Next Article in Journal
Tyrosinase Immobilization Strategies for the Development of Electrochemical Biosensors—A Review
Previous Article in Journal
Selected Area Deposition of High Purity Gold for Functional 3D Architectures
Previous Article in Special Issue
Low Loss Vertical TiO2/Polymer Hybrid Nano-Waveguides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Integrated Optical Filters with Hyperbolic Metamaterials

by
Mas-ud A. Abdulkareem
1,†,
Fernando López-Rayón
2,†,
Citlalli T. Sosa-Sánchez
3,
Ramsés E. Bautista González
4,
Maximino L. Arroyo Carrasco
2,
Marycarmen Peña-Gomar
1,
Victor Coello
3 and
Ricardo Téllez-Limón
5,*
1
Facultad de Ciencias Físico Matemáticas, Universidad Michoacana de San Nicolás de Hidalgo, Avenida Francisco J. Múgica s/n, Ciudad Universitaria, Morelia C. P. 58030, Michoacán, Mexico
2
Facultad de Ciencias Físico-Matemáticas, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y 18 Sur, San Manuel, Puebla C. P. 72570, Puebla, Mexico
3
Centro de Investigación Científica y de Educación Superior de Ensenada, Unidad Monterrey, Alianza Centro 504, PIIT, Apodaca C. P. 66629, Nuevo León, Mexico
4
School of Biological Sciences, The University of Adelaide, Adelaide, SA 5005, Australia
5
CONACYT—Centro de Investigación Científica y de Educación Superior de Ensenada, Unidad Monterrey, Alianza Centro 504, PIIT, Apodaca C. P. 66629, Nuevo León, Mexico
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(4), 759; https://doi.org/10.3390/nano13040759
Submission received: 25 January 2023 / Revised: 8 February 2023 / Accepted: 9 February 2023 / Published: 17 February 2023
(This article belongs to the Special Issue Nanophotonics and Integrated Optics Devices)

Abstract

:
The growing development of nanotechnology requires the design of new devices that integrate different functionalities at a reduced scale. For on-chip applications such as optical communications or biosensing, it is necessary to selectively transmit a portion of the electromagnetic spectrum. This function is performed by the so-called band-pass filters. While several plasmonic nanostructures of complex fabrication integrated to optical waveguides have been proposed, hyperbolic metamaterials remain almost unexplored for the design of integrated band-pass filters at optical wavelengths. By making use of the effective medium theory and finite integration technique, in this contribution we numerically study an integrated device consisting of a one-dimensional hyperbolic metamaterial placed on top of a photonic waveguide. The results show that the filling fraction, period, and number of layers modify the spectral response of the device, but not for type II and effective metal metamaterials. For the proposed Au-TiO2 multilayered system, the filter operates at a wavelength of 760 nm, spectral bandwidth of 100 nm and transmission efficiency above 40%. The designed devices open new perspectives for the development of integrated band-pass filters of small scale for on-chip integrated optics applications.

1. Introduction

Optical bandpass filters are optical devices that selectively transmit a portion of the electromagnetic spectrum while rejecting all other wavelengths. One of the main applications of these devices stands for optical communications, where optical fiber technology requires the transmission of specific bandwidths at given wavelengths. For many years, different photonic waveguides compatible with optical fibers have been designed to properly filter light signals [1,2,3,4,5,6,7]. Even with this development, several factors still hinder the practical use of these devices with current technologies, which require miniaturized functional photonic systems with more advanced and configurable filters with novel characteristics.
With the development of nanotechnology, new opportunities have opened up for the integration of artificially engineered subwavelength materials with enhanced properties not otherwise found in nature, so-called metamaterials [8,9,10], with photonic waveguides. Among the different structures integrated to waveguides for signal filtering that can be mentioned are dielectric and plasmonic ring resonators [11,12], gratings [13,14,15], nanodisk [16,17,18] and asymmetric resonators [19,20], nanostructured plasmonic waveguides [21,22], waveguide cladding modulators [23,24,25,26], and photonic crystals [27,28]. In a previous work, we experimentally demonstrated that a gold nanoslab placed on top of an ion-exchanged glass waveguide serves as a stop-band filter of light for a broad bandwidth at near infrared wavelengths [29].
In recent years, a new kind of metamaterials have attracted the interest of the research community due to their unusual anisotropic nature, the so-called hyperbolic metamaterials (HMM) [30,31,32,33,34]. This growing interest is because isotropic materials have a closed isofrequency surface that limits the wavenumber of the electromagnetic field propagating through these media. For HMM, an extreme anisotropy is induced, leading to higher wavenumbers values in a non-closed hyperbolic isofrequency surface [35,36]. One way to introduce this extreme anisotropy is by alternating dielectric and metallic thin layers [32,37,38,39]. For these one-dimensional periodic structures, intrinsic resonant modes arise from coupling of photonic modes and surface plasmon polaritons at the metal-dielectric interfaces, leading to hybrid photonic-plasmonic modes. If the wavelength and spectral bandwidth of these modes are too close, broad band resonances can take place [37,40]. These broad resonances have been used for the design of bulk bandpass filters operating at telecommunications [41], terahertz (THz) [42], and near infrared [43] wavelengths. Integrated band-pass filters have also been proposed at THz frequencies by using a composite of two different-sized tapered HMM waveguide arrays, with each waveguide operating at wide but different absorption and transmission bands [44].
However, the use of HMM for the development of band-pass filters integrated to optical waveguides operating at visible and near-infrared wavelengths has barely been explored. These spectral bands are of interest, for instance, for on-chip biosensing applications in the first and second biological windows [45]. In this contribution, we numerically explore the design of an integrated band-pass filter by making use of metallic-dielectric multilayered HMM, The structure, as depicted in Figure 1, consists of a Si3N4 multimode waveguide on top of which a finite periodic array of gold (Au) and titanium dioxide ( TiO 2 ) thin layers are placed. It is demonstrated that the transmission for the  TM 0  mode is filtered at a central wavelength  λ = 760  nm of bandwidth  Δ λ F W H M = 100  nm with a transmittance above  40 %  of incident light, when the multilayered system behaves as an effective metal or hyperbolic metamaterial type II [32], while for an effective dielectric metamaterial, the band-pass filtering can be tuned as a function of the period and number of layers. Due to the simplicity of the structure, the proposed devices open new perspectives for the development of size-reduced integrated optical filters.

2. Materials and Methods

2.1. Description of the Integrated System

The device under analysis consists of a finite-sized HMM placed on top of a dielectric photonic waveguide, as depicted in Figure 1.
The waveguide consists of a rectangular silicon nitride ( Si 3 N 4 ) core of width  w c = 750  nm, height  h c = 250  nm, and length  L c = 4.0  μm. This core of refractive index  n c = 2.016  was buried in a glass substrate of refractive index  n s u b = 1.5 . The superstrate was considered as air ( n s u p = 1.0 ). The dispersion curves of the dielectric waveguide and spatial distribution of the electric field of each mode are shown in Figure 2. The modes that can be propagated along the z direction of the waveguide in the spectral range from 500 nm to 1523 nm, have cut-off wavelengths  λ TE 0 = 1523  nm,  λ TM 0 = 1089  nm,  λ TE 1 = 901  nm,  λ TM 1 = 787  nm,  λ TE 2 = 654  nm, and  λ TM 2 = 611  nm.
The HMM has a width of  w H M M = 1.0  μm and length  L H M M = 2.0  μm, and it is constituted by a periodic array of N alternated thin layers of metal (Au) and dielectric ( TiO 2 ) materials, of thickness  t m  ad  t d , respectively. The period of the structure is  T = t d + t m , as shown in the inset of Figure 1. The HMM on top of the waveguide was centered with respect to the center of the core. The dielectric function of gold was calculated from the Drude–Lorentz model as described in Refs. [46,47], while the refractive index of  TiO 2  was taken from the refractive index database using Ref. [48].

2.2. Effective Medium Theory

The effective medium theory describes a system considering the properties of its constituents. For a metamaterial composed of multilayers, if the layers are thinner with respect to the wavelength, it is possible to consider all the multilayers as a whole system whose electrical response can be characterized by an effective permittivity.
We propose an HMM made by an infinite periodic array of ( TiO 2 ) and gold (Au) thin layers. Considering the constituent materials, filling factor, and using the effective medium theory for a multilayer system [35], the effective permittivity phase diagram (Figure 3) was obtained following the classification of Ref. [32], related to the positive or negative values of the effective permittivity components. If  ε x x  and  ε y y  have opposite signs, extreme anisotropy is achieved, giving rise to hyperbolic dispersion curves [30,31,32].
The effective permittivity phase diagram (Figure 3) classifies the effective medium according to the effective dielectric permittivity equation, as a function of the metal filling fraction and wavelength. The phase diagram was obtained using the effective medium theory, where the effective dielectric function for transverse magnetic polarization is given by Ref. [35]:
ϵ x x = ϵ z z = p ϵ m + ( 1 p ) ϵ d ,
ϵ y y = p ϵ m + 1 p ϵ d 1 ,
where  p = t m / T , is the metal filling fraction (portion of metal at each period). The metallic and dielectric layers have permittivities  ϵ m  and  ϵ d , respectively. Using Equations (1) and (2), we computed the phase diagram in Figure 3 varying the Au filling fraction from  p = 0.1  to  p = 1  for a spectral wavelength range from  λ = 500  nm to  λ = 1100  nm. We show different regions depending on the signs of  ϵ x x  and  ϵ y y TiO 2  and Au permittivities were also taken from Ref. [48] and Refs. [46,47], respectively.

2.3. Transfer Matrix Method

To compute the dispersion curves of the multilayered media, we used the transfer matrix method [49]. These curves quantify the number of modes supported by the periodic structure as a function of the propagation constant at a given spectral range. The obtained results for a system of  N = 12  layers (6 Au and 6  TiO 2  layers) with a filling fraction  p = 0.5  and period  T = 80  nm ( t m = t d = 40  nm) are shown in Figure 4a. The green curves represent the modes, the white dotted curve represents the air light-line, and the white dashed curve corresponds to the glass substrate light-line. The map was obtained by equating the fourth element of the general matrix to zero [49] and plotted in logarithmic scale, the maxima values being related to the modes supported by the structure.

2.4. Light Propagation in a 3D Integrated Device

To compute the transmission and reflection spectra of light at the output and input of the integrated system, we performed 3D simulations by means of the finite integration technique [50], using the commercial software CST Studio Suite 2020 (Dassault Systems, Vélizy-Villacoublay, France). For this purpose, first we computed the photonic modes supported by the dielectric waveguide and used the spatial distribution of their electromagnetic field and propagated them trough the integrated device. For the simulations, we used a computational window of width  w x = 3.0  μm, height  h y = 2.4  μm, and length  L z = 4.0  μm, surrounded by perfectly matched layers. The transmission and reflection signals were measured defining port monitors at the input and end of the waveguide.
Figure 4b shows the normalized transmission (red line) and reflection (blue line) curves when the integrated device was excited with the fundamental  TM 0  photonic mode. Two main broad minima bands are observed in the transmission spectrum, located at the spectral position of the broad modes supported by the HMM plotted in the dispersion curves.

3. Results

We firstly analyzed the dependence of the operation of the device as a function of light polarization. For this purpose, we propagated the fundamental  TE 0  and  TM 0  photonic modes through the waveguide. For the  TE 0  mode, the electric field is mainly oriented along the horizontal x direction, while for  TM 0 , the electric field is oriented along the vertical y direction [20]. For these simulations, we considered a system of  N = 8  layers (4 Au layers and 4  TiO 2  layers) with a filling fraction  p = 0.5  ( t m = 40  nm,  t d = 40  nm, period  T = 80  nm), on top of the  Si 3 N 4  dielectric waveguide.
The results are plotted in Figure 5, where normalized transmission and reflection curves for  TM 0  mode (red and blue continuous, respectively) and for the  TE 0  mode (red and blue dashed, respectively) are shown. Vertical lines correspond to the cut-off wavelengths for each mode supported by the waveguide in the spectral region from 550 to 1150 nm:  TM 0  mode has a cut-off wavelength  λ c , TM 0 = 1089  nm (black dashed),  TE 1  mode at  λ c , TE 1 = 901  nm (blue dot-dashed),  TM 1  mode at  λ c , TM 1 = 786  nm (magenta dotted),  TE 2  at  λ c , TE 2 = 655  nm (purple triangles), and  TM 2  at  λ c , TM 2 = 608  nm (green triangles). We must remark that no mode conversion was observed, and scattering losses are around  10 %  of incident light: the signal reduction for the  TM 0  mode is mainly because of the optical losses by absorption.
As observed in Figure 5, the transmission of the  TM 0  mode presents the two main broadband deeps centered around 680 nm ( Δ λ F W H M = 60  nm) and at 900 nm ( Δ λ F W H M = 150  nm). For the  TE 0  mode, these broad deeps disappear.
As the transmission signal was modified only for  TM 0  mode (vertical polarization), we studied the behavior of the transmission and reflection spectra in terms of the filling fraction, the number of layers, and the period of the structure. We first considered a fixed period  T = 80  nm for three filling fractions  p = [ 0.2 , 0.5 , 0.8 ]  and three values for the number of layers  N = [ 8 , 12 , 16 ]  (4, 6, and 8 pairs of Au- TiO 2  interfaces). The results are shown in Figure 6.
The principal observations from transmission (red curves) and reflection (blue curves) spectra of Figure 6 are as follows. For  N = 8  (Figure 6a–c), the transmission spectrum for  p = 0.2  exhibits minima at  λ = 638  nm (guided light transmittance of  2 % ),  λ = 880  nm (transmittance of  4 % ) and  λ = 999  nm (transmittance of  2 % ). For  p = 0.5 , two main broad-band deeps occur at 680 nm (transmittance of  7 % Δ λ F W H M = 80  nm) and at  λ = 908  nm (transmittance of  1 % Δ λ F W H M = 200  nm). For  p = 0.8 , two main broadband deeps also appear, centered at  λ = 680  nm (transmittance of  5 % Δ λ F W H M = 78  nm) and at  λ = 908  nm (transmittance of  2 % Δ λ F W H M = 190  nm). These two deeps generate a band-pass filter with a central wavelength around  λ = 760  nm,  Δ λ F W H M = 100  nm, and signal transmittance of  41 % .
For  N = 12  (Figure 6d–f), when  p = 0.2 , two local minima occur at  λ = 778  nm (transmission of  2 % Δ λ F W H M = 35  nm) and at  λ = 999  nm ( 0.9 %  transmittance,  Δ λ F W H M = 130  nm) and a transparency band is observed centered at  λ = 845  nm ( 30 %  transmittance,  Δ λ F W H M = 110  nm). For  p = 0.5  and  p = 0.8 , the transmission spectra are almost the same as for  N = 8 .
For  N = 16  (Figure 6g–i), if  p = 0.2 , four minima are observed centered at  λ = 638  nm ( 6 %  transmittance,  Δ λ F W H M = 20  nm),  λ = 713  nm ( 2 %  transmittance,  Δ λ F W H M = 40  nm),  λ = 810  nm ( 2 %  transmittance,  Δ λ F W H M = 30  nm), and  λ = 999  nm ( 0.6 %  transmittance,  Δ λ F W H M = 100  nm). For  p = 0.5  and  p = 0.8 , the transmission spectra remain, and, once again, are almost the same as for  N = 8  and  N = 12 .
We then computed the propagation of the  TM 0  mode considering a fixed number of layers  N = 8  (4 pairs of Au- TiO 2  interfaces) for filling fractions  p = [ 0.2 , 0.5 , 0.8 ]  and two periods of the layers  T = [ 50 , 80 ]  nm. The obtained transmission (red curves) and reflection (blue curves) spectra are shown in Figure 7.
For  T = 50  nm (Figure 7a–c), when  p = 0.2  two principal minima occur at  λ = 778  nm ( 0.5 %  transmittance,  Δ λ F W H M = 57  nm) and at  λ = 936  nm ( 1 %  transmittance,  Δ λ F W H M = 47  nm). For  p = 0.5 , two broadband deeps appear centered at  λ = 689  nm ( 9 %  transmittance,  Δ λ F W H M = 75  nm) and at  λ = 920  nm ( 0.9 %  transmittance,  Δ λ F W H M = 170  nm). For  p = 0.8 , two deeps appear centered at  λ = 680  nm ( 5 %  transmittance,  Δ λ F W H M = 65  nm) and at 908 nm ( 2 %  transmittance,  Δ λ F W H M = 150  nm). For  T = 80  nm, the spectra and values are the same as in Figure 6a–c.

4. Discussion

The obtained results show that the transmission spectrum of a dielectric waveguide can be filtered by placing a hyperbolic metamaterial consisting of periodically structured metallic (Au)-dielectric ( TiO 2 ) thin layers integrated on top of a dielectric ( Si 3 N 4 ) waveguide.
This optical integrated filter only operates if light is mainly polarized along the vertical y direction, a situation that can be achieved by propagating the  TM 0  mode of the photonic waveguide, as demonstrated in Figure 5. For this polarization, the electric field is symmetrically compatible for the excitation of surface plasmon polaritons at the dielectric-metallic interfaces [29].
As established by the effective medium theory (Equations (1) and (2)), the filling fraction of the multilayered system determines the behavior of the hyperbolic metamaterial (effective dielectric, effective metal of hyperbolic metamaterial types I or II), as shown in Figure 3. Hence, it is expected that the bands of modes supported by the hyperbolic metamaterial (see Figure 4, for instance) also depend on the number of layers (N). However, as demonstrated in Figure 6, the number of layers only modifies the spectral response (broadband shifting) of the integrated device when the metamaterial behaves as an effective dielectric material ( p = 0.2 ). When the multilayered system behaves as hyperbolic media type II or as effective metal, the number of layers does not significantly modify the central wavelength of two principal broadband resonances centered around  λ = 680  nm and  λ = 908  nm, with a transmittance of  5 %  and  2 % , respectively.
When the period of the multilayered structure was modified from  T = 80  nm to  T = 50  nm, it was also observed that for  p = 0.2 , different broad-band transmission minima arise and are spectrally shifted, while for  p = 0.5  and  p = 0.8  the two main broad-band deeps remain almost unchanged.
It is worth to mention that several small and narrow deeps also appear in transmission spectra. Most of them are due to plasmonic and hybrid photonic-plasmonic modes, which are hard to identify because the modes of the infinite multilayered system are too close to each other (see Figure 4a for instance). In addition, it is possible that some of these small deeps arise from photonic modes, because the hyperbolic metamaterial placed on top of the dielectric waveguide is finite and standing waves can also take place. However, for  p = 0.5  and  p = 0.8 , these perturbations are mounted in two main broadband deeps.
Even when the proposed structure does not present a high transmission efficiency, multilayered media on top of dielectric waveguides are easier to fabricate. For instance, photolithography combined with the thin layers deposition techniques, such as atomic layer deposition, sputtering, and even thermal evaporation, can be employed with high repeatability, being advantageous in comparison with plasmonic nanostructures of complex geometries. These results open up new perspectives in the design of optical integrated filters by making use of 1D hyperbolic metamaterials. Without the loss of generality, the combination of dielectric and metallic thin layers can be modified to tune the central wavelengths of the proposed integrated band-pass filters.

Author Contributions

Conceptualization, M.-u.A.A., F.L.-R., V.C. and R.T.-L.; investigation, M.-u.A.A., F.L.-R. and C.T.S.-S.; methodology, R.T.-L. and M.L.A.C.; software, R.E.B.G., F.L.-R., C.T.S.-S. and R.T.-L.; formal analysis, M.-u.A.A., F.L.-R., R.E.B.G., C.T.S.-S. and R.T.-L.; writing, M.-u.A.A., C.T.S.-S., M.L.A.C., M.P.-G., V.C. and R.T.-L.; supervision, M.L.A.C., M.P.-G., V.C. and R.T.-L.; project administration, V.C. and R.T.-L.; funding acquisition R.T.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This project was partially funded by the “Research fund for education” (CONACYT—Basic Scientific Research, grant No. A1-S-21527). F.L.-R. and M.-u.A.A. thank CONACYT for scholarship grants No. 848883 and No. 624595, respectively. C.T.S.-S. also thanks CONACYT for research scholarship grant No. I1200/320/2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

R.T.-L. thanks to Eugenio R. Méndez Méndez (Optics Department, CICESE) for having enriched the content of this work with his comments.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TMTransverse Magnetic
TETransverse Electric
Δ λ Wavelength spectral bandwidth
FWHMFull width at half maximum

References

  1. Kogelnik, H. An Introduction to Integrated Optics. IEEE Trans. Microw. Theory Tech. 1975, 23, 2–16. [Google Scholar] [CrossRef]
  2. Suhara, T.; Nishihara, H. Integrated optics components and devices using periodic structures. IEEE J. Quantum Electron. 1986, 22, 845–867. [Google Scholar] [CrossRef] [Green Version]
  3. Okamoto, K. Recent progress of integrated optics planar lightwave circuits. Opt. Quantum Electron. 1999, 31, 107–129. [Google Scholar] [CrossRef]
  4. Ma, H.; Jen, A.Y.; Dalton, L. Polymer-Based Optical Waveguides: Materials, Processing, and Devices. Adv. Mater. 2002, 14, 1339–1365. [Google Scholar] [CrossRef]
  5. Broquin, J.E. Glass integrated optics: State of the art and position toward other technologies. In Proceedings of the Integrated Optics: Devices, Materials, and Technologies XI, San Jose, CA, USA, 20–25 January 2007; Sidorin, Y., Waechter, C.A., Eds.; International Society for Optics and Photonics, SPIE: Bellingham, WA, USA, 2007; Volume 6475, p. 647507. [Google Scholar] [CrossRef]
  6. Sohler, W.; Hu, H.; Ricken, R.; Quiring, V.; Vannahme, C.; Herrmann, H.; Büchter, D.; Reza, S.; Grundkötter, W.; Orlov, S.; et al. Integrated Optical Devices in Lithium Niobate. Opt. Photon. News 2008, 19, 24–31. [Google Scholar] [CrossRef]
  7. Tong, X.C. Advanced Materials for Integrated Optical Waveguides; Springer: Cham, Switzerland, 2013; p. 552. [Google Scholar]
  8. Zheludev, N.I.; Kivshar, Y.S. From metamaterials to metadevices. Nat. Mater. 2012, 19, 917–924. [Google Scholar] [CrossRef]
  9. Urbas, A.M.; Jacob, Z.; Negro, L.D.; Engheta, N.; Boardman, A.D.; Egan, P.; Khanikaev, A.B.; Menon, V.; Ferrera, M.; Kinsey, N.; et al. Roadmap on optical metamaterials. J. Opt. 2016, 18, 093005. [Google Scholar] [CrossRef] [Green Version]
  10. Liang, Y.; Koshelev, K.; Zhang, F.; Lin, H.; Lin, S.; Wu, J.; Jia, B.; Kivshar, Y. Bound States in the Continuum in Anisotropic Plasmonic Metasurfaces. Nano Lett. 2020, 20, 6351–6356. [Google Scholar] [CrossRef]
  11. Rabus, D.G.; Sada, C. Integrated Ring Resonators, 2nd ed.; Springer: Cham, Switzerland, 2020; p. 360. [Google Scholar]
  12. Holmgaard, T.; Chen, Z.; Bozhevolnyi, S.I.; Markey, L.; Dereux, A. Dielectric-loaded plasmonic waveguide-ring resonators. Opt. Express 2009, 17, 2968–2975. [Google Scholar] [CrossRef]
  13. Pérez-Galacho, D.; Alonso-Ramos, C.; Mazeas, F.; Roux, X.L.; Oser, D.; Zhang, W.; Marris-Morini, D.; Labonté, L.; Tanzilli, S.; Cassan, É.; et al. Optical pump-rejection filter based on silicon sub-wavelength engineered photonic structures. Opt. Lett. 2017, 42, 1468–1471. [Google Scholar] [CrossRef] [Green Version]
  14. Čtyroký, J.; Wangüemert-Pérez, J.G.; Kwiecien, P.; Richter, I.; Litvik, J.; Schmid, J.H.; Íñigo Molina-Fernández, I.; Moñux, A.O.M.; Dado, M.; Cheben, P. Design of narrowband Bragg spectral filters in subwavelength grating metamaterial waveguides. Opt. Express 2018, 26, 179–194. [Google Scholar] [CrossRef] [Green Version]
  15. Quaranta, G.; Basset, G.; Martin, O.J.F.; Gallinet, B. Recent Advances in Resonant Waveguide Gratings. Laser Photonics Rev. 2018, 12, 1800017. [Google Scholar] [CrossRef]
  16. Matsko, A.; Ilchenko, V. Optical resonators with whispering-gallery modes-part I: Basics. IEEE J. Sel. Top. Quantum Electron. 2006, 12, 3–14. [Google Scholar] [CrossRef]
  17. Lu, H.; Liu, X.; Mao, D.; Wang, L.; Gong, Y. Tunable band-pass plasmonic waveguide filters with nanodisk resonators. Opt. Express 2010, 18, 17922–17927. [Google Scholar] [CrossRef]
  18. Khani, S.; Danaie, M.; Rezaei, P. Realization of single-mode plasmonic bandpass filters using improved nanodisk resonators. Opt. Commun. 2018, 420, 147–156. [Google Scholar] [CrossRef]
  19. Tao, J.; Huang, X.G.; Lin, X.; Zhang, Q.; Jin, X. A narrow-band subwavelength plasmonic waveguide filter with asymmetrical multiple-teeth-shaped structure. Opt. Express 2009, 17, 13989–13994. [Google Scholar] [CrossRef]
  20. López-Rayón, F.; Arroyo Carrasco, M.L.; Rodríguez-Beltrán, R.I.; Salas-Montiel, R.; Téllez-Limón, R. Plasmonic-Induced Transparencies in an Integrated Metaphotonic System. Nanomaterials 2022, 12, 1701. [Google Scholar] [CrossRef]
  21. Lin, X.S.; Huang, X.G. Tooth-shaped plasmonic waveguide filters with nanometeric sizes. Opt. Lett. 2008, 33, 2874–2876. [Google Scholar] [CrossRef] [Green Version]
  22. Neutens, P.; Lagae, L.; Borghs, G.; Dorpe, P.V. Plasmon filters and resonators in metal-insulator-metal waveguides. Opt. Express 2012, 20, 3408–3423. [Google Scholar] [CrossRef] [Green Version]
  23. Tan, D.T.H.; Ikeda, K.; Fainman, Y. Cladding-modulated Bragg gratings in silicon waveguides. Opt. Lett. 2009, 34, 1357–1359. [Google Scholar] [CrossRef]
  24. Yun, H.; Hammood, M.; Chrostowski, L.; Jaeger, N.A.F. Optical Add-drop Filters using Cladding-modulated Sub-wavelength Grating Contra-directional Couplers for Silicon-on-Insulator Platforms. In Proceedings of the 2019 IEEE 10th Annual Information Technology, Electronics and Mobile Communication Conference (IEMCON), Vancouver, BC, Canada, 17–19 October 2019; pp. 0926–0931. [Google Scholar] [CrossRef]
  25. Yen, T.H.; Shih, B.H.; Cheng, N.W.; Hung, Y.J. Linewidth-adjustable bandpass filter based on silicon cladding-modulated waveguide moiré Bragg gratings. In Proceedings of the Conference on Lasers and Electro-Optics, San Jose, CA, USA, 13–18 May 2018; p. JW2A.38. [Google Scholar] [CrossRef]
  26. Yen, T.H.; Wu, C.J.; Yu, C.J.; Hung, Y.J. Silicon photonics multi-channel Bragg reflectors based on narrowband cladding-modulated gratings. In Proceedings of the 2017 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 14–19 May 2017; pp. 1–2. [Google Scholar]
  27. Jin, Y.; Fernez, N.; Pennec, Y.; Bonello, B.; Moiseyenko, R.P.; Hémon, S.; Pan, Y.; Djafari-Rouhani, B. Tunable waveguide and cavity in a phononic crystal plate by controlling whispering-gallery modes in hollow pillars. Phys. Rev. B 2016, 93, 054109. [Google Scholar] [CrossRef] [Green Version]
  28. Mendez-Astudillo, M.; Okayama, H.; Nakajima, H. Silicon optical filter with transmission peaks in wide stopband obtained by anti-symmetric photonic crystal with defect in multimode waveguides. Opt. Express 2018, 26, 1841–1850. [Google Scholar] [CrossRef] [PubMed]
  29. Tellez-Limon, R.; Blaize, S.; Gardillou, F.; Coello, V.; Salas-Montiel, R. Excitation of surface plasmon polaritons in a gold nanoslab on ion-exchanged waveguide technology. Appl. Opt. 2020, 59, 572–578. [Google Scholar] [CrossRef] [PubMed]
  30. Poddubny, A.; Iorsh, I.; Belov, P.; Kivshar, Y. Hyperbolic metamaterials. Nat. Photonics 2013, 7, 948–957. [Google Scholar] [CrossRef]
  31. Shekhar, P.; Atkinson, J.; Jacob, Z. Hyperbolic metamaterials: Fundamentals and applications. Nano Converg. 2014, 1, 1–14. [Google Scholar] [CrossRef] [Green Version]
  32. Ferrari, L.; Wu, C.; Lepage, D.; Zhang, X.; Liu, Z. Hyperbolic metamaterials and their applications. Prog. Quantum Electron. 2015, 40, 1–40. [Google Scholar] [CrossRef]
  33. Dudek, M.; Kowerdziej, R.; Pianelli, A.; Parka, J. Graphene-based tunable hyperbolic microcavity. Sci. Rep. 2021, 11, 74. [Google Scholar] [CrossRef]
  34. Pianelli, A.; Caligiuri, V.; Dudek, M.; Kowerdziej, R.; Chodorow, U.; Sielezin, K.; De Luca, A.; Caputo, R.; Parka, J. Active control of dielectric singularities in indium-tin-oxides hyperbolic metamaterials. Sci. Rep. 2022, 12, 16961. [Google Scholar] [CrossRef]
  35. Rytov, S. Electromagnetic properties of a finely stratified medium. Sov. Phys. Jept 1956, 2, 466–475. [Google Scholar]
  36. Drachev, V.P.; Podolskiy, V.A.; Kildishev, A.V. Hyperbolic metamaterials: New physics behind a classical problem. Opt. Express 2013, 21, 15048–15064. [Google Scholar] [CrossRef]
  37. Kidwai, O.; Zhukovsky, S.V.; Sipe, J.E. Effective-medium approach to planar multilayer hyperbolic metamaterials: Strengths and limitations. Phys. Rev. A 2012, 85, 053842. [Google Scholar] [CrossRef]
  38. Tumkur, T.; Barnakov, Y.; Kee, S.T.; Noginov, M.A.; Liberman, V. Permittivity evaluation of multilayered hyperbolic metamaterials: Ellipsometry vs. reflectometry. J. Appl. Phys. 2015, 117, 103104. [Google Scholar] [CrossRef] [Green Version]
  39. Hu, S.; Du, S.; Li, J.; Gu, C. Multidimensional Image and Beam Splitter Based on Hyperbolic Metamaterials. Nano Lett. 2021, 21, 1792–1799. [Google Scholar] [CrossRef] [PubMed]
  40. Zhukovsky, S.V.; Orlov, A.A.; Babicheva, V.E.; Lavrinenko, A.V.; Sipe, J.E. Photonic-band-gap engineering for volume plasmon polaritons in multiscale multilayer hyperbolic metamaterials. Phys. Rev. A 2014, 90, 013801. [Google Scholar] [CrossRef] [Green Version]
  41. Kalusniak, S.; Orphal, L.; Sadofev, S. Demonstration of hyperbolic metamaterials at telecommunication wavelength using Ga-doped ZnO. Opt. Express 2015, 23, 32555–32560. [Google Scholar] [CrossRef] [PubMed]
  42. Rizza, C.; Ciattoni, A.; Spinozzi, E.; Columbo, L. Terahertz active spatial filtering through optically tunable hyperbolic metamaterials. Opt. Lett. 2012, 37, 3345–3347. [Google Scholar] [CrossRef] [Green Version]
  43. Naik, G.V.; Liu, J.; Kildishev, A.V.; Shalaev, V.M.; Boltasseva, A. Demonstration of Al:ZnO as a plasmonic component for near-infrared metamaterials. Proc. Natl. Acad. Sci. USA 2012, 109, 8834–8838. [Google Scholar] [CrossRef] [Green Version]
  44. Zhou, X.; Yin, X.; Zhang, T.; Chen, L.; Li, X. Ultrabroad terahertz bandpass filter by hyperbolic metamaterial waveguide. Opt. Express 2015, 23, 11657–11664. [Google Scholar] [CrossRef]
  45. Hemmer, E.; Benayas, A.; Légaré, F.; Vetrone, F. Exploiting the biological windows: Current perspectives on fluorescent bioprobes emitting above 1000 nm. Nanoscale Horizons 2016, 13, 168–184. [Google Scholar] [CrossRef]
  46. Vial, A.; Grimault, A.S.; Macías, D.; Barchiesi, D.; de la Chapelle, M.L. Improved analytical fit of gold dispersion: Application to the modeling of extinction spectra with a finite-difference time-domain method. Phys. Rev. B 2005, 71, 085416. [Google Scholar] [CrossRef]
  47. Barchiesi, D.; Grosges, T. Fitting the optical constants of gold, silver, chromium, titanium, and aluminum in the visible bandwidth. J. Nanophotonics 2014, 8, 083097, Errata in J. Nanophotonics 2015, 8, 089996. [Google Scholar] [CrossRef] [Green Version]
  48. DeVore, J.R. Refractive Indices of Rutile and Sphalerite. J. Opt. Soc. Am. 1951, 41, 416–419. [Google Scholar] [CrossRef]
  49. Téllez-Limón, R.; Salas-Montiel, R. Nanowires Integrated to Optical Waveguides. In Nanowires; Peng, X., Ed.; IntechOpen: Rijeka, Croatia, 2021; Chapter 8. [Google Scholar] [CrossRef]
  50. Wittig, T.; Schuhmann, R.; Weiland, T. Model order reduction for large systems in computational electromagnetics. Linear Algebra Appl. 2006, 415, 499–530. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic of the integrated device. Hyperbolic metamaterial consisting of a periodic array of Au- TiO 2  thin layers of thickness  t m  and  t d , respectively, are placed on top of a  Si 3 N 4  waveguide ( w c = 750  nm,  h c = 250  nm) buried in a glass substrate. Photonic modes propagate through the waveguide along the z direction from the input (IN) and the transmission spectrum is measured at the output (OUT) face of the waveguide.
Figure 1. Schematic of the integrated device. Hyperbolic metamaterial consisting of a periodic array of Au- TiO 2  thin layers of thickness  t m  and  t d , respectively, are placed on top of a  Si 3 N 4  waveguide ( w c = 750  nm,  h c = 250  nm) buried in a glass substrate. Photonic modes propagate through the waveguide along the z direction from the input (IN) and the transmission spectrum is measured at the output (OUT) face of the waveguide.
Nanomaterials 13 00759 g001
Figure 2. Modes guided by the waveguide. (a) Dispersion curves of the dielectric  Si 3 N 4  waveguide in the spectral range from 500 nm to 1100 nm. The maps show the  | E |  field distribution and electric field lines, computed at  λ = 550  nm, of the (b TE 0 , (c TM 0 , (d TE 1 , (e TM 1 , (f TE 2 , and (g TM 2  modes launched at the input of the waveguide.
Figure 2. Modes guided by the waveguide. (a) Dispersion curves of the dielectric  Si 3 N 4  waveguide in the spectral range from 500 nm to 1100 nm. The maps show the  | E |  field distribution and electric field lines, computed at  λ = 550  nm, of the (b TE 0 , (c TM 0 , (d TE 1 , (e TM 1 , (f TE 2 , and (g TM 2  modes launched at the input of the waveguide.
Nanomaterials 13 00759 g002
Figure 3. Phase diagram of the metamaterial composed by layers of Au/ TiO 2  as a function of the filling fraction and wavelength. The first region in blue (lower left corner) corresponds to an effective dielectric behavior when  ϵ x x > 0  and  ϵ y y > 0 . The second region in orange (upper left corner) corresponds to an effective metal when  ϵ x x < 0  and  ϵ y y < 0 . The third region in white corresponds to a Type I HMM for which  ϵ x x > 0  and  ϵ y y < 0  and the fourth region in green corresponds to a Type II HMM for which  ϵ x x < 0  and  ϵ y y > 0 . Dotted lines point out the behavior of the multilayerd system for  p = 0.2  (yellow),  p = 0.5  (red) and  p = 0.8  (blue).
Figure 3. Phase diagram of the metamaterial composed by layers of Au/ TiO 2  as a function of the filling fraction and wavelength. The first region in blue (lower left corner) corresponds to an effective dielectric behavior when  ϵ x x > 0  and  ϵ y y > 0 . The second region in orange (upper left corner) corresponds to an effective metal when  ϵ x x < 0  and  ϵ y y < 0 . The third region in white corresponds to a Type I HMM for which  ϵ x x > 0  and  ϵ y y < 0  and the fourth region in green corresponds to a Type II HMM for which  ϵ x x < 0  and  ϵ y y > 0 . Dotted lines point out the behavior of the multilayerd system for  p = 0.2  (yellow),  p = 0.5  (red) and  p = 0.8  (blue).
Nanomaterials 13 00759 g003
Figure 4. (a) Dispersion curves for a HMM of 6 Au and 6  TiO 2  layers with a filling fraction  p = 0.5  and period  T = 80  nm. Dotted and dashed curves represent air and glass light-lines, respectively. (b) Normalized transmission (red solid) and reflection (blue dashed) spectra (normalized units) for an integrated system with a finite HMM ( N = 12 p = 0.5 , and  T = 80  nm) integrated on top of a dielectric waveguide. Several modes in the dispersion curves are associated in the main bands corresponding to the broad-band minima in the normalized transmission spectrum (shaded regions).
Figure 4. (a) Dispersion curves for a HMM of 6 Au and 6  TiO 2  layers with a filling fraction  p = 0.5  and period  T = 80  nm. Dotted and dashed curves represent air and glass light-lines, respectively. (b) Normalized transmission (red solid) and reflection (blue dashed) spectra (normalized units) for an integrated system with a finite HMM ( N = 12 p = 0.5 , and  T = 80  nm) integrated on top of a dielectric waveguide. Several modes in the dispersion curves are associated in the main bands corresponding to the broad-band minima in the normalized transmission spectrum (shaded regions).
Nanomaterials 13 00759 g004
Figure 5. Polarization dependence of transmitted signal. For  TM 0  mode (vertical polarization), the normalized transmission spectrum exhibits two broad deeps due to the excitation of modes in the hyperbolic metamaterial. For  TE 0  mode (horizontal polarization), no deeps are observed as no SPP are excited in the metamaterial.
Figure 5. Polarization dependence of transmitted signal. For  TM 0  mode (vertical polarization), the normalized transmission spectrum exhibits two broad deeps due to the excitation of modes in the hyperbolic metamaterial. For  TE 0  mode (horizontal polarization), no deeps are observed as no SPP are excited in the metamaterial.
Nanomaterials 13 00759 g005
Figure 6. Dependence of the broad deeps as a function of the number of layers (N) and filling fraction (p). (ac N = 8 , (df N = 12 , and (gi N = 16  for  p = [ 0.2 , 0.5 , 0.8 ] , respectively. For  p = 0.2 , the number of deeps and their spectral position depends on the number of layers. For  p = 0.5 , 0.8 , the broad deeps remain almost the same.
Figure 6. Dependence of the broad deeps as a function of the number of layers (N) and filling fraction (p). (ac N = 8 , (df N = 12 , and (gi N = 16  for  p = [ 0.2 , 0.5 , 0.8 ] , respectively. For  p = 0.2 , the number of deeps and their spectral position depends on the number of layers. For  p = 0.5 , 0.8 , the broad deeps remain almost the same.
Nanomaterials 13 00759 g006
Figure 7. Dependence of broad-band deeps as a function of the period (T) and filling fraction (p) for a fixed number of layers ( N = 8  layers). (ac T = 50  nm, and (df T = 80  nm, for  p = [ 0.2 , 0.5 , 0.8 ] , respectively. For  p = 0.2  (a,d), transmission (red solid) and reflection (blue dashed) spectra are modified, while for  p = 0.5  and  p = 0.8 , they remain almost unchanged.
Figure 7. Dependence of broad-band deeps as a function of the period (T) and filling fraction (p) for a fixed number of layers ( N = 8  layers). (ac T = 50  nm, and (df T = 80  nm, for  p = [ 0.2 , 0.5 , 0.8 ] , respectively. For  p = 0.2  (a,d), transmission (red solid) and reflection (blue dashed) spectra are modified, while for  p = 0.5  and  p = 0.8 , they remain almost unchanged.
Nanomaterials 13 00759 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abdulkareem, M.-u.A.; López-Rayón, F.; Sosa-Sánchez, C.T.; Bautista González, R.E.; Arroyo Carrasco, M.L.; Peña-Gomar, M.; Coello, V.; Téllez-Limón, R. Integrated Optical Filters with Hyperbolic Metamaterials. Nanomaterials 2023, 13, 759. https://doi.org/10.3390/nano13040759

AMA Style

Abdulkareem M-uA, López-Rayón F, Sosa-Sánchez CT, Bautista González RE, Arroyo Carrasco ML, Peña-Gomar M, Coello V, Téllez-Limón R. Integrated Optical Filters with Hyperbolic Metamaterials. Nanomaterials. 2023; 13(4):759. https://doi.org/10.3390/nano13040759

Chicago/Turabian Style

Abdulkareem, Mas-ud A., Fernando López-Rayón, Citlalli T. Sosa-Sánchez, Ramsés E. Bautista González, Maximino L. Arroyo Carrasco, Marycarmen Peña-Gomar, Victor Coello, and Ricardo Téllez-Limón. 2023. "Integrated Optical Filters with Hyperbolic Metamaterials" Nanomaterials 13, no. 4: 759. https://doi.org/10.3390/nano13040759

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop