Abstract
All-optical modulators hold significant prospects for future information processing technologies for they are able to process optical signals without the electro-optical convertor which limits the achievable modulation bandwidth. However, owing to the hardly-controlled optical backscattering in the commonly-used device geometries and the weak optical nonlinearities of the conventional material systems, constructing an all-optical modulator with a large bandwidth and a deep modulation depth in an integration manner is still challenging. Here, we propose an approach to achieving an on-chip ultrafast all-optical modulator with ultra-high modulation efficiency and a small footprint by using photonic topological insulators (PTIs) made of metallic quantum wells (MQWs). Since PTIs have attracted significant attention because of their unidirectional propagating edge states, which mitigate optical backscattering caused by structural imperfections or defects. Meanwhile, MQWs have shown a large Kerr nonlinearity, facilitating the development of minimally sized nonlinear optical devices including all-optical modulators. The proposed photonic topological modulator shows a remarkable modulation depth of 15 dB with a substantial modulation bandwidth above THz in a tiny footprint of only 4 × 10 µm2, which manifests itself as one of the most compact optical modulators compared with the reported ones possessing a bandwidth above 100 GHz. Such a high-performance optical modulator could enable new functionalities in future optical communication and information processing systems.
1 Introduction
All-optical modulations have emerged as one of the most essential ingredients for integrated photonics due to their ultrafast operation speed [1], [2], [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13]. Material systems with a relatively large Kerr effect in various spatial configurations, such as the Mach–Zehnder interferometer (MZI) [14], [15], [16] and the ring-assisted MZI [17], [18], have commonly been utilized for the all-optical modulators. Nevertheless, these all-optical modulators suffer from (i) a large device footprint due to the weak optical nonlinearities of the conventional material systems – which limits their application for dense photonic integrated circuits, and (ii) a strong optical backscattering owing to the unavoidable structure defects of the commonly-used device geometries – which results in a small modulation depth. Therefore, new photonic platforms with a robust optical transport and a small feature size are in urgent need for the next-generation high-performance all-optical modulators.
Photonic topological insulators (PTIs) have recently emerged as an intriguing photonic platform that supports a robust optical transport. They are characterized by the presence of unidirectional propagating edge modes which exhibit remarkable robustness against imperfections or defects, thus effectively preventing optical backscattering [19]. Such a robustness feature has enabled an unprecedented evolution in various optical signal processing processes such as signal non-reciprocal transport [20], [21], [22], routing [23] and isolation [24]. As a result, numerous innovative photonic devices have been conceptualized and demonstrated ranging from reflection-free sharply waveguides [25], [26], spin-polarized switches [27], [28], to non-reciprocal circulators and travelling wave amplifiers [29], [30]. However, limited also by the weak optical nonlinearities of commonly-used materials, PTIs based all-optical modulators have not been fully explored.
Many recent efforts have made use of metallic quantum wells (MQWs) to achieve giant ultrafast optical nonlinearities due to the quantum size effect [31], [32], [33], [34]. In this work, we propose and demonstrate an on-chip ultrafast all-optical modulator by using PTIs made of these MQWs. The MQW-based all-optical PTI modulator shows a remarkable modulation depth of 15 dB with a substantial modulation bandwidth above THz [35], and its size is only 4 × 10 µm2 – a great advantage for large-scale integration (LSI) photonic integrated circuits. Moreover, the device size can be further reduced to 1.3 × 10 μm2 at the 3-dB modulation depth with a transmission of −1.68 dB. Such an all-optical modulator holds significant promise for advancing the field of nonlinear topological photonics and expanding its applicability across diverse domains, including optical communications, microwave photonics, and quantum information processing.
2 Results
Figure 1(a) shows the MQW-based PTI used for the proposed all-optical modulator. It is a honeycomb lattice of cylindrical pillars made of TiN/Al2O3 MQWs. The optical nonlinear PTI is designed to have two different topological modes depending on the unit geometry and its refractive index. The intensity-dependent refractive index of the MQWs is adapted from our previous works [33], see details in Supplementary Material Section 1. Without laser pumping, the MQWs are a lossy metal so that PTI in both unit geometries, i.e. the expanded one with a larger distance between adjacent cylindrical pillars R 1 and the shrunken one with a smaller distance between adjacent cylindrical pillars R 2, is a plasmonic structure. In this case, the surface plasmon polariton (SPP) mode is supported at the PTI surface, and its center wavelength is designed to located at the working wavelength around 2.0 μm in order to restrict the signal transmission (see the upper panel of Figure 1(b)). The signal transmission is allowed on the PTI surface under laser pumping (see the lower panel of Figure 1(b)), owing to the high-intensity low loss feature of the MQWs [33]. In the case of pumping, the expanded PTI is designed to have a nontrivial band structure, while the shrunken PTI has a trivial band structure; so, by adjoining the two PTI structures, a zigzag-shaped topological interface is formed where topologically protected pseudospin-dependent edge modes are enabled [36], which can efficiently carry the signal. Figure 1(c) shows the intensity dependence of the signal transmission: Around −21.3 dB without laser pumping while around −5.8 dB under pumping. Such a huge transmission contract (∼15 dB) is essential for an all-optical modulator, as demonstrated in what follows.

MQW-based PTI used for the proposed all-optical modulator. (a) Honeycomb lattice of cylindrical pillars made of TiN/Al2O3 MQWs with two different unit geometries: the expanded one with a/R 1 = 2.75 and the shrunken one with a/R 2 = 3.65, where the lattice period a = 1.58 µm, and the distances between adjacent cylindrical pillars of two different unit geometries R 1 = 0.57 µm and R 2 = 0.43 µm. In both cases, the height of the pillar h = 1 µm, and its diameter d = 0.37 µm. (b) Spectra for expanded and shrunken PTIs without (the upper panels) and with (the lower panels) laser pumping. The horizontal axis is the incidence angle of the incident signal beam onto the PTI waveguide, while the angle of the pump beam remains 45° with respect to the PTI waveguide. Color encodes the magnitude of absorptance (A) of these PTIs for p-polarized incident light. The dashed squares mark the band gap of PTI. Simulation details are presented in Supplementary Material Section 2. (c) Transmittance T = P y_S1/P y_S2 at the wavelength of 2.0 µm as a function of the pump intensity, where P y is the y component of the ponying vector, S1 and S2 are two distant cutting planes, as marked in (a).
Figure 2(a) shows the proposed PTI all-optical modulator, which is sandwiched by two Si topological waveguides. The two Si topological waveguides are designed to maximize the coupling efficiency at the interface to PTI (see details in Supplementary Material Section 3). Figure 2(b) and (d) show the simulation results of P y -field distribution of the right-circularly polarized (RCP) signal light at the wavelength of 2.0 µm without and with laser pumping, respectively, and P y is the y component of the Poynting vector. In the absence of a pumping laser, the input signal light is localized at the PTI interface (Figure 2(b)), indicating that the signal is not allowed to pass the PTI modulator; in this case, the modulator is in the “OFF” state. It is in the “ON” state under laser pumping, and the input signal is then allowed to pass the zigzag-shaped topological interface of PTI, as shown in Figure 2(d). Therefore, these simulation results show clearly that the proposed PTI modulator can be used to efficiently modulate the RCP signal light by alternating the pumping laser. Moreover, the topological protection feature of PTI enables the modulator to completely suppress the transmission of the left circularly polarized (LCP) light, as shown in Figure 2(c), which thus prevents the backscattering of the RCP signal light. Further studies could explore the comparative back scattering properties of topological versus standard waveguide structures to provide a clearer understanding of the suppression mechanisms involved.

All-optical modulator with PTIs made of MQWs. (a) Schematic of the proposed MQW-based PTI all-optical modulator sandwiched by two Si topological waveguides. The input signal, guided by the input Si topological waveguide, undergoes modulation at the PTI modulator by a modulated pumping laser before entering the output Si topological waveguide. (b) P y -field distribution of the RCP signal light without laser pumping. (c, d) P y -field distribution of the LCP (c) and RCP (d) signal light under laser pumping, respectively.
For an all-optical modulator, the transmittance and modulation depth are two of the key performance metrics. Figure 3(a) summarizes the wavelength dependence of the transmittance for the RCP signal light with the proposed PTI all-optical modulator. Without laser pumping, the transmittance remains low in the wavelength range around 2.0 μm – due to the excitation of lossy SPP modes (Figure 1(b)) – with a minimum value at the wavelength of 1.99 μm. Upon laser pumping, the transmittance is significantly increased due to the optically induced metallic-to-dielectric transition of MQWs [33]. In this case, the topological edge modes are responsible for the signal light transmission at the zigzag-shaped topological interface (Figure 1(b)). The calculated modulation depth of the all-optical modulator is summarized in Figure 3(b). The modulation depth reaches 15 dB at the wavelength of 1.99 µm in the topological band gap where the optical backscattering is forbidden. Such a substantial transmittance contract is crucial for the functionality of all-optical modulators. Detailed calculation methods of transmittance and modulation depth are shown in Supplementary Material Section 4. In addition, based on our previous studies of MQWs [37], the modulation speed of the proposed modulator can reach the order of 100 fs, which corresponds to a modulation bandwidth up to the order of 10 THz.

Performance of the proposed PTI all-optical modulator. (a) Wavelength dependence of the transmittance for the RCP signal light without and with laser pumping. (b) Wavelength dependence of the corresponding modulation depth. The dashed squares mark the band gap of the PTI.
PTI-based photonic devices are able to operate at small scales due to their topological nature [35], which is particularly advantageous for the development of compact and integrated optical circuits. To assess the performance of the proposed PTI all-optical modulator under different configurations, the dependence of the number of unit cells and the thickness ratio of Al2O3 to TiN in MQWs on the modulation depth and the transmittance is calculated at the wavelength of 1.99 µm and the results are summarized in Figure 4. It is evident that all the modulation depth is above 6 dB in the available unit cell range at various Al2O3 to TiN thickness ratios (Figure 4(a)), showing the high flexibility in the design of the proposed PTI all-optical modulator. Since a 3-dB modulation depth is sufficient for practical applications, the transmission efficiency becomes the dominant factor in designing the PTI all-optical modulator with a smaller footprint. As shown in Figure 4(b), the transmittance is increased as the Al2O3 to TiN thickness ratio increases and also as the number of unit cells decreases. As the consequence, the proposed PTI all-optical modulator can achieve a footprint down to 1.3 × 10 µm2 (i.e. only one unit cell) at the 6.5-dB modulation depth with a transmittance up to −1.68 dB (at the Al2O3 to TiN thickness ratio of 1.30). All-optical modulators with such a small footprint are suitable for the miniaturization of optical devices in integrated optical networks. It is worth noting that our work, which employs a circularly polarized beam for the PTI waveguide due to pseudo-time-reversal symmetry, can be extended to systems using a linearly polarized beam, such as quantum Hall photonic topological insulators.

Dependence of the performance of the proposed PTI all-optical modulator. (a, b) Modulation depth (a) and transmittance (b) as functions of the number of unit cells and the thickness ratio of Al2O3 to TiN at the wavelength of 1.99 µm. The upper insets show the results at different numbers of unit cells when the Al2O3 to TiN thickness ratio is fixed to 1.2, while the results at different Al2O3 to TiN thickness ratios when only one unit cell is considered are shown in the lower insets.
3 Conclusion
In conclusion, we have presented an on-chip all-optical modulator with PTIs made of MQWs. The all-optical modulator has been designed to have a substantial modulation depth of 15 dB with a modulation bandwidth above THz in a tiny footprint of only 4 × 10 µm2. Such a high-performance all-optical modulator could facilitate the realization of large-scale dense photonic integrated circuits. In regards to the fabrication of the proposed MQW-based PTI all-optical modulator, the growth of high-quality MQWs is critical in the future works. The MQWs-elements consist of multiple pairs of ultrathin TiN and Al2O3 layers, which could be fabricated as the film-stack [32], [33]. Subsequently, the MQWs-element topological structures could be obtained through standard lithography processes such as focused ion beam milling and plasma etching [38]. To this end, MQWs with two-dimensional material heterostructures may be a better alternative, which could further improve the all-optical modulation performance [39], [40], [41].
Funding source: National Key Research and Development Program of China
Award Identifier / Grant number: 2022YFA1404704
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 62005237
-
Research funding: The work at Zhejiang University was sponsored by the National Key Research and Development Program of China under Grant 2022YFA1404704, and the National Natural Science Foundation of China (NNSFC) under Grant No. 62005237.
-
Author contributions: HQ and SL conceived the idea. HW and JN conducted the numerical simulations. HW, SZ, SL, and HQ contributed extensively to the writing of the manuscript. HW, JN, QC, HS, YY, SL, and HQ analyzed data and interpreted the details of the results. HC, SL, and HQ supervised the research.
-
Conflict of interest: Authors state no conflicts of interest.
-
Ethical approval: The conducted research is not related to either human or animals use.
-
Data availability: The datasets generated and/or analyzed during the current study are available from the corresponding author upon reasonable request.
References
[1] V. R. Almeida, C. A. Barrios, R. R. Panepucci, and M. Lipson, “All-optical control of light on a silicon chip,” Nature, vol. 431, no. 7012, pp. 1081–1084, 2004. https://doi.org/10.1038/nature02921.Search in Google Scholar PubMed
[2] D. Pacifici, H. J. Lezec, and H. A. Atwater, “All-optical modulation by plasmonic excitation of CdSe quantum dots,” Nat. Photonics, vol. 1, no. 7, pp. 402–406, 2007. https://doi.org/10.1038/nphoton.2007.95.Search in Google Scholar
[3] C. Koos, et al.., “All-optical high-speed signal processing with silicon–organic hybrid slot waveguides,” Nat. Photonics, vol. 3, no. 4, pp. 216–219, 2009. https://doi.org/10.1038/nphoton.2009.25.Search in Google Scholar
[4] K. Nozaki, et al.., “Sub-femtojoule all-optical switching using a photonic-crystal nanocavity,” Nat. Photonics, vol. 4, no. 7, pp. 477–483, 2010. https://doi.org/10.1038/nphoton.2010.89.Search in Google Scholar
[5] C. Sturm, et al.., “All-optical phase modulation in a cavity-polariton Mach–Zehnder interferometer,” Nat. Commun., vol. 5, no. 1, p. 3278, 2014. https://doi.org/10.1038/ncomms4278.Search in Google Scholar PubMed PubMed Central
[6] J.-H. Chen, B.-C. Zheng, G.-H. Shao, S. J. Ge, F. Xu, and Y. Q. Lu, “An all-optical modulator based on a stereo graphene–microfiber structure,” Light Sci. Appl., vol. 4, no. 12, p. e360, 2015. https://doi.org/10.1038/lsa.2015.133.Search in Google Scholar
[7] P. Guo, R. D. Schaller, L. E. Ocola, B. T. Diroll, J. B. Ketterson, and R. P. H. Chang, “Large optical nonlinearity of ITO nanorods for sub-picosecond all-optical modulation of the full-visible spectrum,” Nat. Commun., vol. 7, no. 1, p. 12892, 2016. https://doi.org/10.1038/ncomms12892.Search in Google Scholar PubMed PubMed Central
[8] M. R. Shcherbakov, et al.., “Ultrafast all-optical tuning of direct-gap semiconductor metasurfaces,” Nat. Commun., vol. 8, no. 1, pp. 1–6, 2017. https://doi.org/10.1038/s41467-017-00019-3.Search in Google Scholar PubMed PubMed Central
[9] G. Grinblat, M. P. Nielsen, P. Dichtl, Y. Li, R. F. Oulton, and S. A. Maier, “Ultrafast sub–30-fs all-optical switching based on gallium phosphide,” Sci. Adv., vol. 5, no. 6, p. eaaw3262, 2019. https://doi.org/10.1126/sciadv.aaw3262.Search in Google Scholar PubMed PubMed Central
[10] G. Grinblat, et al.., “Efficient ultrafast all-optical modulation in a nonlinear crystalline gallium phosphide nanodisk at the anapole excitation,” Sci. Adv., vol. 6, no. 34, p. eabb3123, 2020. https://doi.org/10.1126/sciadv.abb3123.Search in Google Scholar PubMed PubMed Central
[11] M. I. Shalaev, W. Walasik, and N. Litchinitser, “Optically tunable topological photonic crystal,” Optica, vol. 6, no. 7, pp. 839–844, 2019. https://doi.org/10.1364/optica.6.000839.Search in Google Scholar
[12] H. Wang, et al.., “All-optical ultrafast polarization switching with nonlinear plasmonic metasurfaces,” Sci. Adv., vol. 10, no. 8, p. eadk3882, 2024. https://doi.org/10.1126/sciadv.adk3882.Search in Google Scholar PubMed PubMed Central
[13] J. Gao, et al.., “All-optical tunable wavelength conversion in opaque nonlinear nanostructures,” Nanophotonics, vol. 11, no. 17, pp. 4027–4035, 2022. https://doi.org/10.1515/nanoph-2022-0078.Search in Google Scholar
[14] S. Yu, et al.., “All-optical graphene modulator based on optical Kerr phase shift,” Optica, vol. 3, no. 5, pp. 541–544, 2016. https://doi.org/10.1364/optica.3.000541.Search in Google Scholar
[15] K. J. Ooi, P. C. Leong, L. K. Ang, and D. T. H. Tan, “All-optical control on a graphene-on-silicon waveguide modulator,” Sci. Rep., vol. 7, no. 1, p. 12748, 2017. https://doi.org/10.1038/s41598-017-13213-6.Search in Google Scholar PubMed PubMed Central
[16] Y. Wang, et al.., “All-optical phosphorene phase modulator with enhanced stability under ambient conditions,” Laser Photonics Rev., vol. 12, no. 6, p. 1800016, 2018. https://doi.org/10.1002/lpor.201800016.Search in Google Scholar
[17] Z. Qin, et al.., “Black phosphorus as saturable absorber for the Q-switched Er: ZBLAN fiber laser at 2.8 μm,” Opt. Express, vol. 23, no. 19, pp. 24713–24718, 2015. https://doi.org/10.1364/oe.23.024713.Search in Google Scholar PubMed
[18] Z. Luo, et al.., “Two-dimensional material-based saturable absorbers: towards compact visible-wavelength all-fiber pulsed lasers,” Nanoscale, vol. 8, no. 2, pp. 1066–1072, 2016. https://doi.org/10.1039/c5nr06981e.Search in Google Scholar PubMed
[19] L. Lu, J. D. Joannopoulos, and M. Soljačić, “Topological photonics,” Nat. Photonics, vol. 8, no. 11, pp. 821–829, 2014. https://doi.org/10.1038/nphoton.2014.248.Search in Google Scholar
[20] Z. Shen, et al.., “Experimental realization of optomechanically induced non-reciprocity,” Nat. Photonics, vol. 10, no. 10, pp. 657–661, 2016. https://doi.org/10.1038/nphoton.2016.161.Search in Google Scholar
[21] D. L. Sounas and A. Alù, “Non-reciprocal photonics based on time modulation,” Nat. Photonics, vol. 11, no. 7880, pp. 774–783, 2017. https://doi.org/10.1038/s41566-017-0051-x.Search in Google Scholar
[22] Z. Zhang, P. Delplace, and R. Fleury, “Superior robustness of anomalous non-reciprocal topological edge states,” Nature, vol. 598, no. 7880, pp. 293–297, 2021. https://doi.org/10.1038/s41586-021-03868-7.Search in Google Scholar PubMed PubMed Central
[23] L. Lu, H. Gao, and Z. Wang, “Topological one-way fiber of second Chern number,” Nat. Commun., vol. 9, no. 1, p. 5384, 2018. https://doi.org/10.1038/s41467-018-07817-3.Search in Google Scholar PubMed PubMed Central
[24] R. El-Ganainy and M. Levy, “Optical isolation in topological-edge-state photonic arrays,” Opt. Lett., vol. 40, no. 22, pp. 5275–5278, 2015. https://doi.org/10.1364/ol.40.005275.Search in Google Scholar PubMed
[25] K. Lai, T. Ma, X. Bo, S. Anlage, and G. Shvets, “Experimental realization of a reflections-free compact delay line based on a photonic topological insulator,” Sci. Rep., vol. 6, no. 1, p. 28453, 2016. https://doi.org/10.1038/srep28453.Search in Google Scholar PubMed PubMed Central
[26] T. Ma, A. B. Khanikaev, S. H. Mousavi, and G. Shvets, “Guiding electromagnetic waves around sharp corners: topologically protected photonic transport in metawaveguides,” Phys. Rev. Lett., vol. 114, no. 12, p. 127401, 2015. https://doi.org/10.1103/physrevlett.114.127401.Search in Google Scholar
[27] D. Leykam and Y. D. Chong, “Edge solitons in nonlinear-photonic topological insulators,” Phys. Rev. Lett., vol. 117, no. 14, p. 143901, 2016. https://doi.org/10.1103/physrevlett.117.143901.Search in Google Scholar PubMed
[28] X. Cheng, C. Jouvaud, X. Ni, S. H. Mousavi, A. Z. Genack, and A. B. Khanikaev, “Robust reconfigurable electromagnetic pathways within a photonic topological insulator,” Nat. Mater., vol. 15, no. 5, pp. 542–548, 2016. https://doi.org/10.1038/nmat4573.Search in Google Scholar PubMed
[29] V. Peano, M. Houde, F. Marquardt, and A. A. Clerk, “Topological quantum fluctuations and traveling wave amplifiers,” Phys. Rev. X, vol. 6, no. 4, p. 041026, 2016. https://doi.org/10.1103/physrevx.6.041026.Search in Google Scholar
[30] M. Hafezi, E. A. Demler, M. D. Lukin, and J. M. Taylor, “Robust optical delay lines with topological protection,” Nat. Phys., vol. 7, no. 11, pp. 907–912, 2011. https://doi.org/10.1038/nphys2063.Search in Google Scholar
[31] H. Qian, Y. Xiao, and Z. Liu, “Giant Kerr response of ultrathin gold films from quantum size effect,” Nat. Commun., vol. 7, no. 1, p. 13153, 2016. https://doi.org/10.1038/ncomms13153.Search in Google Scholar PubMed PubMed Central
[32] H. Qian, et al.., “Large optical nonlinearity enabled by coupled metallic quantum wells,” Light Sci. Appl., vol. 8, no. 1, p. 13, 2019. https://doi.org/10.1038/s41377-019-0123-4.Search in Google Scholar PubMed PubMed Central
[33] H. Qian, et al.., “Nanoscale optical pulse limiter enabled by refractory metallic quantum wells,” Sci. Adv., vol. 6, no. 20, p. eaay3456, 2020. https://doi.org/10.1126/sciadv.aay3456.Search in Google Scholar PubMed PubMed Central
[34] A. Mekawy and A. Alù, “Giant midinfrared nonlinearity based on multiple quantum well polaritonic metasurfaces,” Nanophotonics, vol. 10, no. 1, pp. 667–678, 2020. https://doi.org/10.1515/nanoph-2020-0408.Search in Google Scholar
[35] Y. Zhang, et al.., “High-speed electro-optic modulation in topological interface states of a one-dimensional lattice,” Light Sci. Appl., vol. 12, no. 1, p. 206, 2023. https://doi.org/10.1038/s41377-023-01251-x.Search in Google Scholar PubMed PubMed Central
[36] L.-H. Wu and X. Hu, “Scheme for achieving a topological photonic crystal by using dielectric material,” Phys. Rev. Lett., vol. 114, no. 22, p. 223901, 2015. https://doi.org/10.1103/physrevlett.114.223901.Search in Google Scholar PubMed
[37] S. Li, H. Qian, and Z. Liu, “Anomalous nonlinear optical selection rules in metallic quantum wells,” Adv. Funct. Mater., vol. 30, no. 21, p. 2000829, 2020. https://doi.org/10.1002/adfm.202000829.Search in Google Scholar
[38] D. Lu, et al.., “Nanostructuring multilayer hyperbolic metamaterials for ultrafast and bright green InGaN quantum wells,” Adv. Mater., vol. 30, no. 15, p. 1706411, 2018. https://doi.org/10.1002/adma.201706411.Search in Google Scholar PubMed
[39] P. Schmidt, et al.., “Nano-imaging of intersubband transitions in van der Waals quantum wells,” Nat. Nanotechnol., vol. 13, no. 11, pp. 1035–1041, 2018. https://doi.org/10.1038/s41565-018-0233-9.Search in Google Scholar PubMed
[40] Y. Liang, et al.., “Hybrid anisotropic plasmonic metasurfaces with multiple resonances of focused light beams,” Nano Lett., vol. 21, no. 20, pp. 8917–8923, 2021. https://doi.org/10.1021/acs.nanolett.1c02751.Search in Google Scholar PubMed
[41] C. Pan, et al.., “Large area single crystal gold of single nanometer thickness for nanophotonics,” Nat. Commun., vol. 15, no. 1, p. 2840, 2024. https://doi.org/10.1038/s41467-024-47133-7.Search in Google Scholar PubMed PubMed Central
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2024-0197).
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Editorial
- New frontiers in nonlinear nanophotonics
- Reviews
- Tailoring of the polarization-resolved second harmonic generation in two-dimensional semiconductors
- A review of gallium phosphide nanophotonics towards omnipotent nonlinear devices
- Nonlinear photonics on integrated platforms
- Nonlinear optical physics at terahertz frequency
- Research Articles
- Second harmonic generation and broad-band photoluminescence in mesoporous Si/SiO2 nanoparticles
- Second harmonic generation in monolithic gallium phosphide metasurfaces
- Intrinsic nonlinear geometric phase in SHG from zincblende crystal symmetry media
- CMOS-compatible, AlScN-based integrated electro-optic phase shifter
- Symmetry-breaking-induced off-resonance second-harmonic generation enhancement in asymmetric plasmonic nanoparticle dimers
- Nonreciprocal scattering and unidirectional cloaking in nonlinear nanoantennas
- Metallic photoluminescence of plasmonic nanoparticles in both weak and strong excitation regimes
- Inverse design of nonlinear metasurfaces for sum frequency generation
- Tunable third harmonic generation based on high-Q polarization-controlled hybrid phase-change metasurface
- Phase-matched third-harmonic generation in silicon nitride waveguides
- Nonlinear mid-infrared meta-membranes
- Phase-matched five-wave mixing in zinc oxide microwire
- Tunable high-order harmonic generation in GeSbTe nano-films
- Si metasurface supporting multiple quasi-BICs for degenerate four-wave mixing
- Cryogenic nonlinear microscopy of high-Q metasurfaces coupled with transition metal dichalcogenide monolayers
- Giant second-harmonic generation in monolayer MoS2 boosted by dual bound states in the continuum
- Quasi-BICs enhanced second harmonic generation from WSe2 monolayer
- Intense second-harmonic generation in two-dimensional PtSe2
- Efficient generation of octave-separating orbital angular momentum beams via forked grating array in lithium niobite crystal
- High-efficiency nonlinear frequency conversion enabled by optimizing the ferroelectric domain structure in x-cut LNOI ridge waveguide
- Shape unrestricted topological corner state based on Kekulé modulation and enhanced nonlinear harmonic generation
- Vortex solitons in topological disclination lattices
- Dirac exciton–polariton condensates in photonic crystal gratings
- Enhancing cooperativity of molecular J-aggregates by resonantly coupled dielectric metasurfaces
- Symmetry-protected bound states in the continuum on an integrated photonic platform
- Ultrashort pulse biphoton source in lithium niobate nanophotonics at 2 μm
- Entangled photon-pair generation in nonlinear thin-films
- Directionally tunable co- and counterpropagating photon pairs from a nonlinear metasurface
- All-optical modulator with photonic topological insulator made of metallic quantum wells
- Photo-thermo-optical modulation of Raman scattering from Mie-resonant silicon nanostructures
- Plasmonic electro-optic modulators on lead zirconate titanate platform
- Miniature spectrometer based on graded bandgap perovskite filter
- Far-field mapping and efficient beaming of second harmonic by a plasmonic metagrating
Articles in the same Issue
- Frontmatter
- Editorial
- New frontiers in nonlinear nanophotonics
- Reviews
- Tailoring of the polarization-resolved second harmonic generation in two-dimensional semiconductors
- A review of gallium phosphide nanophotonics towards omnipotent nonlinear devices
- Nonlinear photonics on integrated platforms
- Nonlinear optical physics at terahertz frequency
- Research Articles
- Second harmonic generation and broad-band photoluminescence in mesoporous Si/SiO2 nanoparticles
- Second harmonic generation in monolithic gallium phosphide metasurfaces
- Intrinsic nonlinear geometric phase in SHG from zincblende crystal symmetry media
- CMOS-compatible, AlScN-based integrated electro-optic phase shifter
- Symmetry-breaking-induced off-resonance second-harmonic generation enhancement in asymmetric plasmonic nanoparticle dimers
- Nonreciprocal scattering and unidirectional cloaking in nonlinear nanoantennas
- Metallic photoluminescence of plasmonic nanoparticles in both weak and strong excitation regimes
- Inverse design of nonlinear metasurfaces for sum frequency generation
- Tunable third harmonic generation based on high-Q polarization-controlled hybrid phase-change metasurface
- Phase-matched third-harmonic generation in silicon nitride waveguides
- Nonlinear mid-infrared meta-membranes
- Phase-matched five-wave mixing in zinc oxide microwire
- Tunable high-order harmonic generation in GeSbTe nano-films
- Si metasurface supporting multiple quasi-BICs for degenerate four-wave mixing
- Cryogenic nonlinear microscopy of high-Q metasurfaces coupled with transition metal dichalcogenide monolayers
- Giant second-harmonic generation in monolayer MoS2 boosted by dual bound states in the continuum
- Quasi-BICs enhanced second harmonic generation from WSe2 monolayer
- Intense second-harmonic generation in two-dimensional PtSe2
- Efficient generation of octave-separating orbital angular momentum beams via forked grating array in lithium niobite crystal
- High-efficiency nonlinear frequency conversion enabled by optimizing the ferroelectric domain structure in x-cut LNOI ridge waveguide
- Shape unrestricted topological corner state based on Kekulé modulation and enhanced nonlinear harmonic generation
- Vortex solitons in topological disclination lattices
- Dirac exciton–polariton condensates in photonic crystal gratings
- Enhancing cooperativity of molecular J-aggregates by resonantly coupled dielectric metasurfaces
- Symmetry-protected bound states in the continuum on an integrated photonic platform
- Ultrashort pulse biphoton source in lithium niobate nanophotonics at 2 μm
- Entangled photon-pair generation in nonlinear thin-films
- Directionally tunable co- and counterpropagating photon pairs from a nonlinear metasurface
- All-optical modulator with photonic topological insulator made of metallic quantum wells
- Photo-thermo-optical modulation of Raman scattering from Mie-resonant silicon nanostructures
- Plasmonic electro-optic modulators on lead zirconate titanate platform
- Miniature spectrometer based on graded bandgap perovskite filter
- Far-field mapping and efficient beaming of second harmonic by a plasmonic metagrating