• Photonics Research
  • Vol. 13, Issue 6, 1554 (2025)
Fangyuan Peng1,†, Hongxiang Chen2,†, Lipeng Wan1, Xiao-Dong Chen2,3..., Jianwen Dong2, Weimin Deng1,* and Tianbao Yu1,4|Show fewer author(s)
Author Affiliations
  • 1School of Physics and Materials Science & Jiangxi Provincial Key Laboratory of Photodetectors, Nanchang University, Nanchang 330031, China
  • 2School of Physics & State Key Laboratory of Optoelectronic Materials and Technologies, Sun Yat-sen University, Guangzhou 510275, China
  • 3e-mail: chenxd67@mail.sysu.edu.cn
  • 4e-mail: yutianbao@ncu.edu.cn
  • show less
    DOI: 10.1364/PRJ.555997 Cite this Article Set citation alerts
    Fangyuan Peng, Hongxiang Chen, Lipeng Wan, Xiao-Dong Chen, Jianwen Dong, Weimin Deng, Tianbao Yu, "Dual-band dislocation modes in a topological photonic crystal," Photonics Res. 13, 1554 (2025) Copy Citation Text show less

    Abstract

    Introducing topological lattice defects, such as dislocations, into topological photonic crystals enables the emergence of many interesting phenomena, including robust bound states and fractional charges. Previous studies have primarily focused on the realization of dislocation modes within a single band gap, which limits the number of dislocation modes and their applications. Here, we design a topological photonic crystal with two topologically non-trivial band gaps. By introducing a dislocation defect into this system, we observe the emergence of localized dislocation modes in both band gaps. Furthermore, we demonstrate a two-channel add-drop filter by coupling two dislocation modes with topological edge modes. These findings are rigorously validated through full-wave numerical simulations and experimental pump-probe transmission measurements. Our results provide a foundation for further exploration of dislocation modes and unlock the potential for harnessing other topological defect modes in dual-band-gap systems.

    1. INTRODUCTION

    Since the proposal of the photonic quantum Hall effect in 2008, topological photonics has attracted increasing attention [17]. Photonic crystal (PC) is one of the systems used to realize photonic topological states. By designing photonic bands with non-zero topological invariants, unique optical effects and transport properties in PCs can be achieved, such as topological edge states without backscattering [3,4,812]. In topological PCs, the transport of topological edge states is protected by the band topology, thereby ensuring that light propagation remains imperturbable in the face of large-angle bending and defects in the edge. This unique property enables the realization of waveguides that are both low-loss and capable of withstanding sharp bends [1113].

    On the other hand, optical microcavities [14,15], which enable optical localization, are also one of the important photonic devices. Traditionally, optical microcavity modes in PCs are achieved by introducing point defects [15] in the PCs. However, these modes lack topological protection, making them susceptible to fabrication errors that can significantly alter the resonance frequency [16]. This sensitivity hinders the realization of optical microcavities whose resonance frequencies are less sensitive to fabrication errors. In recent years, researchers have explored the design of optical microcavities using second-order topological corner states [1721] or topological defects with non-trivial real-space topology [2230]. For example, robust topological defect modes can be achieved by the introduction of a dislocation defect [26,3135] in a single topological PC. A dislocation is characterized by the Burgers vector B. Due to the conservation of the Burgers vector, the dislocations cannot be eliminated through small local perturbations and are protected by topology [31]. In 2018, researchers observed dislocation modes in magnetic PCs in the microwave regime [26]. Due to the weak response of magnetic materials in the optical regime, this method is challenging to implement in the optical regime. Recently, dislocation modes based on two-dimensional dielectric PCs have been experimentally observed in the microwave [32] and terahertz regimes [33]. Besides, fractional topological numbers and bulk-dislocation correspondence are also observed in photonic crystals [35]. However, previous studies mostly focused on realizing dislocation modes within a single band gap, which typically supports only one dislocation mode. This limits dislocation mode applications that require two resonant modes, such as a dual-band filter and a dual-mode laser [36]. Therefore, there is an urgent need for a scheme to achieve dislocation modes in dual band gaps.

    In this work, a topological PC with two topologically non-trivial band gaps is designed. By introducing a dislocation with non-trivial real-space topology into the topological PC, zero-dimensional dislocation modes emerge in both band gaps. These dislocation modes exhibit a notable degree of robustness, as their frequencies do not change significantly when the radius or position of the dielectric rod near the dislocation is modified. Furthermore, by integrating topological waveguides with two dislocations, a two-channel add-drop filter is realized. The theoretical results presented here are corroborated by experimental results in the microwave regime. Our results may promote the applications of dislocation modes in multimode microcavities and on-chip add-drop filters.

    2. THEORETICAL DESIGN AND EXPERIMENTAL OBSERVATION OF DUAL-BAND DISLOCATION MODES

    To achieve dual dislocation modes, we initially design a PC with two topologically non-trivial band gaps. Here, we consider the PC consisting of two dielectric semicylinders with a refractive index of 2.9, named as PC1, as depicted in Fig. 1(a). The lattice constant a is 12.5 mm and the radius of the semicylinder r is 3.75 mm. The gap between two semicylinders is denoted as 2d. Figure 1(b) illustrates the bulk bands of the PC with d=2.5  mm. There are two photonic band gaps, as highlighted in yellow. The topological properties of each band can be determined by calculating the 2D Zak phase [3741]: Zj=dkxdkyTr[A^j],where j=x or y, A^j=iu(k)|kj|u(k), and |u(k) denotes the periodic part of the Bloch function. For PCs exhibiting mirror symmetry along the j direction, Zj is quantized to 0 and π. In addition, the Zak phase of each band can be derived from the mirror eigenvalues of the eigenmodes at high-symmetry points [40]. Figure 1(c) shows the Ez fields of bulk modes of the first, second, and third bands at the Γ,X, and Y points. According to the eigenmodes’ mirror symmetries with respect to the boundary of the unit cell, Zx of each bulk band is determined, as labeled on each band. Figures 1(d) and 1(e) depict the evolution of the modes at the Γ and X points for varying values of d. As d increases from 0 to 2.5 mm, mode inversion occurs between band1 and band2 as well as band3 and band4 at the X point. In contrast, mode inversion does not appear between the adjacent bands at the Γ point. This phenomenon can be understood by mapping the 2D PC to a Su-Schrieffer-Heeger model [38,42]. The distance between two dielectric semicylinders modulates the hopping amplitude. Since we alter the distance solely along the x direction, mode inversion occurs exclusively at the X point. Mode inversion usually results in the change of a band’s Zak phase. Figure 1(f) illustrates the evolution of Zx for each band as d increases from 0 to 2.5 mm, showing that Zx of each band changes by π after the mode inversion.

    (a) Unit cell of the photonic crystal (PC). Lattice constant of the PC is a=12.5 mm. The radius and refractive index of the dielectric semicylinder and background are r=3.75 mm, n1=2.9, and n0=1.03, respectively. The lower panel shows the Brillouin zone. (b) Bulk bands of the PC1 with d=2.5 mm, where yellow rectangles highlight the band gaps. (c) Eigen-field profiles of bulk modes at high-symmetry points in the first, second, and third bands. (d), (e) Evolution of the modes at the (d) Γ and (e) X points as a function of d. (f) Evolution of Zx for each band when d increases from 0 to 2.5 mm.

    Figure 1.(a) Unit cell of the photonic crystal (PC). Lattice constant of the PC is a=12.5  mm. The radius and refractive index of the dielectric semicylinder and background are r=3.75  mm, n1=2.9, and n0=1.03, respectively. The lower panel shows the Brillouin zone. (b) Bulk bands of the PC1 with d=2.5  mm, where yellow rectangles highlight the band gaps. (c) Eigen-field profiles of bulk modes at high-symmetry points in the first, second, and third bands. (d), (e) Evolution of the modes at the (d) Γ and (e) X points as a function of d. (f) Evolution of Zx for each band when d increases from 0 to 2.5 mm.

    Based on the results in Fig. 1(a), we select a PC with an appropriate Zak phase to design a dislocation defect, characterized by the Burgers vector B. The number of dislocation mode is determined by the topological invariant Q, defined as [26,33,34] Q=1πθ·B:mod2,where θ=(θΓYa,θΓXa). θΓY and θΓX are the Zak phases of the bulk bands along the ΓY and ΓX directions with contributions from all bands below the gap. Dislocation modes emerge exclusively when the topological invariant Q equals one. According to Eq. (2), a dislocation mode is achievable when θ is parallel to the Burgers vector B. For PC1, θ=(0,πa) for both gaps. To realize dislocation modes in both gaps, a dislocation defect with a Burgers vector B=(0,a) should be introduced. As shown in Fig. 2(a), a dislocation defect is created by removing a row of unit cells and shifting the adjacent lower and upper unit cells towards the center. The red arrows form a closed loop around the dislocation and the orange arrow indicates the Burgers vector B. Note that some adjacent dielectric semicylinders are staggered. For experimental convenience, we slightly tune the positions of these adjacent dielectric semicylinders to make them form a circular dielectric cylinder, as shown in Fig. 2(b). Figure 2(c) presents the eigenfrequencies of the modes near the first bulk gap, with the yellow region highlighting the bulk band gap. The red circle indicates the dislocation mode at the frequency of 6.741 GHz and the corresponding |Ez| field is strongly localized around the dislocation defect [Fig. 2(d)]. The eigenfrequencies of the modes near the second bulk gap are displayed in Fig. 2(f), also featuring a dislocation mode, highlighted by the red circle. Figure 2(g) shows the localized electric field distribution of this dislocation mode. Besides, two trivial cavity modes are present in the gap, indicated by the blue dots in Fig. 2(f). A discussion about the identification of these trivial cavity modes is provided in Appendix B.

    (a) Schematic of the dislocation and (b) modified dislocation constructed by the PC1. (c) Eigenfrequencies of the modes near the first bulk gap, where the red circle indicates the dislocation mode. The yellow region highlights the bulk gap. (d) Eigen-field profile of dislocation modes in (c). (e) Experimentally measured spectrum when the excitation source is placed around the dislocation. (f)–(h) Similar to (c)–(e), but for the second bulk gap. The blue dots indicate the trivial modes in the second band gap.

    Figure 2.(a) Schematic of the dislocation and (b) modified dislocation constructed by the PC1. (c) Eigenfrequencies of the modes near the first bulk gap, where the red circle indicates the dislocation mode. The yellow region highlights the bulk gap. (d) Eigen-field profile of dislocation modes in (c). (e) Experimentally measured spectrum when the excitation source is placed around the dislocation. (f)–(h) Similar to (c)–(e), but for the second bulk gap. The blue dots indicate the trivial modes in the second band gap.

    To further confirm the localized dislocation modes, we fabricate a microwave photonic crystal whose structural parameters are the same as those descripted in Fig. 1. In the experiment, the dielectric cylinders are made of alumina and the background is filled with foam whose refractive index is 1.03. Both the alumina cylinders and foam background have a thickness of 10 mm, and are sandwiched between upper and lower metallic plates. By employing the zero-order transverse magnetic waveguide modes, which are uniform along the z direction, this PC waveguide is designed to replicate the band dispersion of TM modes of a 2D PC. A source antenna is inserted through the bottom plate to excite the eigenmode, while a probe antenna is introduced from the top plate to capture the resulting electromagnetic fields. This setup enables efficient excitation and measurement of the localized modes within the photonic crystal structure (see details in Appendix D). Figure 2(e) shows the measured transmission spectrum in the first band gap by putting the source and probe antenna around the dislocation. A resonance peak near 6.777 GHz confirms the existence of the dislocation mode. In the same way, the transmission spectrum in the second band gap is also measured, as shown in Fig. 2(h). Three resonance peaks are observed. The peak observed near 12.5 GHz corresponds to the excitation of the dislocation mode within the second band gap. The other two resonance peaks are attributed to the excitation of trivial cavity modes, which lack topological protection.

    Next, we investigate the robustness of these dislocation modes. In the first case, we introduce a point defect by increasing the radius of a dielectric cylinder to 5.25 mm, as highlighted by the yellow circle in Fig. 3(a). Figures 3(b) and 3(c) show the electric field of the dislocation modes at the frequency of 6.701 GHz and 12.724 GHz, respectively. For another case, we shift the positions of two dielectric cylinders with a displacement vector of (a/5, a/5) and (a/5,a/5), as highlighted by the yellow circles in Fig. 3(d). Now, the frequencies of these two dislocation modes are 6.714 GHz [Fig. 3(e)] and 12.477 GHz [Fig. 3(f)], respectively. The dislocation modes persist in both cases and their frequencies do not change significantly, indicating the robustness of the dislocation modes.

    (a) Schematic of the dislocation with a point defect. The yellow circle highlights the point defect with a radius of 5.25 mm. (b), (c) Electric field distribution of two dislocation modes for the structure in (a). (d) Schematic of the dislocation with two shifted dielectric cylinders. The yellow circles highlight the shifted dielectric cylinders with a displacement vector of (−a/5, a/5) and (−a/5, −a/5), respectively. (e), (f) Electric field distribution of two dislocation modes for the structure in (d).

    Figure 3.(a) Schematic of the dislocation with a point defect. The yellow circle highlights the point defect with a radius of 5.25 mm. (b), (c) Electric field distribution of two dislocation modes for the structure in (a). (d) Schematic of the dislocation with two shifted dielectric cylinders. The yellow circles highlight the shifted dielectric cylinders with a displacement vector of (a/5, a/5) and (a/5, a/5), respectively. (e), (f) Electric field distribution of two dislocation modes for the structure in (d).

    3. REALIZATION OF A TWO-CHANNEL ADD-DROP FILTER

    Based on two dislocation modes, a two-channel add-drop filter is designed. The bus waveguide and drop waveguide are constructed by the edges between PC1 and another topologically trivial PC2. The unit cell of PC2 can be obtained by shifting PC1’s unit cell by 0.5a along the x direction. The bulk bands of these two PCs are identical, but each bulk band’s Zx differs by π [43]. The bulk bands and the eigenmodes at high-symmetry points of PC2 are provided in Appendix A. Due to their difference in Zx, topological edge states emerge at the y-direction edge formed by PC1 and PC2 [39]. Figure 4(a) shows the projected band, with the inset illustrating the edge schematic. Topological edge states emerge in both band gaps, with the red (blue) line representing the edge states in the first (second) band gap. The lower panel shows the electric fields of two edge states at ky=0, indicating strong confinement of electric fields around the interface. By utilizing two dislocation defects and the edge between PC1 and PC2, a two-channel add-drop filter is designed [Fig. 4(b)]. Here, compared with dislocation1, dislocation2’s center is extended by one more unit cell; see more details in Appendix C. Consequently, the modes supported by dislocation1 and dislocation2 have different frequencies. When the incident light at port A matches the dislocation mode’s frequency, the excited edge state is directed to either port C or port D through coupling to the dislocation mode. The coupling between two edge states on the right of two dislocations is blocked by PC2. To minimize the coupling between the waveguide at port A and the one at port C, the latter is strategically designed as a stepwise waveguide. Figure 4(e) shows the simulated transmission spectra at ports C and D. At the frequency of 6.708 GHz, the majority of the light is directed to port C, with the corresponding simulated electric field distribution illustrated in Fig. 4(d). In the frequency range of 13.265 GHz to 13.275 GHz, the majority of the light is routed to port D at a frequency of 13.27 GHz [Fig. 4(h)], with the corresponding simulated electric field distribution depicted in Fig. 4(g). The function of the filter is also observed in experiment. Figure 4(c) presents a photograph of the sample, with the structural parameters identical to those depicted in Fig. 4(b). Unlike the setup for measuring localized dislocation modes, here the source antenna is positioned near port A, while the probe antenna is placed at either port C or D to measure the transmitted spectra. The measured results at ports C and D within the first band gap are presented in Fig. 4(f), with the curve’s trend closely resembling the simulation results shown in Fig. 4(e). For the second band gap, a resonant peak is also observed in the measured result at port D [Fig. 4(i)]. Note that the linewidth of the resonant peak in Figs. 4(f) and 4(i) is broader than that of the simulation results. This discrepancy may arise from the fact that the top metallic plate does not tightly contact with the dielectric rods, leading to microwave leakage through the air gap.

    (a) Calculated edge states, which are highlighted in red and blue. The inset shows the schematic of the edge. Lower panel shows the |Ez| field of two edge states at ky=0. (b) Schematic of the add-drop filter, where the input port is denoted as A, and the output ports are denoted as C and D. (c) Photograph of the sample. (d) Simulated |Ez| field distribution at the frequency of 6.708 GHz. (e) Simulated transmission spectrum of the output ports C and D. (f) Experimentally measured electric field intensity at ports C and D. (g)–(i) Similar to (d)–(f) but for the frequency range in the second bulk gap.

    Figure 4.(a) Calculated edge states, which are highlighted in red and blue. The inset shows the schematic of the edge. Lower panel shows the |Ez| field of two edge states at ky=0. (b) Schematic of the add-drop filter, where the input port is denoted as A, and the output ports are denoted as C and D. (c) Photograph of the sample. (d) Simulated |Ez| field distribution at the frequency of 6.708 GHz. (e) Simulated transmission spectrum of the output ports C and D. (f) Experimentally measured electric field intensity at ports C and D. (g)–(i) Similar to (d)–(f) but for the frequency range in the second bulk gap.

    4. CONCLUSION

    In conclusion, we theoretically and experimentally observed the emergence of dual dislocation modes by introducing a dislocation into the PC with dual topologically non-trivial band gaps. These dislocation modes exhibit a significant degree of robustness against variations in the radius or position of the dielectric rods in their vicinity. Furthermore, we have successfully realized an add-drop filter by integrating topological waveguides with two dislocations, demonstrating the functionality of the designed device with experimental results. Our results may promote the applications of dislocation modes, such as on-chip add-drop filters. Besides, the proposed topological PC may serve as a platform for the realization of other topological defect modes within dual band gaps, such as vortex nanolasers based on dual-frequency disclination mode [24]. Note that our microwave experimental scheme is not suitable for the near-infrared band. In the future, one potential direction is studying how to realize dual-band dislocation modes in a dielectric photonic crystal slab and explore potential applications of these dislocation modes, such as on-chip dual-band filters and dual-mode lasers.

    APPENDIX A: BULK BANDS AND THEIR ZAK PHASES OF PC2

    The unit cell of PC2 is shown in the inset of Fig. 5(a). The lattice constant is a=12.5  mm. The radius and refractive index of the rod are r=3.75  mm and n1=2.9, respectively. The refractive index of the background medium is 1.03. Figure 5(a) shows the calculated bulk band, where yellow rectangles highlight two band gaps. Figure 5(b) shows the Ez field distributions at the high-symmetry points. According to the symmetry of the eigenmodes, Zak phases of each band can be determined, as labeled on each band in Fig. 5(a).

    (a) Bulk band structure of PC2. The inset shows the structure of the unit cell. The lattice constant is a=12.5 mm. The radius and refractive index of the rod are r=3.75 mm and n1=2.9, respectively. (b) Eigen-field profiles at high-symmetry points in the first, second, and third bands.

    Figure 5.(a) Bulk band structure of PC2. The inset shows the structure of the unit cell. The lattice constant is a=12.5  mm. The radius and refractive index of the rod are r=3.75  mm and n1=2.9, respectively. (b) Eigen-field profiles at high-symmetry points in the first, second, and third bands.

    APPENDIX B: IDENTIFICATION OF THE TRIVIAL CAVITY MODES

    To identify the trivial cavity modes in Fig. 2(e), we construct a dislocation by PC2, as shown in Fig. 6(a). The calculated eigenmodes and field distributions are shown in Figs. 6(b) and 6(c). According to Eq. (2) in the main text, when θΓX=0, the topological invariant Q is zero, indicating there is no topological dislocation mode in the band gap. Therefore, two modes in Fig. 6(c) are topologically trivial cavity modes. For comparison, two extra modes in the dislocation formed by PC1 are plotted in Fig. 6(f). One can see that the mode profiles in Fig. 6(f) are similar to that in Fig. 6(c). Thus, the extra two modes in Fig. 6(f) are trivial cavity modes.

    (a) Dislocation constructed by PC2. (b) Eigenfrequencies of the modes near the second bulk gap, where the red dots indicate the trivial cavity modes in the gap. The yellow region highlights the bulk gap. (c) Eigen-field profile of two trivial cavity modes in (b). (d)–(f) Similar to (a)–(c), but for the dislocation constructed by PC1.

    Figure 6.(a) Dislocation constructed by PC2. (b) Eigenfrequencies of the modes near the second bulk gap, where the red dots indicate the trivial cavity modes in the gap. The yellow region highlights the bulk gap. (c) Eigen-field profile of two trivial cavity modes in (b). (d)–(f) Similar to (a)–(c), but for the dislocation constructed by PC1.

    APPENDIX C: MODIFIED DISLOCATION IN THE ADD-DROP FILTER

    To increase the coupling efficiency between the edge state and the dislocation mode in the second band gap, we extended the lower dislocation’s center by one more unit cell, as highlighted in red in Fig. 7(a). Figure 7(b) shows the frequencies of the modes in the modified dislocation. The yellow region highlights the bulk gap. The dislocation modes are highlighted by red dots and their mode profiles are shown in Figs. 7(c) and 7(d).

    (a) Dislocation with the center extended by one more unit cell, as highlighted in red. (b) Eigenfrequencies of the modes in two band gaps. (c), (d) Eigen-field profiles of two dislocation modes.

    Figure 7.(a) Dislocation with the center extended by one more unit cell, as highlighted in red. (b) Eigenfrequencies of the modes in two band gaps. (c), (d) Eigen-field profiles of two dislocation modes.

    APPENDIX D: EXPERIMENTAL SETUP

    Figure 8 shows the photograph of the experimental setup. The sample is sandwiched between two metallic plates. A source antenna is inserted through the bottom plate to excite the eigenmode, while a probe antenna is introduced from the top plate to capture the resulting electromagnetic fields.

    Photograph of the experimental setup.

    Figure 8.Photograph of the experimental setup.

    References

    [1] F. D. M. Haldane, S. Raghu. Analogs of quantum-Hall-effect edge states in photonic crystals. Phys. Rev., 78, 033834(2008).

    [2] S. Raghu, F. D. M. Haldane. Possible realization of directional optical waveguides in photonic crystals with broken time-reversal symmetry. Phys. Rev. Lett., 100, 013904(2008).

    [3] Z. Wang, Y. Chong, J. D. Joannopoulos. Observation of unidirectional backscattering-immune topological electromagnetic states. Nature, 461, 772-775(2009).

    [4] L. Lu, J. D. Joannopoulos, M. Soljačić. Topological photonics. Nat. Photonics, 8, 821-829(2014).

    [5] S. Zhang, X. Qi. Topological insulators and superconductors. Rev. Mod. Phys., 83, 1057-1110(2011).

    [6] T. Ozawa, H. M. Price, A. Amo. Topological photonics. Rev. Mod. Phys., 91, 015006(2019).

    [7] G. Tang, X. He, F. Shi. Topological photonic crystals: physics, designs, and applications. Laser Photonics Rev., 16, 2100300(2022).

    [8] Z. Wang, Y. D. Chong, J. D. Joannopoulos. Reflection-free one-way edge modes in a gyromagnetic photonic crystal. Phys. Rev. Lett., 100, 013905(2008).

    [9] X. Ao, Z. Lin, C. T. Chan. One-way edge mode in a magneto-optical honeycomb photonic crystal. Phys. Rev., 80, 033105(2009).

    [10] S. A. Skirlo, L. Lu, Y. Igarashi. Experimental observation of large Chern numbers in photonic crystals. Phys. Rev. Lett., 115, 253901(2015).

    [11] M. Wang, R. Zhang, L. Zhang. Topological one-way large-area waveguide states in magnetic photonic crystals. Phys. Rev. Lett., 126, 067401(2021).

    [12] Q. Chen, L. Zhang, F. Chen. Photonic topological valley-locked waveguides. ACS Photon., 8, 1400-1406(2021).

    [13] B. Yan, Y. Peng, J. Xie. Multifrequency and multimode topological waveguides in a Stampfli-triangle photonic crystal with large valley Chern numbers. Laser Photon. Rev., 18, 2470035(2024).

    [14] T. Asano, S. Noda. Photonic crystal devices in silicon photonics. Proc. IEEE., 106, 2183-2195(2018).

    [15] T. Yoshie, J. Vučković, A. Scherer. High quality two-dimensional photonic crystal slab cavities. Appl. Phys. Lett., 79, 4289-4291(2001).

    [16] J. Noh, W. A. Benalcazar, S. Huang. Topological protection of photonic mid-gap defect modes. Nat. Photonics, 12, 408-415(2018).

    [17] X. He, M. Li, H. Qiu. In-plane excitation of a topological nanophotonic corner state at telecom wavelengths in a cross-coupled cavity. Photonics Res., 9, 1423-1431(2021).

    [18] W. Zhang, X. Xie, H. Hao. Low-threshold topological nanolasers based on the second-order corner state. Light: Sci. Appl., 9, 109(2020).

    [19] C. Han, M. Kang, H. Jeon. Lasing at multidimensional topological states in a two-dimensional photonic crystal structure. ACS Photonics, 7, 2027-2036(2020).

    [20] L. Tao, Y. Liu, L. Du. Evolution of the edge states and corner states in a multilayer honeycomb valley-Hall topological metamaterial. Phys. Rev. B, 107, 035431(2023).

    [21] B. Xie, G. Su, H. Wang. Visualization of higher-order topological insulating phases in two-dimensional dielectric photonic crystals. Phys. Rev. Lett., 122, 233903(2019).

    [22] Z. Lin, Q. Wang, Y. Liu. Topological phenomena at defects in acoustic, photonic and solid-state lattices. Nat. Rev. Phys., 5, 483-495(2023).

    [23] S. Tang, Y. Xu, F. Ding. Continuously tunable topological defects and topological edge states in dielectric photonic crystals. Phys. Rev. B., 107, L041403(2023).

    [24] M. Hwang, H. Kim, J. Kim. Vortex nanolaser based on a photonic disclination cavity. Nat. Photonics, 18, 286-293(2024).

    [25] Z. Zhang, S. Yang, S. Yan. Homogeneous dislocation-induced rainbow concentrating for elastic waves. Phys. Rev. Appl., 18, 044015(2022).

    [26] F. F. Li, H. X. Wang, Z. Xiong. Topological light-trapping on a dislocation. Nat. Commun., 9, 2462(2018).

    [27] Y. Liu, S. Leung, F. Li. Bulk–disclination correspondence in topological crystalline insulators. Nature, 589, 381-385(2021).

    [28] Z. Lin, Y. Wu, B. Jiang. Topological Wannier cycles induced by sub-unit-cell artificial gauge flux in a sonic crystal. Nat. Mater., 21, 430-437(2022).

    [29] V. Juričić, B. Roy. Dislocation as a bulk probe of higher-order topological insulators. Phys. Rev. Res., 3, 033107(2021).

    [30] Z. Hu, D. Bongiovanni, Z. Wang. Topological orbital angular momentum extraction and twofold protection of vortex transport. Nat. Photonics, 19, 162-169(2025).

    [31] L. Ye, C. Qiu, M. Xiao. Topological dislocation modes in three-dimensional acoustic topological insulators. Nat. Commun., 13, 508(2022).

    [32] X. Chen, F. Shi, J. Liu. Second Chern crystals with inherently non-trivial topology. Natl. Sci. Rev., 10, nwac289(2023).

    [33] J. Lu, K. G. Wirth, W. Gao. Observing 0D subwavelength-localized modes at ∼100 THz protected by weak topology. Sci. Adv., 7, eabl3903(2021).

    [34] Y. Ran, Y. Zhang, A. Vishwanath. One-dimensional topologically protected modes in topological insulators with lattice dislocations. Nat. Phys., 5, 298-303(2009).

    [35] X. Zhang, C. Liang, F. Li. Observation of bulk-dislocation correspondence in photonic crystals. Phys. Rev. Appl., 22, 064097(2024).

    [36] H. Z. Weng, Y. D. Yang, J. L. Wu. Dual-mode microcavity semiconductor lasers. IEEE J. Sel. Top. Quantum Electron., 25(2019).

    [37] J. Zak. Berry’s phase for energy bands in solids. Phys. Rev. Lett., 62, 2747-2750(1989).

    [38] F. Liu, K. Wakabayashi. Novel topological phase with a zero Berry curvature. Phys. Rev. Lett., 118, 076803(2017).

    [39] Y. He, Y. Gao, H. Lin. Topological edge and corner states based on the transformation and combination of photonic crystals with square lattice. Opt. Commun., 512, 128038(2022).

    [40] M. Xiao, Z. Q. Zhang, C. T. Chan. Surface impedance and bulk band geometric phases in one-dimensional systems. Phys. Rev. X., 4, 021017(2014).

    [41] M. Xiao, G. Ma, Z. Yang. Geometric phase and band inversion in periodic acoustic systems. Nat. Phys., 11, 240-244(2015).

    [42] X. Chen, W. Deng, F. Shi. Direct observation of corner states in second-order topological photonic crystal slabs. Phys. Rev. Lett., 122, 233902(2019).

    [43] Y. Yang, T. Xu, Y. F. Xu. Zak phase induced multiband waveguide by two-dimensional photonic crystals. Opt. Lett., 42, 3085-3088(2017).

    Fangyuan Peng, Hongxiang Chen, Lipeng Wan, Xiao-Dong Chen, Jianwen Dong, Weimin Deng, Tianbao Yu, "Dual-band dislocation modes in a topological photonic crystal," Photonics Res. 13, 1554 (2025)
    Download Citation