding snapshot of electric field distributions. Again, it is evident that the PTX-symmetric metasurfaces can form a pencil-beam in any arbitrary direction, which can be adapted to any background dielectrics. Our results show that a highly directive and reconfigurable RF antenna/lens or optical emitter can be realized using the CPAL singularity, at which the transverse resonance relation is satisfied at any point of arbitrary cross sections of the metachannel. Leaky-wave antennas based on guided-wave devices with periodic grids/slots have been widely studied in RF, microwave, and even optical regions. However, their effective aperture size, especially for optical applications, is limited by the non-negligible attenuation rate. Also, the occurrence of higher-order spatial harmonics could produce unwanted grating lobes. These long-standing challenges may be addressed using the PTX-synthetic leaky-wave structures, which are homogeneous and non-graded surfaces.

 figure: Fig. 5.

Fig. 5. (a) Far-field radiation pattern and electric field distribution for PTX metasurfaces excited by a line source, which is operated at the CPAL condition with $\alpha = 0$ and different values of κ. (b) Similar to (a), but for α = π/4.

Download Full Size | PDF

We also briefly discuss the practical implementation of PTX-symmetric metasurfaces. The positive surface resistance can be readily achieved by a passive metasurface with suitable loss. In the optical region, an active metasurface could be a patterned thin layer of material with negative conductivity such optically-pumped 2D materials [[3941].], organic dyes, or semiconductors. In the microwave region, the active metasurface could be a metasurface loaded with negative-impedance converters [41,42]. The low-index host medium, required for PTX-symmetry, can be realized with the Drude-type material, which could be wire- or resonator-based metamaterials in the microwave regions or doped semiconductor in the optical regions.

3. Conclusions

We have presented the superdirective antenna or emitter based on the PTX-symmetric metasurfaces. We have first conducted the eigenmodal analysis for this non-Hermitian open structure, showing that a nearly undamped fast-wave mode (leaky mode) can exist at the CPAL point. Then, we have theoretically and numerically studied radiation from PTX-symmetric metasurfaces fed by a line source, demonstrating that it is possible to produce ultradirective beam, with a controllable beam angle ranging from broadside to end-fire direction, while vanishingly small side lobes. Furthermore, for different background dielectrics (even with extreme permittivity), the focused beam can be locked to a specific direction by varying the reciprocally scaling factor and the impedance profile of metasurfaces. Our results may open up a new pathway for building the adaptive and reconfigurable superdirective metasurface antennas and emitters, as well as the unbounded low-index artificial media.

Funding

National Science Foundation (2210977).

Disclosures

The authors declare no conflict of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. R. El-Ganainy, K. G. Makris, D. N. Christodoulides, and Z. H. Musslimani, “Theory of coupled optical PT-symmetric structures,” Opt. Lett. 32(17), 2632 (2007). [CrossRef]  

2. C. E. Rüter, K. G. Makris, R. El-Ganainy, D. N. Christodoulides, M. Segev, and D. Kip, “Observation of parity–time symmetry in optics,” Nat. Phys. 6(3), 192–195 (2010). [CrossRef]  

3. A. Regensburger, Ch. Bersch, M.A. Miri, G. Onishchukov, D. N. Christodoulides, and U. Peschel, “Parity–time synthetic photonic lattices,” Nature 488(7410), 167–171 (2012). [CrossRef]  

4. L. Feng, Y. L. Xu, W. S. Fegadolli, M. H. Lu, J. E. B. Oliveira, V. R. Almeida, Y. F. Chen, and A. Scherer, “Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies,” Nat. Mater. 12(2), 108–113 (2013). [CrossRef]  

5. B. Peng, Ş. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys. 10(5), 394–398 (2014). [CrossRef]  

6. S. Savoia, G. Castaldi, V. Galdi, A. Alù, and N. Engeta, “Tunneling of obliquely-incident waves through PT-symmetric epsilon-near-zero bi-layers,” Phys. Rev. B 89(8), 085105 (2014). [CrossRef]  

7. S. Savoia, G. Castaldi, V. Galdi, A. Alù, and N. Engheta, “PT-symmetry-induced wave confinement and guiding in ɛ-near-zero metamaterials,” Phys. Rev. B 91(11), 115114 (2015). [CrossRef]  

8. Ş. K. Özdemir, S. Rotter, F. Nori, and L. Yang, “Parity–time symmetry and exceptional points in photonics,” Nat. Mater. 18(8), 783–798 (2019). [CrossRef]  

9. S. Longhi, “PT-symmetric laser absorber,” Phys. Rev. A 82(3), 031801 (2010). [CrossRef]  

10. Y. D. Chong, L. Ge, and A. D. Stone, “PT-symmetry breaking and laser-absorber modes in optical scattering systems,” Phys. Rev. Lett. 106(9), 093902 (2011). [CrossRef]  

11. Y. Sun, W. Tan, H. Q. Li, J. Li, and H. Chen, “Experimental demonstration of a Coherent perfect absorber with PT phase transition,” Phys. Rev. Lett. 112(14), 143903 (2014). [CrossRef]  

12. J. Schindler, Z. Lin, J. M. Lee, H. Ramezani, F. M. Ellis, and T. Kottos, “PT−symmetric electronics,” J. Phys. A: Math. Theor. 45(44), 444029 (2012). [CrossRef]  

13. D. G. Baranov, A. Krasnok, T. Shegai, A. Alù, and Y. Chong, “Coherent perfect absorbers: linear control of light with light,” Nat. Rev. Mater. 2(12), 17064 (2017). [CrossRef]  

14. C. M. Bender and S. Boettcher, “Real spectra in non-hermitian hamiltonians having PT symmetry,” Phys. Rev. Lett. 80(24), 5243–5246 (1998). [CrossRef]  

15. X. Zhu, H. Ramezani, C. Shi, J. Zhu, and X. Zhang, “PT-symmetric acoustics,” Phys. Rev. X 4(3), 031042 (2014). [CrossRef]  

16. M. Li and T. Xu, “Dark and antidark soliton interactions in the nonlocal nonlinear Schr ¨odinger equation with the self-induced parity-time-symmetric potential,” Phys. Rev. E 91(3), 033202 (2015). [CrossRef]  

17. R. Driben and B. A Malomed, “Stability of solitons in parity–time-symmetric couplers,” Opt. Lett. 36(22), 4323–4325 (2011). [CrossRef]  

18. J. Schindler, A. Li, M.C. Zheng, F.M. Ellis, and T. Kottos, “Experimental study of active LRC circuits with PT symmetries,” Phys. Rev. A 84(4), 040101 (2011). [CrossRef]  

19. P. Y. Chen, M. Sakhdari, M. Hajizadegan, Q. Cui, M. Cheng, R. El-Ganainy, and A. A. Alù, “Generalized parity-time symmetry condition for enhanced sensor telemetry,” Nat. Electron. 1(5), 297–304 (2018). [CrossRef]  

20. X. Yin and X. Zhang, “Unidirectional light propagation at exceptional points,” Nat. Mater. 12(3), 175–177 (2013). [CrossRef]  

21. R. Fleury, D.L. Sounas, and A. Alù, “Negative refraction and planar focusing based on parity-time symmetric metasurfaces,” Phys. Rev. Lett. 113(2), 023903 (2014). [CrossRef]  

22. R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, “Non-Hermitian physics and PT symmetry,” Nat. Phys. 14(1), 11–19 (2018). [CrossRef]  

23. Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, “Unidirectional invisibility induced by PT-symmetric periodic structures,” Phys. Rev. Lett. 106(21), 213901 (2011). [CrossRef]  

24. M. Sakhdari, N. Mohammadi Estakhri, H. Bagci, and P. Y. Chen, “Low-threshold lasing and coherent perfect absorption in generalized PT-symmetric optical structures,” Phys. Rev. Appl. 10(2), 024030 (2018). [CrossRef]  

25. M. Hajizadegan, M. Sakhdari, S. Liao, and P.-Y. Chen, “High-sensitivity wireless displacement sensing enabled by PT-symmetric telemetry,” IEEE Trans. Antennas Propag. 67(5), 3445–3449 (2019). [CrossRef]  

26. H. Hodaei, A. U. Hassan, S. Wittek, H. Garcia-Gracia, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, “Enhanced sensitivity at higher-order exceptional points,” Nature 548(7666), 187–191 (2017). [CrossRef]  

27. Z. Dong, Z. Li, F. Yang, C. W. Qiu, and J. S. Ho, “Sensitive readout of implantable microsensors using a wireless system locked to an exceptional point,” Nat. Electron. 2(8), 335–342 (2019). [CrossRef]  

28. M. Hajizadegan, L. Zhu, and P. Y. Chen, “Superdirective leaky radiation from a PT-synthetic metachannel,” Opt. Express 29(8), 12330–12343 (2021). [CrossRef]  

29. Z. Ye, M. Yang, L. Zhu, and P.-Y. Chen, “PTX-symmetric metasurfaces for sensing applications,” Frontiers Optoelectron. 14(2), 211–220 (2021). [CrossRef]  

30. H. Shen, B. Zhen, and L. Fu, “Topological band theory for non-Hermitian Hamiltonians,” Phys. Rev. Lett. 120(14), 146402 (2018). [CrossRef]  

31. M. Yang, Z. Ye, M. Farhat, and P.-Y. Chen, “Ultrarobust wireless interrogation for sensors and transducers: a non- Hermitian telemetry technique,” IEEE Trans. Instrum. Meas. 70, 1–13 (2021). [CrossRef]  

32. A. Alù, M. G. Silveirinha, A. Salandrino, and N. Engheta, “Epsilon-near-zero metamaterials and electromagnetic sources: tailoring the radiation phase pattern,” Phys. Rev. B 75(15), 155410 (2007). [CrossRef]  

33. C. Argyropoulos, P. Y. Chen, G. D’Aguanno, N. Engheta, and A. Alu, “Boosting optical nonlinearities in ε-near-zero plasmonic channels,” Phys. Rev. B 85(4), 045129 (2012). [CrossRef]  

34. G. Lovat, P. Burghignoli, F. Capolino, D. R. Jackson, and D. R. Wilton, “Analysis of directive radiation from a line source in a metamaterial slab with low permittivity,” IEEE Trans. Antennas Propag. 54(3), 1017–1030 (2006). [CrossRef]  

35. D. M. Pozar, Microwave Engineering, 4th edition (Wiley, 2012).

36. J. Lee, M. Tymchenko, C. Argyropoulos, P. Y. Chen, F. Lu, F. Demmerle, G. Boehm, M. C. Amann, A. Alù, and M. A. Belkin, “Giant nonlinear response from plasmonic metasurfaces coupled to intersubband transitions,” Nature 511(7507), 65–69 (2014). [CrossRef]  

37. C. A. Balanis, Antenna Theory: Analysis and Design (Wiley, 2016).

38. D. R. Jackson and A. A. Oliner, “A leaky-wave analysis of the high-gain printed antenna configuration,” IEEE Trans. Antennas Propag. 36(7), 905–910 (1988). [CrossRef]  

39. T. Low, P. Y. Chen, and D. N. Basov, “Superluminal plasmons with resonant gain in population inverted bilayer graphene,” Phys. Rev. B 98(4), 041403 (2018). [CrossRef]  

40. T. J. Guo, L. Zhu, P. Y. Chen, and C. Argyropoulos, “Tunable terahertz amplification based on photoexcited active graphene hyperbolic metamaterials,” Opt. Mater. Express 8(12), 3941–3952 (2018). [CrossRef]  

41. F. Monticone, C. A. Valagiannopoulos, and A. Alù, “Parity-time symmetric nonlocal metasurfaces: all-angle negative refraction and volumetric imaging,” Phys. Rev. X 6(4), 041018 (2016). [CrossRef]  

42. P. Y. Chen, C. Argyropoulos, and A. Alu, “Broadening the cloaking bandwidth with non-foster metasurfaces,” Phys. Rev. Lett. 111(23), 233001 (2013). [CrossRef]  

43. Z. Li, G. Cao, C. Li, S. Dong, Y. Deng, X. Liu, J. S. Ho, and C. W. Qiu, “Non-Hermitian electromagnetic metasurfaces at exceptional points (invited review),” Prog. Electromagn. Res. 171, 1–20 (2021). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. (a) Schematics of PTX-symmetric metasurfaces composed of an active metasurface (-R2) and a passive metasurface (R1). The PTX-symmetric metasurfaces can be regarded as a metachannel having a longitudinal propagation constant $\beta $, varied between zero and the wavenumber of the background medium 1, resulting in an effective permittivity, $0 \le {\varepsilon _{{eff}}} \le \varepsilon .$ (b) Transmission-line network model for scattering of plane waves by PTX-symmetric metasurfaces, and (c) radiation from a line source inside PTX-symmetric metasurface.
Fig. 2.
Fig. 2. Evolution of two eigenvalues of the scattering matrix of the PTX-symmetric metasurfaces shown in Fig. 1(b).
Fig. 3.
Fig. 3. Dispersion relations for the PTX-symmetric metachannel when the electrical length between two metasurface is x = 2 (black), x = 1 (red), and x = 0.5 (green) in Fig. 1 under the CPAL condition (Eqs. (2) and (3) with $\gamma = 1/\sqrt 2$) with (a)$\alpha = 0$ (b)$\alpha = \pi /6$ and (c) $\alpha = \pi /3$; here, the solid and dashed lines represents the real and imaginary parts, respectively.
Fig. 4.
Fig. 4. Far-field radiation pattern and electric field distribution for PTX metasurfaces excited by a line source, which is operated at the CPAL condition with (a)$\alpha = 0$ (b)$\alpha = \pi /6$ and (c) $\alpha = \pi /3$.
Fig. 5.
Fig. 5. (a) Far-field radiation pattern and electric field distribution for PTX metasurfaces excited by a line source, which is operated at the CPAL condition with $\alpha = 0$ and different values of κ. (b) Similar to (a), but for α = π/4.

Equations (16)

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

Z s ( y ) = ( κ Z ) δ ( y + d / 2 ) + ( Z / κ ) δ ( y d / 2 )
Z ( y ) = { Z / κ if y > d / 2 Z if d / 2 < y < d / 2 κ Z if y < d / 2 ,
ε 1 = ε 1 + cot 2 α 1 + κ 2 cot 2 α
ε 2 = ε 1 + κ 4 cot 2 α 1 + κ 2 cot 2 α
R 1 = κ γ 1 + tan 2 α / κ 2 η
R 2 = ( 1 / κ ) γ 1 + tan 2 α / κ 2 η
d = 1 + tan 2 α / κ 2 4 λ 0
( 1 R 2 + k 2 2 β 2 ω μ ) 1 + 1 + j ( ω μ R 1 k 2 β 2 + k 1 2 β 2 k 2 β 2 ) tan ( k 2 β 2 d ) 1 R 1 + k 1 2 β 2 ω μ + j k 2 β 2 ω μ tan ( k 2 β 2 d ) = 0
E z ( x , y ) = 1 2 π + E ~ z ( k i , x ) e j ( k i , x x + k i , y y ) d k i , x
E ~ z ( + ) ( k x ) = I g Z e j β d 2 ( e j β d + Γ 1 ) ( 1 + Γ 2 ) 2 ( e 2 j β d Γ 1 Γ 2 )
Γ i = ( Z i / / R i ) Z ( Z i / / R i ) + Z
E ~ z ( ) ( k x ) = I g Z e j β d 2 ( 1 + Γ 1 ) ( e j β d + Γ 2 ) 2 ( e 2 j β d Γ 1 Γ 2 )
E z ( ± ) ( ρ , θ ) = E z f f , ( ± ) ( θ ) e j k i ρ ρ
E z f f , ( + ) ( θ ) = cos θ j k 2 2 π E ~ z ( + ) ( k 2 sin θ )
E z f f , ( ) ( θ ) = cos θ j k 1 2 π E ~ z ( ) ( k 1 sin θ )
P ( ± ) ( θ ) = | E z f f , ( ± ) ( θ ) | 2 2 η
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.