Citation:

Hybrid-integrated chalcogenide photonics


  • Light: Advanced Manufacturing  4, Article number: (2023)
More Information
  • Corresponding author:
    Bin Zhang (zhangbin5@mail.sysu.edu.cn)Zhaohui Li (lzhh88@mail.sysu.edu.cn)
  • Received: 27 March 2023
    Revised: 26 July 2023
    Accepted: 27 July 2023
    Accepted article preview online: 28 July 2023
    Published online: 22 September 2023

doi: https://doi.org/10.37188/lam.2023.024

  • High-quality photonic materials are critical for promoting integrated photonic devices with broad bandwidths, high efficiencies, and flexibilities for high-volume chip-scale fabrication. Recently, we designed a home-developed chalcogenide glass (ChG)-Ge25Sb10S65 (GeSbS) for optical information processing chips and systems, which featured an ultrabroad transmission window, a high Kerr nonlinearity and photoelastic coefficient, and compatibility with the photonic hybrid integration technology of silicon photonics. Chip-integrated GeSbS microresonators and microresonator arrays with high quality factors and lithographically controlled fine structures were fabricated using a modified nanofabrication process. Moreover, considering the high Kerr nonlinearity and photoelastic effect of ChGs, we realised a novel ChG hybrid integrated chip, inspired by recent advances in integrated soliton microcombs and acousto-optic (AO) modulators.

  • 加载中
  • [1] Spencer, D. T. et al. An optical-frequency synthesizer using integrated photonics. Nature 557, 81-85 (2018). doi: 10.1038/s41586-018-0065-7
    [2] Feldmann, J. et al. Parallel convolutional processing using an integrated photonic tensor core. Nature 589, 52-58 (2021). doi: 10.1038/s41586-020-03070-1
    [3] Atabaki, A. H. et al. Integrating photonics with silicon nanoelectronics for the next generation of systems on a chip. Nature 556, 349-354 (2018). doi: 10.1038/s41586-018-0028-z
    [4] Sun, C. et al. Single-chip microprocessor that communicates directly using light. Nature 528, 534-538 (2015). doi: 10.1038/nature16454
    [5] Yang, Z. Y. et al. Single-nanowire spectrometers. Science 365, 1017-1020 (2019). doi: 10.1126/science.aax8814
    [6] Liu, J. Q. et al. High-yield, wafer-scale fabrication of ultralow-loss, dispersion-engineered silicon nitride photonic circuits. Nature Communications 12, 2236 (2021). doi: 10.1038/s41467-021-21973-z
    [7] Yang, Z. Y. et al. Miniaturization of optical spectrometers. Science 371, eabe0722 (2021). doi: 10.1126/science.abe0722
    [8] Gaeta, A. L., Lipson, M. & Kippenberg, T. J. Photonic-chip-based frequency combs. Nature Photonics 13, 158-169 (2019). doi: 10.1038/s41566-019-0358-x
    [9] Xiang, C. et al. Laser soliton microcombs heterogeneously integrated on silicon. Science 373, 99-103 (2021). doi: 10.1126/science.abh2076
    [10] Guidry, M. A. et al. Quantum optics of soliton microcombs. Nature Photonics 16, 52-58 (2022). doi: 10.1038/s41566-021-00901-z
    [11] Liu, X. W. et al. Aluminum nitride nanophotonics for beyond-octave soliton microcomb generation and self-referencing. Nature Communications 12, 5428 (2021). doi: 10.1038/s41467-021-25751-9
    [12] Chang, L. et al. Ultra-efficient frequency comb generation in AlGaAs-on-insulator microresonators. Nature Communications 11, 1331 (2020). doi: 10.1038/s41467-020-15005-5
    [13] Zhang, M. et al. Broadband electro-optic frequency comb generation in a lithium niobate microring resonator. Nature 568, 373-377 (2019). doi: 10.1038/s41586-019-1008-7
    [14] Chen, G. Y. et al. Advances in lithium niobate photonics: development status and perspectives. Advanced Photonics 4, 034003 (2022).
    [15] Eggleton, B. J. et al. Brillouin integrated photonics. Nature Photonics 13, 664-677 (2019). doi: 10.1038/s41566-019-0498-z
    [16] Eggleton, B. J., Luther-Davies, B. & Richardson, K. Chalcogenide photonics. Nature Photonics 5, 141-148 (2011). doi: 10.1038/nphoton.2011.309
    [17] Büttner, T. F. S. et al. Phase-locked, chip-based, cascaded stimulated Brillouin scattering. Optica 1, 311-314 (2014). doi: 10.1364/OPTICA.1.000311
    [18] Giacoumidis, E. et al. Chip-based Brillouin processing for carrier recovery in self-coherent optical communications. Optica 5, 1191-1199 (2018). doi: 10.1364/OPTICA.5.001191
    [19] Gao, J. N. et al. Near-infrared to ultra-violet frequency conversion in chalcogenide metasurfaces. Nature Communications 12, 5833 (2021). doi: 10.1038/s41467-021-26094-1
    [20] Petersen, C. R. et al. Mid-infrared supercontinuum covering the 1.4–13.3 μm molecular fingerprint region using ultra-high NA chalcogenide step-index fibre. Nature Photonics 8, 830-834 (2014).
    [21] Zhang, W. et al. Designing crystallization in phase-change materials for universal memory and neuro-inspired computing. Nature Reviews Materials 4, 150-168 (2019). doi: 10.1038/s41578-018-0076-x
    [22] Wang, S. K. et al. Chalcogenide glass IR artificial compound eyes based on femtosecond laser microfabrication. Advanced Materials Technologies 8, 2200741 (2023). doi: 10.1002/admt.202200741
    [23] Asobe, M. et al. Third‐order nonlinear spectroscopy in As2S3 chalcogenide glass fibers. Journal of Applied Physics 77, 5518-5523 (1995). doi: 10.1063/1.359256
    [24] Wan, L. et al. Highly efficient acousto-optic modulation using nonsuspended thin-film lithium niobate-chalcogenide hybrid waveguides. Light:Science & Applications 11, 145 (2022).
    [25] Zhang, Y. F. et al. Broadband transparent optical phase change materials for high-performance nonvolatile photonics. Nature Communications 10, 4279 (2019). doi: 10.1038/s41467-019-12196-4
    [26] Zhang, Y. F. et al. Electrically reconfigurable non-volatile metasurface using low-loss optical phase-change material. Nature Nanotechnology 16, 661-666 (2021). doi: 10.1038/s41565-021-00881-9
    [27] Yu, Y. et al. Experimental demonstration of linearly polarized 2–10 μm supercontinuum generation in a chalcogenide rib waveguide. Optics Letters 41, 958-961 (2016). doi: 10.1364/OL.41.000958
    [28] Xia, D. et al. On-chip broadband mid-infrared supercontinuum generation based on highly nonlinear chalcogenide glass waveguides. Frontiers in Physics 9, 598091 (2021). doi: 10.3389/fphy.2021.598091
    [29] Broaddus, D. H. et al. Silicon-waveguide-coupled high-Q chalcogenide microspheres. Optics Express 17, 5998-6003 (2009). doi: 10.1364/OE.17.005998
    [30] Zhang, B. et al. On-chip chalcogenide microresonators with low-threshold parametric oscillation. Photonics Research 9, 1272-1279 (2021). doi: 10.1364/PRJ.422435
    [31] Morrison, B. et al. Compact Brillouin devices through hybrid integration on silicon. Optica 4, 847-854 (2017). doi: 10.1364/OPTICA.4.000847
    [32] Hô, N. et al. Photosensitivity of As2S3 chalcogenide thin films at 1.5 µm. Optics Letters 28, 965-967 (2003).
    [33] Knotek, P. et al. On the ultraviolet light induced oxidation of amorphous As2S3 film. Thin Solid Films 520, 5472-5478 (2012). doi: 10.1016/j.tsf.2012.03.116
    [34] Xia, D. et al. Integrated chalcogenide photonics for microresonator soliton combs. Laser & Photonics Reviews 17, 2200219 (2023).
    [35] Wang, T. et al. Systematic z-scan measurements of the third order nonlinearity of chalcogenide glasses. Optical Materials Express 4, 1011-1022 (2014). doi: 10.1364/OME.4.001011
    [36] Lihachev, G. et al. Platicon microcomb generation using laser self-injection locking. Nature Communications 13, 1771 (2022). doi: 10.1038/s41467-022-29431-0
    [37] Pu, M. H. et al. Ultra‐Efficient and broadband nonlinear AlGaAs‐on‐insulator chip for low‐power optical signal processing. Laser & Photonics Reviews 12, 1800111 (2018).
    [38] Xia, D. et al. Engineered Raman lasing in photonic integrated chalcogenide microresonators. Laser & Photonics Reviews 16, 2100443 (2022).
    [39] Gundavarapu, S. et al. Sub-hertz fundamental linewidth photonic integrated Brillouin laser. Nature Photonics 13, 60-67 (2019). doi: 10.1038/s41566-018-0313-2
    [40] Ji, X. C. et al. Ultra-low-loss on-chip resonators with sub-milliwatt parametric oscillation threshold. Optica 4, 619-624 (2017). doi: 10.1364/OPTICA.4.000619
    [41] Kim, D. G. et al. Universal light-guiding geometry for on-chip resonators having extremely high Q-factor. Nature Communications 11, 5933 (2020). doi: 10.1038/s41467-020-19799-2
    [42] Yu, M. J. et al. Raman lasing and soliton mode-locking in lithium niobate microresonators. Light:Science & Applications 9, 9 (2020).
    [43] Madden, S. J. et al. Long, low loss etched As2S3 chalcogenide waveguides for all-optical signal regeneration. Optics Express 15, 14414-14421 (2007). doi: 10.1364/OE.15.014414
    [44] Zhang, B. et al. High brightness 2.2–12 µm mid-infrared supercontinuum generation in a nontoxic chalcogenide step‐index fiber. Journal of the American Ceramic Society 99, 2565-2568 (2016).
    [45] Riemensberger, J. et al. A photonic integrated continuous-travelling-wave parametric amplifier. Nature 612, 56-61 (2022). doi: 10.1038/s41586-022-05329-1
    [46] Ye, Z. C. et al. Overcoming the quantum limit of optical amplification in monolithic waveguides. Science Advances 7, eabi8150 (2021). doi: 10.1126/sciadv.abi8150
    [47] Morrison, B. et al. Tunable microwave photonic notch filter using on-chip stimulated Brillouin scattering. Optics Communications 313, 85-89 (2014). doi: 10.1016/j.optcom.2013.09.065
    [48] Marpaung, D. et al. Low-power, chip-based stimulated Brillouin scattering microwave photonic filter with ultrahigh selectivity. Optica 2, 76-83 (2015). doi: 10.1364/OPTICA.2.000076
    [49] Lin, H. T. et al. Demonstration of high-Q mid-infrared chalcogenide glass-on-silicon resonators. Optics Letters 38, 1470-1472 (2013). doi: 10.1364/OL.38.001470
    [50] Zhang, R. Z. et al. High quality, high index-contrast chalcogenide microdisk resonators. Optics Express 29, 17775-17783 (2021). doi: 10.1364/OE.427054
    [51] Zhu, Y. et al. Effects of shallow suspension in low-loss waveguide-integrated chalcogenide microdisk resonators. Journal of Lightwave Technology 38, 4817-4823 (2020).
    [52] Du, Q. Y. et al. Low-loss photonic device in Ge–Sb–S chalcogenide glass. Optics Letters 41, 3090-3093 (2016). doi: 10.1364/OL.41.003090
    [53] Yang, Z. et al. High-Q, submicron-confined chalcogenide microring resonators. Optics Express 29, 33225-33233 (2021). doi: 10.1364/OE.434808
    [54] Jean, P. et al. Etchless chalcogenide microresonators monolithically coupled to silicon photonic waveguides. Optics Letters 45, 2830-2833 (2020). doi: 10.1364/OL.392879
    [55] Gai, X. et al. Dispersion engineered Ge115As24Se64. 5 nanowires with a nonlinear parameter of 136W−1m−1 at 1550 nm. Optics Express 18, 18866-18874 (2010). doi: 10.1364/OE.18.018866
    [56] Serna, S. et al. Nonlinear optical properties of integrated GeSbS chalcogenide waveguides. Photonics Research 6, B37-B42 (2018). doi: 10.1364/PRJ.6.000B37
    [57] Ma, P. et al. Low-loss chalcogenide waveguides for chemical sensing in the mid-infrared. Optics Express 21, 29927-29937 (2013). doi: 10.1364/OE.21.029927
    [58] Shen, W. H. et al. Chalcogenide glass photonic integration for improved 2 µm optical interconnection. Photonics Research 8, 1484-1490 (2020). doi: 10.1364/PRJ.398957
    [59] Hô, N. et al. Single-mode low-loss chalcogenide glass waveguides for the mid-infrared. Optics Letters 31, 1860-1862 (2006). doi: 10.1364/OL.31.001860
    [60] Jin, T. N. et al. Mid-infrared chalcogenide waveguides for real-time and nondestructive volatile organic compound detection. Analytical Chemistry 91, 817-822 (2019). doi: 10.1021/acs.analchem.8b03004
    [61] Newman, Z. L. et al. Architecture for the photonic integration of an optical atomic clock. Optica 6, 680-685 (2019). doi: 10.1364/OPTICA.6.000680
    [62] Del'Haye, P. et al. Phase-coherent microwave-to-optical link with a self-referenced microcomb. Nature Photonics 10, 516-520 (2016). doi: 10.1038/nphoton.2016.105
    [63] Papp, S. B. et al. Microresonator frequency comb optical clock. Optica 1, 10-14 (2014). doi: 10.1364/OPTICA.1.000010
    [64] Drake, T. E. et al. Terahertz-rate Kerr-microresonator optical clockwork. Physical Review X 9, 031023 (2019).
    [65] Geng, Y. et al. Coherent optical communications using coherence-cloned Kerr soliton microcombs. Nature Communications 13, 1070 (2022). doi: 10.1038/s41467-022-28712-y
    [66] Fülöp, A. et al. High-order coherent communications using mode-locked dark-pulse Kerr combs from microresonators. Nature Communications 9, 1598 (2018). doi: 10.1038/s41467-018-04046-6
    [67] Hu, H. et al. Single-source chip-based frequency comb enabling extreme parallel data transmission. Nature Photonics 12, 469-473 (2018). doi: 10.1038/s41566-018-0205-5
    [68] Marin-Palomo, P. et al. Microresonator-based solitons for massively parallel coherent optical communications. Nature 546, 274-279 (2017). doi: 10.1038/nature22387
    [69] Lundberg, L. et al. Phase-coherent lightwave communications with frequency combs. Nature Communications 11, 201 (2020). doi: 10.1038/s41467-019-14010-7
    [70] Liu, J. Q. et al. Photonic microwave generation in the X-and K-band using integrated soliton microcombs. Nature Photonics 14, 486-491 (2020). doi: 10.1038/s41566-020-0617-x
    [71] Bao, C. Y. et al. Architecture for microcomb-based GHz-mid-infrared dual-comb spectroscopy. Nature Communications 12, 6573 (2021). doi: 10.1038/s41467-021-26958-6
    [72] Guo, H. R. et al. Nanophotonic supercontinuum-based mid-infrared dual-comb spectroscopy. Optica 7, 1181-1188 (2020). doi: 10.1364/OPTICA.396542
    [73] Dutt, A. et al. On-chip dual-comb source for spectroscopy. Science Advances 4, e1701858 (2018). doi: 10.1126/sciadv.1701858
    [74] Suh, M. G. et al. Microresonator soliton dual-comb spectroscopy. Science 354, 600-603 (2016). doi: 10.1126/science.aah6516
    [75] Liang, Q. Z. et al. Ultrasensitive multispecies spectroscopic breath analysis for real-time health monitoring and diagnostics. Proceedings of the National Academy of Sciences of the United States of America 118, e2105063118 (2021).
    [76] Riemensberger, J. et al. Massively parallel coherent laser ranging using a soliton microcomb. Nature 581, 164-170 (2020). doi: 10.1038/s41586-020-2239-3
    [77] Trocha, P. et al. Ultrafast optical ranging using microresonator soliton frequency combs. Science 359, 887-891 (2018). doi: 10.1126/science.aao3924
    [78] Jin, W. et al. Hertz-linewidth semiconductor lasers using CMOS-ready ultra-high-Q microresonators. Nature Photonics 15, 346-353 (2021). doi: 10.1038/s41566-021-00761-7
    [79] Zhang, S. Y. et al. Sub-milliwatt-level microresonator solitons with extended access range using an auxiliary laser. Optica 6, 206-212 (2019). doi: 10.1364/OPTICA.6.000206
    [80] Luke, K. et al. Broadband mid-infrared frequency comb generation in a Si3N4 microresonator. Optics Letters 40, 4823-4826 (2015). doi: 10.1364/OL.40.004823
    [81] Griffith, A. G. et al. Silicon-chip mid-infrared frequency comb generation. Nature Communications 6, 6299 (2015). doi: 10.1038/ncomms7299
    [82] Brasch, V. et al. Photonic chip–based optical frequency comb using soliton Cherenkov radiation. Science 351, 357-360 (2016). doi: 10.1126/science.aad4811
    [83] Brasch, V. et al. Bringing short-lived dissipative Kerr soliton states in microresonators into a steady state. Optics Express 24, 29312-29320 (2016). doi: 10.1364/OE.24.029312
    [84] Guo, H. et al. Universal dynamics and deterministic switching of dissipative Kerr solitons in optical microresonators. Nature Physics 13, 94-102 (2017). doi: 10.1038/nphys3893
    [85] Joshi, C. et al. Thermally controlled comb generation and soliton modelocking in microresonators. Optics Letters 41, 2565-2568 (2016). doi: 10.1364/OL.41.002565
    [86] Yi, X. et al. Active capture and stabilization of temporal solitons in microresonators. Optics Letters 41, 2037-2040 (2016). doi: 10.1364/OL.41.002037
    [87] Zhou, H. et al. Soliton bursts and deterministic dissipative Kerr soliton generation in auxiliary-assisted microcavities. Light:Science & Applications 8, 50 (2019).
    [88] Miao, R. L. et al. Repetition rate locked single-soliton microcomb generation via rapid frequency sweep and sideband thermal compensation. Photonics Research 10, 1859-1867 (2022). doi: 10.1364/PRJ.458472
    [89] Shen, B. Q. et al. Integrated turnkey soliton microcombs. Nature 582, 365-369 (2020). doi: 10.1038/s41586-020-2358-x
    [90] Lihachev, G. et al. Low-noise frequency-agile photonic integrated lasers for coherent ranging. Nature Communications 13, 3522 (2022). doi: 10.1038/s41467-022-30911-6
    [91] Gong, Z. et al. Photonic dissipation control for Kerr soliton generation in strongly Raman-active media. Physical Review Letters 125, 183901 (2020). doi: 10.1103/PhysRevLett.125.183901
    [92] Huang, S. W. et al. Globally stable microresonator turing pattern formation for coherent high-power THz radiation on-chip. Physical Review X 7, 041002 (2017).
    [93] Vanholsbeeck, F., Emplit, P. & Coen, S. Complete experimental characterization of the influence of parametric four-wave mixing on stimulated Raman gain. Optics Letters 28, 1960-1962 (2003). doi: 10.1364/OL.28.001960
    [94] Vanholsbeeck, F. et al. Numerical modeling of a four-wave-mixing-assisted Raman fiber laser. Optics Letters 29, 2719-2721 (2004). doi: 10.1364/OL.29.002719
    [95] Vanholsbeeck, F. et al. Cascaded Raman generation in optical fibers: influence of chromatic dispersion and Rayleigh backscattering. Optics Letters 29, 998-1000 (2004). doi: 10.1364/OL.29.000998
    [96] Sylvestre, T. et al. Raman-assisted parametric frequency conversion in a normally dispersive single-mode fiber. Optics Letters 24, 1561-1563 (1999). doi: 10.1364/OL.24.001561
    [97] Okubo, S. et al. Offset-free optical frequency comb self-referencing with an f-2f interferometer. Optica 5, 188-192 (2018). doi: 10.1364/OPTICA.5.000188
    [98] Brasch, V. et al. Self-referenced photonic chip soliton Kerr frequency comb. Light:Science & Applications 6, e16202 (2017).
    [99] Wang, Z. F. et al. Engineered octave frequency comb in integrated chalcogenide dual-ring microresonators. Frontiers in Photonics 4, 1066993 (2023). doi: 10.3389/fphot.2023.1066993
    [100] Kim, S. et al. Dispersion engineering and frequency comb generation in thin silicon nitride concentric microresonators. Nature Communications 8, 372 (2017). doi: 10.1038/s41467-017-00491-x
    [101] Moille, G. et al. Phased-locked two-color single soliton microcombs in dispersion-engineered Si3N4 resonators. Optics Letters 43, 2772-2775 (2018). doi: 10.1364/OL.43.002772
    [102] Ji, H. L. et al. Design of Partially Etched GaP-OI Microresonators for Two-Color Kerr Soliton Generation at NIR and MIR. Proceedings of 2022 Asia Communications and Photonics Conference. Shenzhen, China: IEEE, 2022, 1622-1625.
    [103] Parra-Rivas, P. et al. Dynamics of localized and patterned structures in the Lugiato-Lefever equation determine the stability and shape of optical frequency combs. Physical Review A 89, 043813 (2014). doi: 10.1103/PhysRevA.89.043813
    [104] Lugiato, L. A. et al. From the Lugiato–Lefever equation to microresonator-based soliton Kerr frequency combs. Philosophical Transactions of the Royal Society A:Mathematical,Physical and Engineering Sciences 376, 20180113 (2018). doi: 10.1098/rsta.2018.0113
    [105] Chembo, Y. K. & Menyuk, C. R. Spatiotemporal Lugiato-Lefever formalism for Kerr-comb generation in whispering-gallery-mode resonators. Physical Review A 87, 053852 (2013). doi: 10.1103/PhysRevA.87.053852
    [106] Balram, K. C. et al. Coherent coupling between radiofrequency, optical and acoustic waves in piezo-optomechanical circuits. Nature Photonics 10, 346-352 (2016). doi: 10.1038/nphoton.2016.46
    [107] Sarabalis, C. J. et al. Acousto-optic modulation of a wavelength-scale waveguide. Optica 8, 477-483 (2021). doi: 10.1364/OPTICA.413401
    [108] Savage, N. Acousto-optic devices. Nature Photonics 4, 728-729 (2010). doi: 10.1038/nphoton.2010.229
    [109] Munk, D. et al. Surface acoustic wave photonic devices in silicon on insulator. Nature Communications 10, 4214 (2019). doi: 10.1038/s41467-019-12157-x
    [110] Balram, K. C. et al. Acousto-optic modulation and optoacoustic gating in piezo-optomechanical circuits. Physical Review Applied 7, 024008 (2017). doi: 10.1103/PhysRevApplied.7.024008
    [111] Qi, Y. F. & Li, Y. Integrated lithium niobate photonics. Nanophotonics 9, 1287-1320 (2020). doi: 10.1515/nanoph-2020-0013
    [112] Boes, A. et al. Status and Potential of Lithium Niobate on Insulator (LNOI) for Photonic Integrated Circuits. Laser & Photonics Reviews 12, 1700256 (2018).
    [113] Shin, H. et al. Tailorable stimulated Brillouin scattering in nanoscale silicon waveguides. Nature communications 4, 1944 (2013). doi: 10.1038/ncomms2943
    [114] Safavi-Naeini, A. H. et al. Controlling phonons and photons at the wavelength scale: integrated photonics meets integrated phononics. Optica 6, 213-232 (2019). doi: 10.1364/OPTICA.6.000213
    [115] Song, J. C. et al. Stimulated Brillouin scattering in low-loss Ge25Sb10S65 chalcogenide waveguides. Journal of Lightwave Technology 39, 5048-5053 (2021). doi: 10.1109/JLT.2021.3078722
    [116] Yang, Z. Q. et al. Efficient acousto-optic modulation using a microring resonator on a thin-film lithium niobate–chalcogenide hybrid platform. Optics Letters 47, 3808-3811 (2022). doi: 10.1364/OL.464482
    [117] Hassanien, A. E. et al. Efficient and wideband acousto-optic modulation on thin-film lithium niobate for microwave-to-photonic conversion. Photonics Research 9, 1182-1190 (2021). doi: 10.1364/PRJ.421612
    [118] Lin, H. T. et al. Chalcogenide glass-on-graphene photonics. Nature Photonics 11, 798-805 (2017). doi: 10.1038/s41566-017-0033-z
    [119] Song, J. C. et al. Ultrasound Measurement Using On-Chip Optical Micro-Resonators and Digital Optical Frequency Comb. Journal of Lightwave Technology 38, 5293-5301 (2020). doi: 10.1109/JLT.2020.2982211
    [120] Cao, Z. F. et al. Investigation of the acousto‐optical properties of Ge–As–Te–(Se) chalcogenide glasses at 10.6 μm wavelength. Journal of the American Ceramic Society 104, 3224-3234 (2021). doi: 10.1111/jace.17767
    [121] Cai, L. T. et al. Acousto-optical modulation of thin film lithium niobate waveguide devices. Photonics Research 7, 1003-1013 (2019). doi: 10.1364/PRJ.7.001003
    [122] Khan, S. I. et al. Extraction of elastooptic coefficient of thin-film arsenic trisulfide using a Mach–Zehnder acoustooptic modulator on lithium niobate. Journal of Lightwave Technology 38, 2053-2059 (2020). doi: 10.1109/JLT.2019.2960396
    [123] Shao, L. B. et al. Microwave-to-optical conversion using lithium niobate thin-film acoustic resonators. Optica 6, 1498-1505 (2019). doi: 10.1364/OPTICA.6.001498
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Figures(10) / Tables(3)

Research Summary

Hybrid-integrated chalcogenide photonics: recent breakthroughs and opportunities

Multifunctional photonic integrated chips, inspired by the hybrid and heterogeneous integration of various materials, have achieved unique optoelectronic properties by precisely controlling the structure of materials and their interaction with the optical field at the chipscale. Chalcogenide glasses (ChGs) are promising candidates for Highly efficient nonlinear photonics and acousto-optic modulator devices due to their broadband transmittance window, high nonlinearity and photoelastic coefficient. Bin Zhang from Sun-Yat Sen University and colleagues, carefully reviewed their recent advancements in the development of integrated ChG photonic chips with ultralow optical losses and the applications of these chips in nonlinear photodevices and acousto-optic modulators. These advancements have effectively addressed key issues such as improving the nonlinear interaction within highly confined waveguides and efficiently modulating heterogeneously integrated optical waveguides using surface acoustic waves. These results demonstrate that ChG-based photonics devices offer new opportunities for achieving higher levels of integration, larger bandwidths, and lower power consumption in photonic chips.


show all

Article Metrics

Article views(1736) PDF downloads(863) Citation(0) Citation counts are provided from Web of Science. The counts may vary by service, and are reliant on the availability of their data.

Hybrid-integrated chalcogenide photonics

  • 1. Guangdong Provincial Key Laboratory of Optoelectronic Information Processing Chips and Systems, School of Electrical and Information Technology, Sun Yat-sen University, Guangzhou 510275, China
  • 2. Key Laboratory of Optoelectronic Materials and Technologies, Sun Yat-sen University, Guangzhou 510275, China
  • 3. Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), Zhuhai, 519000, China
  • 4. Department of Electronic Engineering, College of Information Science and
  • Corresponding author:

    Bin Zhang, zhangbin5@mail.sysu.edu.cn

    Zhaohui Li, lzhh88@mail.sysu.edu.cn

doi: https://doi.org/10.37188/lam.2023.024

Abstract: 

High-quality photonic materials are critical for promoting integrated photonic devices with broad bandwidths, high efficiencies, and flexibilities for high-volume chip-scale fabrication. Recently, we designed a home-developed chalcogenide glass (ChG)-Ge25Sb10S65 (GeSbS) for optical information processing chips and systems, which featured an ultrabroad transmission window, a high Kerr nonlinearity and photoelastic coefficient, and compatibility with the photonic hybrid integration technology of silicon photonics. Chip-integrated GeSbS microresonators and microresonator arrays with high quality factors and lithographically controlled fine structures were fabricated using a modified nanofabrication process. Moreover, considering the high Kerr nonlinearity and photoelastic effect of ChGs, we realised a novel ChG hybrid integrated chip, inspired by recent advances in integrated soliton microcombs and acousto-optic (AO) modulators.

Research Summary

Hybrid-integrated chalcogenide photonics: recent breakthroughs and opportunities

Multifunctional photonic integrated chips, inspired by the hybrid and heterogeneous integration of various materials, have achieved unique optoelectronic properties by precisely controlling the structure of materials and their interaction with the optical field at the chipscale. Chalcogenide glasses (ChGs) are promising candidates for Highly efficient nonlinear photonics and acousto-optic modulator devices due to their broadband transmittance window, high nonlinearity and photoelastic coefficient. Bin Zhang from Sun-Yat Sen University and colleagues, carefully reviewed their recent advancements in the development of integrated ChG photonic chips with ultralow optical losses and the applications of these chips in nonlinear photodevices and acousto-optic modulators. These advancements have effectively addressed key issues such as improving the nonlinear interaction within highly confined waveguides and efficiently modulating heterogeneously integrated optical waveguides using surface acoustic waves. These results demonstrate that ChG-based photonics devices offer new opportunities for achieving higher levels of integration, larger bandwidths, and lower power consumption in photonic chips.


show all
    • Silicon photonics is a rapidly growing field that combines the optical and electrical properties of silicon to create various systems and devices that are fabricated using the mature complementary metal-oxide semiconductor (CMOS) technique, which plays a critical role in next-generation communication systems, data centres, optical computing, and biosensing1-5. Recently, multifunctional photonic integrated chips created via the hybrid and heterogeneous photonic integration of various materials have promoted broadband laser source generation, high-capacity information processing and storage, and chip-scale quantum optics with miniaturised footprints and low power consumption6,7. A series of integrated photonic materials, including III-V materials such as silicon nitride (Si3N4), lithium niobate (LiNbO3), silicon carbide (SiC), aluminum nitride (AlN), and chalcogenide glass (ChGs), have been explored for promoting the high-volume fabrication of integrated chips with a broad bandwidth, high efficiency, and flexibility8-14.

      Among these, ChGs are promising candidates for broadband laser sources and exhibit highly effective photoelastic effects. They are amorphous compounds composed of one or more chalcogenide elements, including sulfur (S), selenium (Se), selenium (Te), and other metal or non-oxide elements15-19. Currently, commercial ChG bulk materials, thin films, and optical fibres are widely used in infrared imaging lenses, phase-change memory chips, and in situ biomolecular detection16,20-22. Demonstrations involving ChGs have revealed the following characteristics: an absence of two-photon absorption (TPA) and free-carrier effects in a wide transmission window (0.5–25 μm)16; a high linear refractive index ($ {n}_{0}\approx 2.2-3.5 $)16; a nonlinear refractive index23 ($ {n}_{2} > {10}^{-18}{m}^{2}/W $, 2–3 orders of magnitude higher than that of silica); and high photoelastic coefficients ($ {{p}_{11}\approx p}_{12}\approx 0.238 $, approximately two times that of LiNbO324). Additionally, some ChGs called ChG-based phase-change materials (PCMs) exhibit rapid and reversible phase transitions between the amorphous and crystalline phases under external thermal stimulation. Examples include Ge2Sb2Te5 (GST), Ge2Sb2Se4Te1 (GSST), and Sb2S3. Owing to the significant differences in the optical properties (i.e. refractive index and extinction coefficient) between their different phases, photonic devices based on PCMs can quickly change the amplitude and phase of light, supporting their widespread use in applications such as optical switches, optical storage, and optical computing2,25,26. Moreover, the optical and material properties of ChGs can be tailored using different glass components to meet the requirements of various photonic applications. Wafer-scale ChG thin films have been prepared directly on crystal or amorphous platforms by low-temperature deposition technology (< 350℃). Owing to their amorphous nature, wafer-scale ChG films can be prepared without additional processes such as wafer bonding and crack mitigation. Moreover, devices integrated with ChGs as both the core and cladding are promising for eliminating SiO2-induced optical absorption in the infrared wavelength range and enabling broadband (> 4 μm) MIR-integrated photonics27,28.

      However, in the past decade, developing integrated ChG photonic chips with ultralow optical losses and high performances has been exceptionally challenging8, 16, 29-31. Typical arsenic (As)-based components (As2S3) suffer from photooxidation and photoinduced changes in refractive indices32,33. In our previous study30, we presented a light-annealing method and developed an on-chip optical parametric oscillator (OPO). The low laser-induced damage (LDT) threshold has hindered the development of advanced nonlinear applications such as soliton microcombs. Moreover, the low glass transition temperature of As2S3 (~200 °C) leads to a low fabrication tolerance, similar to that under SiO2 cladding deposition based on the ICP-CVD method at a higher temperature of 300 oC34. Recently, we have developed a new ChG photonic material-Ge25Sb10S65 (GeSbS), which exhibits a ultra-wide transparency (from 0.5 to 10 μm without TPA), a large linear index (n0 $ \approx $ 2.2), a strong Kerr nonlinearity (1.4 × 10−18 m2W−1 at 1550 nm)35, a relatively low thermo-optic coefficient ($ \approx $ 3.1 ×10−5 K−1), and a large bandgap (2.64 eV), as shown in Fig. 1. Additionally, it displays a high laser damage threshold ($ \approx $ 820.7 GW cm−2), and a high glass transition temperature (>350 °C), compared with the properties of typical As2S3. Nonetheless, as a new photonically integrated material, GeSbS faces several challenges in realising high-performance hybrid photonic integrated devices with other materials. First, the relationship between tailored ChG components and the properties of ChG-integrated devices remains unclear. The adjustable optical and material properties of ChGs, which are obtained through tailored glass compositions, are one of their most attractive properties. These properties enable the adjustment of the target performance of integrated chips. Second, hybrid integrated devices based on ChGs and different materials can be geometrically engineered to combine their material and optical properties, resulting in a design framework with complementary advantages. Most importantly, it is imperative to establish a fabrication route, which encompasses optical materials and devices in emerging systems, for hybrid ChG photonic-integrated devices that are compatible with silicon photonics. A stable manufacturing process that fits into a well-established semiconductor device fabrication infrastructure is clearly advantageous over novel manufacturing processes for new nanomaterial-based devices7. By further reducing the optical transmission loss, the pumping power of various nonlinear processes can be reduced to the submilliwatt level, and fully integrated photonic devices free of external optical amplifiers can be realised. Solving these critical issues and realising these technologies will determine whether ChG devices can be integrated into hybrid silicon chips.

      Fig. 1  Optical properties of GeSbS materials. Adapted with permission from Ref. 34, copyright 2022, WILEY. a Measured transmission window and refractive index of a GeSbS bulk material. Inset: Measured LDT at five different positions of the film. b Z-scan (close aperture) trace for determining the nonlinear refractive index of a GeSbS film. c Measured resonance frequency shift versus temperature for determining the thermo-optic coefficient (TOC) of a GeSbS resonator. d Measured Tauc’s plot for determining the bandgap of the GeSbS film.

      In this article, we review recent advances in the optimised fabrication of new GeSbS-integrated devices and their applications based on their prominent Kerr nonlinearity and the photoelastic effect. We developed a modified ChG waveguide fabrication process compatible with silicon photonics integration that leverages higher quality (Q) factors (more than 5×106) with respect to state-of-the-art ChG-based microring resonators34. Furthermore, based on the high nonlinearity and photoelastic coefficient of GeSbS, we experimentally demonstrated the integrated soliton microcombs and acousto-optic (AO) modulators in improved ChG-integrated devices.

    • Low-propagation-loss photonic integrated devices are crucial in various photonic applications such as low-threshold frequency combs34,36, optical signal processing18,37, and on-chip lasers38,39. Since they increase the scattering of light confined in devices, surface and sidewall roughness are crucial for realising low-loss photonic integrated devices40. By optimising thermal annealing and dry etching processes, we developed an improved fabrication method for ChG photonic integrated devices to minimise the surface roughnesses of both top and sidewalls34.

      ChG thin films are typically prepared using thermal evaporation at a low temperature (Tg), through which amorphous ChGs can directly adhere to silicon substrates41, free of the wafer-bonding process12,42. This process is simple and low-cost and allows the film to be easily controlled43. A GeSbS film was deposited on a 4-inch silicon wafer using an optimised thermal evaporation technique at a low temperature (350 °C) (Fig. 2a). The uniformity of the film thickness and refractive index was characterised on a 4-inch GeSbS wafer (Fig. 2b, c). There were fluctuations in the thickness (≈ ±6.5 nm) and refractive index (≈ ±0.0075) of the film, and it had a thickness of 850 nm. However, since these fluctuations were small, they did not affect the application of this wafer in on-chip photonic integrated circuits. Then, a thermal annealing process was directly performed to reduce the surface roughness and improve the optical quality of the GeSbS film30,34,38. The effect of different annealing temperatures on the surfaces of the films were explored, including the as-deposited and thermal annealed states at 300, 350, and 400 °C. After thermal annealing at 350 °C, the RMS roughness of the GeSbS film significantly decreased from 0.623 nm for the as-deposited film to 0.241 nm34 (Fig. 2d). Annealing can cause changes in the density and molecular structure of ChGs, leading to an increase in the film thickness (less than 10 nm) and a decrease in the refractive index (less than 0.08) for films with a thickness of 850 nm. Considering the minimal changes in surface roughness at 400 °C, 350 °C was selected as the optimal thermal annealing temperature.

      Fig. 2  Characterisation of the GeSbS film. Adapted with permission from Ref. 34, copyright 2023, WILEY. a The 4-inch GeSbS film wafer. b Thickness mapping and c refractive index mapping of a 4-inch GeSbS wafer. d Atomic force microscopy (AFM) images of GeSbS thin films in a 5 × 5 μm2 area: i As-deposited and after thermal-annealing at ii 300, iii 350, and iv 400 °C.

      Subsequently, the dry-etching process was optimised to reduce the roughness of the sidewalls and improve the verticality of the waveguides. Here, the CF4/CHF3/Ar gas-based ICP-RIE etching process was modified by adjusting the flow rates of O2 and CF4 gases to reduce the scattering losses caused by in situ polymer deposition on the waveguide sidewalls (Fig. 3a, b)34. A vertical sidewall can also be achieved through an improved etching process, which is critical for the precise dispersion engineering of waveguides (Fig. 3c). Additionally, the coupling conditions of the microresonator could be finely controlled by carefully designing the width of the bus waveguide, coupling gap, and coupling length. For instance, a GeSbS microresonator with an integrated pulley bus waveguide was finely fabricated with a 500-nm target coupling gap and good verticality (Fig. 3df).

      Fig. 3  Scanning electron microscopy (SEM) images of the fabricated device. Adapted with permission from Ref. 34, copyright 2023, WILEY. a, b The sidewall of the waveguides before and after the improved dry etching process. Inset: The schematic of the waveguide structure; c the cross-section with a silica cladding; d a GeSbS microresonator with a radius of 100 μm with an integrated pulley bus waveguide; e the pulley coupling region and f the cross-section for the etching width of 500 nm. The cross-section of the microresonator is 2.4 × 0.8 μm.

      We have proposed a complete fabrication process for low-loss ChG-integrated devices, through which a 1× 2 cm GeSbS photonic chip can be obtained (Fig. 4a, b). Initially, GeSbS bulk glass was synthesised from high-purity elements using a mature melt-quenching technique and further purified using an improved physical and chemical purification method44. Secondly, the high-purity GeSbS, which was the deposition material source, was thermally deposited on a silicon wafer with a 3-μm SiO2 layer, followed by thermal annealing under an inert atmosphere. Then, the pattern was imprinted on the mask layer using electron-beam lithography (EBL). After development, a thermal reflow process was applied to remove the roughness of the pattern sidewalls. Next, CF4/CHF3/Ar gas-based inductively coupled plasma reactive ion etching (ICP-RIE) was used to transfer the pattern onto the GeSbS layer. Finally, a 3-μm silica layer was deposited on the top as cladding by inductively coupled plasma chemical vapour deposition (ICP-CVD). Two microresonators with typical free spectral ranges (FSR) of 200 and 6.9 GHz were fabricated using the abovementioned fabrication process. For the microresonator with a FSR of 200 GHz shown in Fig. 4ce, a typical resonant-loaded linewidth of ≈56 MHz was measured in the critical coupling regime using the spectral scanning method, indicating an intrinsic Q-factor of 3.43 × 106. The mean intrinsic Q-factor of the TE00 mode was ≈ 1.80 × 106. For the microresonator with a FSR of 6.9 GHz shown in Fig. 4fh, the measured intrinsic Q-factor of the TE00 mode was as high as 6.48 × 106, and the average intrinsic Q-factor was ≈ 2.36 × 106. Furthermore, the Q factor of the microresonator was insensitive to the wavelength, exhibiting a low loss over a wide range in the communication band (Fig. 4e, h).

      Fig. 4  Fabrication and characterisation of high-performance ChG photonic integrated devices. Adapted with permission from Ref. 34, copyright 2023, WILEY. a Fabrication process of high-performance ChG photonic integrated devices. b A photograph of a fabricated 1 × 2 cm GeSbS photonic chip. c, f The measured transmission spectra of the TE00 mode and the corresponding Lorentzian fits of two microresonators with FSR ≈ 200 and 6.9 GHz, respectively. d, g Histograms of intrinsic Q-factors. e, h Wavelength dependence of the loaded Q-factor and intrinsic Q-factor of the TE00 mode from 1540–1620 nm. i Optical micrograph of a GeSbS spiral waveguide. j Measured insertion losses of GeSbS spiral waveguides.

      Our modified fabrication technology can also meet the requirements of some integrated nonlinear applications, such as integrated travelling-wave parametric amplifiers45,46 and stimulated Brillouin scattering (SBS) filters47,48, which require nearly metre-long photonic circuits. We fabricated spiral waveguides with different lengths (7, 14, 20, 25, and 30 cm) using the proposed fabrication process. An optical micrograph of a spiral waveguide with a length of 7 cm is shown in Fig. 4i. The insertion loss was measured, and a transmission loss as low as 0.2 dB/cm was obtained by linear fitting (Fig. 4j). As shown in Tables 1 and 2, we compared the losses of the microresonators and waveguides reported in previous studies using different ChG materials and methods. The results indicated that our GeSbS photonic integrated devices achieved the highest Q-factor and lowest optical loss among those of the state-of-the-art ChG microrings and planar waveguides used for comparison (order of tens of centimetres).

      MaterialFabrication methodWavelength (μm)Dimension (width×height) (μm)Q-factorRef
      As2S3 microresonatorTrapezoidal-TE1.5510 × 1.31.44 × 10741
      As2Se3 microdiskTE, Lift-off5.22.5 × 1.12 × 10549
      Ge28Sb12Se60 microdiskTE, ICP1.551 (height)5 × 10550
      Ge11.5As24Se64.5 microdiskTE, RIE1.551 (height)1.1 × 10651
      Ge23Sb7S70 microdiskTE, RIE1.550.8 × 0.451.2 × 10652
      Ge28Sb12Se60 microringTE, ICP1.550.3 (height)4.1 × 10553
      As20S80 microringMicro-trench, EBE1.552.0 ×1.56 × 10554
      Ge25Sb10S65 microringTE, ICP-RIE1.552.4 × 0.81.97 × 10638
      Ge25Sb10S65 microringTE, ICP-RIE1.552.4 × 0.82.2 × 10634
      Ge25Sb10S65 microringTE, ICP-RIE1.552.4 × 0.856.48 × 106This work
      TE: thermal evaporation; ICP: inductively coupled plasma; RIE: reactive ion etching; EBE: electron beam evaporation.

      Table 1.  Fabrication method, working wavelength, dimensions, and Q-factors of several typical chalcogenide microresonators

      MaterialFabrication methodWavelength (μm)Dimension (width×height×length)Propagation lossRef
      Ge11.5As24Se64.5TE, ICP-RIE1.550.63 μm × 0.5 μm × 0.18 cm2.6 dB/cm55
      Ge23Sb7S70TE, RIE1.550.8 μm × 0.42 μm × 1 cm0.5 dB/cm56
      Ge11.5As24Se64.5TE, ICP3 - 54 μm × 2.5 μm × 1.4 cm~0.5 dB/cm57
      Ge11.5As24Se64.5TE, ICP3.8 - 54.0 μm × 4.4 μm × 1.8 cm~0.6 dB/cm27
      As2S3TE, ICP21.2 μm × 0.6 μm × 2.4 cm1.45 dB/cm58
      As2Se3TE, wet etching8.45.4 μm × 4.53 μm× 3.55 cm0.5 dB/cm59
      As2Se3Sputter, Lift-off3.510 μm × 1 μm × 5 cm0.16 dB/cm60
      Ge25Sb10S65TE, ICP-RIE1.552.5 μm × 0.7 μm × 30 cm0.2 dB/cmThis work
      TE: Thermal evaporation; ICP: inductive coupled plasma; RIE: reactive ion etching.

      Table 2.  Fabrication method, working wavelengths, dimensions, and propagation losses of several typical chalcogenide waveguides

    • Mode-locked optical frequency combs have revolutionised a wide variety of applications including optical atomic clocks61-64, coherent communication systems65-69, microwave and optical frequency synthesis1,70, molecular footprint detection71-75 and light detection and ranging (LiDAR)76,77. In particular, integrated frequency combs based on Kerr nonlinear microresonators (microcombs) are an attractive choice for realising frequency comb sources with a high coherence, chip-scale sizes, and low power consumption at the milliwatt level78,79. In recent decades, dissipative Kerr soliton microcombs have been demonstrated in various photonic integrated platforms, greatly promoting the development of nonlinear optics and supporting new developments in microcomb applications. In contrast, there is motivation to shift focus from exploring telecommunication bands to seeking new materials for simultaneously realising ultrahigh nonlinear efficiencies based on both material nonlinearities and high Q factors, as well as new broadband spectral windows. Microcombs have been demonstrated in the mid infrared (MIR) bands of silicon and Si3N4 microresonators80,81; the spectral windows of more than 4 μm in these microcombs are currently limited by the SiO2 substrate and cladding of the current platforms8. Additionally, there were no reports of optical frequency combs based on ChG photonic devices till the development of GeSbSs. We conducted a series of studies on the generation and manipulation of microcombs based on GeSbS microresonators.

    • Precise dispersion engineering and high Q factors strongly support the role of microresonators in the generation of third-order nonlinear optical processes and Kerr microcombs. The dispersion of GeSbS microresonators was tailored using geometric parameters and theoretically calculated using the finite element method. The integrated dispersion Dint can be used to characterise the full-order dispersion of GeSbS microresonators; this parameter can be defined as82:

      $$ {D}_{int}={\omega }_{\mu }-{\omega }_{0}-{D}_{1}\mu =\frac{{D}_{2}{\mu }^{2}}{2!}+\frac{{D}_{3}{\mu }^{3}}{3!}+\sum _{m > 3}\frac{{D}_{m}{\mu }^{m}}{m!} $$ (1)

      where μ and ωμ are the relative mode numbers and angular frequencies of the resonances, respectively; D1/2π is the FSR; D2 is the second-order microresonator dispersion; and D3 is the third-order microresonator dispersion. The measured integrated dispersion curves were characterised by calibrating the resonant frequency of the microresonators using a fibre-based Mach-Zehnder interferometer (MZI) and removing the offset frequency and linear dispersion term, which were consistent with the simulation results (Fig. 5a, b). The second-order dispersions were calculated to be −8.2 and 2.1 MHz for the TE00 mode and TM00 mode, respectively, in the GeSbS microresonators with a radius of 100 μm and a cross-section of 2.4 × 0.8 μm (width × height).

      Fig. 5  Dispersion engineering of the two orthogonal polarization fundamental modes in the high-Q microresonator and low threshold OPO generation. Adapted with permission from Ref. 34, copyright 2023, WILEY. a, b The calculated and measured integrated dispersion (Dint) values for the TE00 and TM00 modes with a radius of 100 μm and a cross-section of 2.4 × 0.8 μm (width × height). The second-order dispersions (D2/2π) are −8.2 MHz (normal dispersion) and 2.1 MHz (anomalous dispersion). The insets are the optical field distributions in the cross-section. c, d The pump power dependence of the 1st OPO sideband, revealing low power thresholds of 1.3 and 3.5 mW. The insets are the experimentally measured output power spectra of the OPO sidebands.

      The threshold of optical parametric oscillation is estimated by34:

      $$ {P}_{th}=\frac{\pi }{8}\frac{n}{{n}_{2}}\frac{{v}_{0}}{{v}_{FSR}}\frac{{A}_{eff}}{{Q}_{i}^{2}}\frac{{\left(1+{\textit κ} \right)}^{3}}{{\textit κ} } $$ (2)

      where n is the linear refractive index; n2 is the nonlinear refractive index; v0 is the pump frequency; vFSR is the FSR of the microresonator; Aeff is the effective mode area of the microresonator; and the coupling factor κ = κex/κi, where κex is the coupling rate, and κi is the intrinsic rate of the microresonator. The extract intrinsic Q-factors of the fabricated microresonators were 1.6 × 106 and 2.3 × 106 for the TM00 and TE00 modes, respectively, contributing to lower power requirements for triggering the OPO. The pump thresholds for the TM00 and TE00 modes were 1.3 and 3.5 mW, respectively, as measured by the power dependence of the generated first-order FWM sidebands with the input pump power (Fig. 5c, d).

    • Owing to the distinct dispersion characteristics of the two fundamental polarised modes in the GeSbS microresonator, bright soliton and dark-pulse microcombs could be achieved in a single GeSbS microresonator, as shown in Fig. 6a. With an input power of ~ 20 mW, a bright dissipative Kerr microcomb was experimentally obtained by pumping the TM00 mode family, which featured a broad spectrum ranging from 1440 to 1680 nm and a comb repetition rate of 197 GHz, as shown in Fig. 6b. Due to the low thermo-optic coefficient (TOC) (e.g., dn/dT ≈ 3.1 × 10−5 K−1) of GeSbS, the soliton microcomb could be stably captured by manually sweeping the pump frequency from blue detuning to red detuning without any complicated pumping schemes79,83-88. Moreover, a dark-pulse microcomb assisted by the avoided mode-crossing effect (AMX) was demonstrated in the same microresonator. In general, the origin of the AMX is the multimode nature of ChG microresonators and random fabrication imperfections, which cause linear coupling between different mode families. Only ~ 25 mW of input power was used to drive the dark-pulse microcomb with a microcomb bandwidth of 80 nm (1510–1590 nm) and a repetition rate of 200 GHz, as shown in Fig. 6d. The low pump power requirements of both microcomb generations facilitated the full integration of our chip with commercial on-chip DFB lasers. Furthermore, the formation of dissipative solitons could dramatically reduce the comb intensity noise, as shown in the radio frequency (RF) spectra collected from the electrical spectrum analyser (ESA) in Fig. 6c, e. In combination with well-developed self-injection-locking technologies, the performance of versatile bright and dark soliton microcombs generated in GeSbS microresonators may be further improved and operated in a turnkey scheme, which is favourable for practical applications36, 78, 89, 90.

      Fig. 6  Low-power bright and dark soliton microcombs based on integrated GeSbS microresonators. Adapted with permission from Ref. 34, copyright 2023, WILEY. a Schematic illustration of soliton microcomb generations in an integrated ChG microresonator, including a dark-pulse soliton in the TE00 mode and a bright soliton in the TM00 mode. b, d Output power spectra as the pump frequency sweeps across the resonances, with the soliton transition steps. c, e Experimental output optical spectra of a bright soliton in TM00 mode families and a dark-pulse microcomb in TE00 mode families.

    • We also systematically studied the interplay between stimulated Raman scattering (SRS) and the Kerr nonlinear process of GeSbS microresonators (Fig. 7a, b)38. Owing to the broadband Raman gain spectra of GeSbS materials38, the significant Raman effect affects the formation of Kerr microcombs, especially in the weak anomalous dispersion and normal dispersion regime91,92. The underlying physical mechanism that mediates the nonlinear interaction of Raman lasers and Kerr combs can be attributed to the degenerate FWM process among the pump, Stokes, and second-order Stokes resonant modes and the nondegenerate FWM process involving the pump wave, first and second Stokes waves, and anti-Stokes wave, as shown in Fig. 7c. If the cavity dispersion of a microresonator is improperly designed, it results in large frequency mismatching ($ \mathrm{\Delta }v $) among the three participant resonant modes, which hinders the efficiency of the FWM process and results in the cascading of stimulated Raman scattering (SRS) (Fig.7d). Consequently, the spontaneous FWM process dominates and leads to the formation of Raman-Kerr frequency combs. Otherwise, if the frequency mismatch is sufficiently small, the cascading of SRS occurs93-96. Therefore, a broadband Raman-Kerr microcomb spanning 1450–1700 nm was obtained in the GeSbS microresonators by pumping with an input power of 30 mW, as shown in Fig. 7e. A broadband Raman gain of approximately 1630 nm simultaneously extended the spectral range and enhanced the comb power, providing potential high-power microcomb sources for molecular spectroscopy. The results provided a clear insight into the Kerr-Raman nonlinear effect in integrated ChG microresonators and showed that using this effect was a viable approach for future Kerr microcombs in the near-IR and MIR regions.

      Fig. 7  Broadband Raman-Kerr comb in GeSbS microresonators. Adapted with permission from Ref. 38, copyright 2022, WILEY. a A schematic of engineered Raman lasing in GeSbS microresonators. b Raman response spectrum of a GeSbS film. c Schematic diagram of the degenerate four-wave mixing (dFWM) between a pump wave and the first- and second-order Stokes waves. d Calculated phase mismatch curves for the TE00 mode of two microresonators with microresonator widths of 1.7 and 2.4 μm. The film thickness is fixed at 0.8 μm. e Measured broadband Raman-Kerr comb in the GeSbS microresonator with a width of 2.4 μm.

    • An octave-spanning microcomb enables an f-2f self-reference scheme to completely stabilise the optical frequency comb, which is critical for timing, metrology, and molecular applications11, 62, 97, 98. Soliton Cherenkov radiation or dispersive wave (DW) emission is a key route for reliably extending the bandwidths of optical frequency combs in photonic integrated microresonators62. Generally, in the absence of third- and higher-order dispersions, a soliton microcomb spectrum features a smooth sech2 envelope corresponding to a bright soliton pulse in the time domain. In a small group velocity dispersion (GVD) region, the higher-order dispersion dominates the second-order dispersion, leading to the reshaping of the microcomb spectrum. The pump power can be efficiently transferred to comb lines far from the pump frequency if the phase-matching condition is satisfied. Considering the integrated dispersion in microresonators, the generation of DWs is spectrally located at Dint = 082. Currently, broadband microcombs, especially those with DWs, suffer from the deterioration of comb spectral flatness owing to the large phase mismatch in certain spectral ranges99. Mode hybridisation in waveguides and microresonators has enabled advanced dispersion engineering through fine geometrical tuning, which has been utilised for bright soliton microcombs in normal-dispersion Si3N4 microresonators100 and two-colour soliton microcomb generation101,102. Thus, to further extend the bandwidths and flatten the spectra of microcombs, we propose a GeSbS microresonator with dual-ring microresonators (DRMs) to optimise microcavity dispersion over a wide spectral range using mode hybridisation (Fig. 8a)99. We systematically investigated arbitrary dispersion shaping at specific wavebands and engineered octave frequency combs based on mode hybridisation in mutually coupled concentric microresonators99. When the optical path lengths (OPL) of the inner and outer microresonators were equal, the eigenfrequencies of the antisymmetric supermode in the DRMs were shifted owing to mode hybridisation, generating a new local anomalous dispersion window in the strong normal dispersion region. As a result, an octave-spanning Kerr soliton microcomb with multidispersive waves was numerically achieved with an input power of 40 mW based on the Lugiato-Lefever equation (LLE)103-105, covering wavelengths from 1224 to 2913 nm, as shown in Fig. 8b. The FSR of the dual-ring microresonator was approximately 620 GHz. Moreover, the number of comb lines within the bandwidth of −40 dB could reach 126, and the total comb power was ~ 1.7 mW. The integrated dispersion profile can be flattened over a broadband wavelength range by tailoring the geometric parameters of DRMs. The flexible structures of DRMs enhance the dynamics of frequency combs in GeSbS platforms and enable the generation of octave soliton microcombs with dispersion-engineered multiple-dispersive waves.

      Fig. 8  Broadband microcomb generation in dispersion-engineered chalcogenide dual-ring microresonators. Adapted with permission from Ref. 99, copyright 2023, FRONTIERS. a Schematic of a dual-ring microresonator. b A microcomb with multiple dispersive waves at a red detuning of −20 κ (Upper panel) and a beyond-octave microcomb at a detuning of −32 κ (Lower panel) in the DRMs.

    • Regarding AO modulators, they enable the manipulation of confined photons in photonic materials through the tuning of the refractive indices of the materials using radio frequency (RF)-driven acoustic waves106,107. Traditional AO devices based on bulk materials suffer from weak light and acoustic wave confinement, which leads to AO modulators having high pump powers108. Surface acoustic waves (SAWs) can be well-confined near photonic integrated circuits (PICs), such as integrated waveguides and microresonators, and exhibit a high-energy overlap within a wavelength-scale chip109,110. Thin-film lithium niobate (TFLN) has attracted increasing interest for the realisation of high-performance AO modulators in PICs, owing to its superior advantages in electro-optical conversion, including a strong electro-optic effect, a large refractive index, and a wide transparency wavelength111,112. However, owing to acoustic leakage, the modulation efficiencies of AO modulators become weak, which is a bottleneck in TFLN-based AO modulators. Suspended LN acoustic resonator construction is commonly used to enhance the overlap coefficient between optical and acoustic modes; this method requires strict integrated fabrication processing. Being a typical ChG, GeSbS is resistant to acidic environments and can be easily etched using alkaline reagents. Therefore, a GeSbS waveguide can be suspended by the HF etching of a SiO2 layer, enabling higher-frequency AO modulation. Indeed, suspended structures can enhance the bounds of acoustic modes and improve the AO overlap in a waveguide because optical modes in common optical-material-based waveguides can be bound by total internal reflection. However, acoustic modes can leak into substrates because the acoustic wave velocities in waveguides resemble those in the substrates, resulting in reduced AO overlaps in the waveguides113. However, GeSbS has a high refractive index and a low acoustic velocity (~2.6 km/s), allowing the high confinement of the optical and acoustic modes within a waveguide without suspended structures114,115. Thus, high-frequency (higher than 1 GHz) AO modulation can also be achieved by changing a interdigital electrode structure while preserving the current waveguide structure. Moreover, the refractive indices of Ge25Sb10S65 (n=2.23) and TFLN (ne=2.13) at 1550 nm were similar. In addition, the photoelastic coefficients of GeSbS were anisotropic, and those of p11, p12 and p44 were 0.25, 0.24 and 0.05, respectively. Therefore, we have proposed a hybrid TFLN-ChG integrated waveguide, which benefits from the electrical, optical, and material advanced properties of TFLN and ChG thin films, for high-efficiency integrated AO modulators24,116.

      We extracted the Vπ of an AO modulator to quantitatively evaluate the AO modulation characteristics based on non-suspended TFLN-ChG MZIs from a measured opto-acoustic S21 spectrum, as expressed through the following equation117:

      $$ {V}_{\pi }=\frac{\pi {R}_{PD}{I}_{\text{rec}}}{\left|{S}_{21}\right|} $$ (3)

      where RPD is the sensitivity of the photoreceiver, and Irec is the DC optical power at the bias point with a π/2 phase difference between the two arms of the MZI. Amorphous ChGs have nearly two times higher photoelastic coefficients than those of TFLNs, potentially providing smaller Vπ values in TFLN-ChG waveguide-based AO devices16,118,119.

      To improve the modulation efficiency, a high Q-factor microresonator with an interdigital transducer (IDT) was used to realise strong interactions between the photons and phonons (Fig. 9a, c)116. To exploit the photoelastic effect of the ChG material, we designed a ChG waveguide to confine most of the optical energy (Fig. 9b). The intensities of the three peaks at frequencies of 0.843, 0.88, and 1.464 GHz in the S21 spectrum were more than −50 dB (Fig 9d)116. As a result, the calculated Vπ was 1.74 V, corresponding to a half-wave-voltage-length product of 0.02 V·cm, which indicated that AO conversion had been enhanced at 0.843 GHz.

      Fig. 9  Efficient AO modulation using a microring resonator on a TFLN-ChG hybrid platform. Adapted with permission from Ref. 116, copyright © The Optical Society. a Schematic of the AO modulator with microresonator modulation configurations. b Electric field of the fundamental TE mode. c Lorentz fitting (red curve) of the resonance dip at 1551.198 nm corresponding to a loaded Q-factor of 5 × 105. d Acoustic S11 and S21 spectra of the AO modulators with the microresonator modulation configuration, in which the Irec amplified by EDFA is −5.6 dBm, the S21 is −16.52 dB at 0.843 GHz, and RPD is 300 V/W.

      To improve the modulation efficiency of a MZI-based AO modulator consisting of waveguide arms, we further propose a built-in push-pull AO modulation structure (called a double-arm-modulated device) based on an anti-symmetric SAW mode (Fig. 10a, b). The geometric dimensions of the waveguide were optimised using an antisymmetric Rayleigh SAW (Fig. 10c, f). The optimised IDT was placed in the two waveguide arms of a MZI, which effectively utilised the energy of the acoustic waves in both directions to achieve more efficient AO modulation (Fig. 10d, e). Moreover, we concentrated approximately 83% of the optical energy in the ChG waveguides, significantly improving the AO overlap factor (Fig. 10f). As a result, in the case of a sound phase difference up to π, the modulation efficiency was twice that of the single-arm-modulated device (Fig. 10h)24.

      Fig. 10  Highly-efficient AO modulation using nonsuspended built-in push-pull structure. Adapted with permission from Ref. 24, copyright 2022, NPG. a Schematic diagram of the proposed device. The thickness of the TFLN is fixed at 400 nm, and the thickness of the oxide layer is 2 μm. b Cross-section image of the proposed device. c SEM image of the cross-section of the modulated waveguide. d SEM image of the modulation area, including the IDT and waveguide. e Microscopic image of a TFLN-ChG hybrid MZI-based AO modulator. f Relation between the energy confinement factor Γ in the ChG waveguide and waveguide geometry. The optimized H and W of the waveguides are chosen as 850 nm and 2.05 μm, respectively. g Numerical simulation results of the dominant Sxx strain components of the SAW modes in the heterogeneous-integration waveguide platform with the double arm modulation configuration. h S21 spectra of the TFLN-ChG hybrid MZI-based AO modulators with single-arm and double-arm modulation configurations.

      By properly designing the IDT structure with odd fingers, an opposite refractive-index change can be achieved in the double-arm waveguide, maximising the conversion between the acoustic and optical waves. For instance, the SAW modes at 0.833 GHz and 0.805 GHz corresponded to 5.5 (number of fingers NIDT=11, odd) and 5 (NIDT=10, even) pairs of IDT fingers, respectively (Fig. 10g, h). Therefore, an IDT with 50.5 pairs of fingers (NIDT=101) was designed to realize the effective excitation of the SAWs. Moreover, compared to our single-arm-modulated device, the intensity of the frequency peak at 0.844 GHz in the optoacoustic S21 spectrum of the double-arm-modulated device increased by 12 dB (Fig. 10h). The modulation efficiency of the push-pull AO modulator was up to 0.03 V·cm (VπL), which was three times that of the single-arm modulation device.

      We compared our AO modulators, including the single-arm-modulated and double-arm-modulated ones based on the non-suspended TFLN-ChG hybrid waveguides, with those based on the TFLN in Table 3. Our results showed the highest AO modulation efficiency based on a non-suspended TFLN waveguide. Presently, the on-chip loss of our devices was as high as 5 dB, which was mainly due to the four compact 90° bending waveguides. Additionally, AO efficiency depends on the material density120, and annealing can change the density, thereby affecting the AO coefficient. To date, we have not performed experiments to compare the AO changes before and after annealing. We plan to further optimise the device structure and fabrication process to achieve a higher AO modulation efficiency.

      PlatformAcoustic cavityFrequency (GHz)1-|S11|2 (%)L (μm)αp (rad/√mW)VπL (V cm)Ref./device
      LNa0.114212000.0732.5121
      As2S3/SiO2/LNa0.119524000.260.94122
      LNb3.33641000.270.046123
      LNb1.1619.3450.540.019117
      LN×1.950450.0170.38117
      LN×1.990450.0180.27117
      SA: Ge25Sb10S65/LN×0.8498.51200.120.124
      DA: Ge25Sb10S65/LN×0.8498.51200.40.0324
      a In-plane metal grating reflectors have been fabricated to construct an acoustic cavity. b: Suspended TFLN is etched as an acoustic cavity.

      Table 3.  Comparison of modulation metrics for TFLN MZI-based AO modulators24.

    Conclusion and outlook
    • In this review, we have discussed recent progress in the improved fabrication of hybrid integrated ChG photonic devices and related integrated applications based on Kerr nonlinearity and the AO effect. High-Q integrated microresonators and low-loss planar waveguides were achieved through improved hybrid chip fabrication, including the preparation of a GeSbS thin film and an improved fabrication process compatible with that of silicon photonics. The advantages of low loss, high nonlinearity, superior AO interactions, and versatile Kerr microcomb interactions were demonstrated in the generation of versatile Kerr microcombs and highly-efficient integrated AO modulators.

      To date, this novel hybrid ChG–silicon device is still in its early stages of development. There is ample room to apply the potential of this device in broadband spectral coverage, high nonlinearity, and photoelastic effects, as well as multifunctional hybrid or heterogeneous integration with other material platforms. The fabrication of this device should reach a point where reproducible, high-yield, and wafer-scale processing can be accomplished while increasing the Q-factor up to the material absorption limit. Moreover, low-loss ‘all-ChGs’ waveguides with both ChGs as the core and claddings need to be developed for flexible dispersion engineering and extending operating wavelengths to the MIR region. Additionally, owing to their low-temperature deposition features and compatibility with silicon photonic fabrication, integrating active materials, such as III-V materials and rare-earth elements, with ChG photonic circuits, can open up opportunities for realising narrow-linewidth lasers, on-chip amplifiers, and fully integrated microcomb sources. We believe that ChG-based photonic integration will revolutionise existing photonic applications through the development of a new class of devices in the near-IR-to-MIR spectral range.

    Acknowledgements
    • This work was supported by the National Key R&D Program of China under Grant (2019YFA0706301), the Key Project in Broadband Communication and New Network of the Ministry of Science and Technology (MOST) (2018YFB1801003), the National Science Foundation of China (NSFC) (U2001601, 61975242, 61525502, 62175095, 62375292), and the Natural Science Foundation of Guangdong Province for Distinguished Young Scholars (2023B1515020028).

Reference (123)

Catalog

    /

    DownLoad:  Full-Size Img PowerPoint
    Return
    Return