Symmetry plays a fundamental role in topological photonic crystals, and topological phase transitions induced by disorder have also been extensively explored in recent years. However, in this work, we find anisotropy can be induced by reducing symmetry in a symmetry triangular photonic crystal. We investigate that anisotropy-induced interfaces profoundly affect edge states and enable the realization of slow light dispersion. Numerical simulations reveal a transition from gapless chiral edge modes to gapped flat band dispersion. Furthermore, we observe higher-order corner states in corner structures constructed by anisotropic interfaces. The corner states can be induced and localized at different lattice positions, thereby realizing multiple types of higher-order topological states. We demonstrate the significance of anisotropic geometry in topological photonics. These findings open new possibilities for steering wave transport in multiple dimensions and offer, to our knowledge, a novel research perspective on the transformation of topological states induced by anisotropic lattices.
【AIGC One Sentence Reading】:In C2v triangular photonic crystals, reducing symmetry induces anisotropy. This affects edge states, enabling slow light dispersion. Numerical simulations show a transition to flat band dispersion, and higher-order corner states emerge. Anisotropic geometry is key for topological photonics, opening new wave transport possibilities.
【AIGC Short Abstract】:In C2v symmetry triangular photonic crystals, reducing symmetry induces anisotropy. This anisotropy impacts edge states, enabling slow light dispersion. Numerical simulations show a shift from gapless chiral edge modes to gapped flat band dispersion. Higher-order corner states emerge in corner structures with anisotropic interfaces, localized at different lattice positions. The study highlights the significance of anisotropic geometry in topological photonics, offering new ways for wave transport control.
Note: This section is automatically generated by AI . The website and platform operators shall not be liable for any commercial or legal consequences arising from your use of AI generated content on this website. Please be aware of this.
1. INTRODUCTION
Topological insulators represent a unique quantum state in condensed matter systems, revealing nontrivial phases [1–6]. Their electronic properties, such as Dirac fermions [7–9] and valley polarization [10–12], are determined by atomic constituents and bonding arrangements. Photonic crystals (PCs), analogous to conventional crystals, exhibit band gaps due to Bragg diffraction [13–15]. In these systems, periodicity and symmetry play a pivotal role, making topological photonic crystals (TPCs) a burgeoning field in the realm of topological states, ensured by inversion () and time-reversal () symmetries [16–24]. When time-reversal symmetry is conserved, classical analogues of the quantum spin-Hall effect [25–27] and the valley-Hall effect [23,24,28,29] rely on broken inversion symmetry, with edge states protected by pseudospin conservation. Gapless edge states at interfaces of different topological phases are stable and non-local [25–30]. In addition, besides the conventional descriptions of bulk-edge correspondence, higher-order topological insulators have been introduced [31–38]. These topological states are relatively insensitive to external disturbances, offering a stable and reliable avenue for the manipulation of information processing.
A topological photonics insulator based on graphene-like structures has Dirac cones typically in hexagonal lattices of or symmetry. Topological phenomena in lower-symmetry systems are rare. Recently, symmetry breaking in electronics has induced novel nonlinear quantum phenomena by modifying the energy band structure through the introduction of anisotropy [39–42]. Chen et al. have achieved terahertz emission in films through symmetry breaking, uncovering a universal pathway for efficient nonlinear transport [40]. Duan et al. have disrupted the rotational symmetry of materials, clearly presenting a novel academic concept for realizing Berry curvature dipole moments in energy bands by exploiting lattice symmetry breaking at heterogeneous interfaces [41]. Yuan et al. have demonstrated that FeSe films exhibit anisotropy at low temperatures, triggering nonlinear phenomena [42]. In addition, disorder can induce or alter topological states, leading to unique phase transition phenomena, which are also crucial for understanding and designing materials with desired topological properties [43–47].
As an additional degree of freedom, anisotropy has profound implications. Achieving in-plane anisotropy in PCs typically involves breaking and reducing rotational symmetry, a feature rarely reported in low-symmetry TPCs. In this work, we propose a triangular lattice composed of elliptical metal scatterers. By rotating these scatterers, we induce strong in-plane anisotropy in a symmetric triangular PC, allowing flexible control over the band structure. We investigate the interplay between topological edge dispersion and anisotropy at the interface. Notably, we theoretically examine anisotropy-induced domain walls and observe a significant flat band in the edge state dispersion as the interface’s geometric orientation changes. Furthermore, we find novel multi-type localized states, which are associated with higher-order topological phenomena in corners formed by anisotropic domain walls. The result challenges conventional views on space inversion symmetrical interfaces, which are compatible with the supercells for flattened edge dispersion and higher-order corner states. Our findings highlight that anisotropy has a critical role in topological effects stemming from the reduction of structural symmetry. It can be expected that our findings will provide new insights into flat band dispersion and higher-order topology.
Sign up for Photonics Research TOC Get the latest issue of Advanced Photonics delivered right to you!Sign up now
2. RESULTS AND DISCUSSION
To induce in-plane anisotropy in the crystal structure, we start with a symmetric elliptical metal scatterer while keeping the unit cell hexagonal, as depicted in the right panel of Fig. 1(a). The proposed PC consists of a parallel-plate metal waveguide filled with a periodically arranged triangular array of elliptical metal copper pillars connected to the top and bottom metal plates, with a height of . Each unit cell of the PC contains a pillar with an ellipticity , minor axis lengths , and the lattice constant . In contrast, as the triangular lattice of uniform circular pillars shown in the left panel of Fig. 1(a), the air channels surrounding them can form a graphene-like honeycomb lattice; therefore, such valley degree of freedom is existent [48]. However, as the pillars are transformed into asymmetric triangular sublattices such as ellipses, strong symmetric anisotropic scatterers are induced.
Figure 1.Proposed symmetry triangular PC with anisotropy. (a) Schematic illustration of the metal pillars. (b) Periodically arranged triangular array of the elliptical metal pillars (, , and are the basis vectors of three mirror symmetrical directions). The right panel presents the complex first Brillouin zone. (c) Dispersion relation of the and paths.
Figure 1(b) illustrates the crystal structure of the circular pillars elongated along the horizontal direction, with the symmetry reduced from to or from to . The corresponding complex first Brillouin zone is presented in the right panel in Fig. 1(b). In fact, the choice of elliptical metal pillars as a unit cell is a visualization of the anisotropic primitive cell. Our design strategy for anisotropy is exemplified by the artificial modification of band dispersion in different directions. By solving Maxwell equations, we explicitly reveal the harmonic transverse magnetic modes (TM modes with nonvanishing field components , , and ) hosted by the symmetry triangular PC. By retaining the lowest-order transverse modes, we can assume the field decomposition as follows [49]: where , are the Bloch wave numbers inside the Brillouin zone, are the normalized periodic field profiles, and index refers to lower (upper) propagation bands.
By setting up boundaries and electromagnetic wave modes corresponding to Eq. (1), we use the commercial software ANSYS HFSS to calculate the band dispersion diagram of the proposed crystal. Figure 1(c) presents the numerical results of photonic bands along the specific direction of the lattice. Significantly, we observe an inconsistency in the dispersion relation corresponding to the and paths in Fig. 1(c), which is certainly absent in the case of conventional photonic Dirac systems. Hence, as the symmetry of the triangular photonic lattice is reduced from to or from to in the triangular photonic lattice, the structure exhibits strong in-plane anisotropy. The ellipticity factor characterizes the shape deviation from a perfect circle to an ellipse, which determines the strength of the crystal structure anisotropy (see Appendix A for supporting content). Furthermore, the anisotropic phase diagrams of elliptical PCs are presented in Appendix B for supporting content.
We examine the spectral characteristics of a two-dimensional (2D) PC arranged in a triangular lattice, influenced by anisotropy. Conversely, we can implement the different arrangements of the PC lattice with different crystal basis vector directions, which also introduces anisotropy. The upper panel of Fig. 2(a) represents the arrangement of the PC relative to various crystal orientation vectors as . The orange, blue, and dark gray ellipses represent the lattices aligned with the , , and basis vectors, respectively. The calculated band structure of the unit cell is presented in the lower panel of Fig. 2(a), where a pair of Dirac points is generated in a specific direction. The emergence of the Dirac points at and is not deterministic, and the position of these degenerate points is influenced by the spatial filling rate of the crystal structure, which in turn depends on the ellipticity in our proposed structure. A detailed evolution is provided in Appendix C to support content.
Figure 2.Arrangement of the PC structures and the band structures of the different configurations with (a) , (b) , and (c) . Upper panels represent the arrangement of the PC with various crystal orientation vectors , , and induced by anisotropy. The orange, blue, and dark gray ellipses represent the , , and basis vectors, respectively. Lower panels show the band diagram of the different unit cells. Orange shaded regions represent the band gap.
A Dirac mass is introduced to characterize the distinct bulk topology gap in our 2D triangular lattice [50]. As , a topological gap emerges, as depicted in the lower panels of Figs. 2(b) and 2(c). There are two opposite chiral vortex phases: the pseudospin topological insulator with and . We can understand the doubly degenerate dispersions around the Dirac point utilizing the theory: [36]. The effective Hamiltonian for a gapped Dirac point is written as , where is the unitary transformation matrix, is the coupling matrix, and and refer to the pseudospin-up and pseudospin-down matrices, respectively [50]. We have the expressions of where is the Dirac degeneracy frequency, is the value of the band gap, and and are determined by the first-order and second-order perturbation terms, respectively. The parity inversion at the point leads to the formation of the two opposite chiral vortex phases. The sign of the Dirac mass is controlled by the band flipping of the first and second bands, and this flipping can be regulated by the rotation angle of the ellipse scattering. Topological edge states can emerge at the interface between two PCs with opposite signs of the Dirac mass. When anisotropy is not considered, the detailed process of the dispersion relation of the edge states is proved in Appendix D for supporting content. Similarly, the crystal structure induced by crystal anisotropy can exhibit markedly different arrangement shapes based on different basis vector considerations, as shown by the permutation of orange, blue, and dark gray shapes in the upper panels of Figs. 2(b) and 2(c). Although the presence of anisotropy does not affect bulk-boundary correspondence, it provides new possibilities for the arrangement of the structure domain wall interfaces.
To investigate the topological edge states changing with the transformation of the anisotropy interface, we construct three types of zigzag domain walls along lattice directions corresponding to the , , and basis vectors as shown in the upper panels of Figs. 3(a)–3(c), named as , , and interfaces, respectively. We numerically simulate the corresponding dispersion of these edge states, as depicted in the lower panels in Figs. 3(d)–3(e). At the zigzag interface along the basis vector, the edge states display a pair of almost linear lines, and near-gapless dispersion curves are observed in Fig. 3(d). Due to the chiral vortex locking, the topologically projected edge modes exhibit chirality, implying that the two counter-propagating modes can be selectively excited by chiral sources with different chiralities. The underlying physics of this phenomenon mimics that of quantum valley-Hall insulators in photonic systems.
Figure 3.Anisotropy-induced zigzag domain walls. (a)–(c) Simulated electric field component at 14.47 GHz corresponding to , , and interface, respectively. (d)–(e) Dispersion diagrams of the anisotropy-induced zigzag edge states along the and basis vectors. (f) Left panel shows the simulated transmission of the zigzag domain walls, and right panel shows the density of states (DOS) of the band dispersion. (g) Group velocity of the edge states and . Black dotted line represents the zero-horizontal line of zero value. (h) Enlarged view of the flat dispersion curve of (e). (i) Intrinsic field intensity distribution at the frequencies of 14.462 GHz, 14.482 GHz, and 15.502 GHz. Electric field strength distribution of edge mode is along the direction.
However, a mode gap emerges between the upper and lower bulk modes for the zigzag domain walls along the () basis vector, as illustrated in Fig. 3(e). It is notable that the group velocity of the edge states decreases and a flat band dispersion is observed, which creates a slow light region within the Brillouin zone. In particular, the slow light dispersions of the domain walls constructed along the lattice directions of and are the same due to mirror symmetry. The simulated electric field amplitude components for the edge modes at 14.47 GHz are shown in Figs. 3(a)–3(c), where the edge mode is strongly localized at the interface. We investigate that anisotropy-induced interfaces profoundly affect edge states and can achieve slow light dispersion. Numerical simulations reveal a transition from gapless chiral edge modes to the gapped flat band dispersion. This is due to the strong anisotropy in the crystal plane induced by the low symmetry of the structure, which provides a new way to reconfigure the lattice for the topological waveguide interface. It is also the key factor to achieve the flat dispersion. Although the topological phases of the two structures forming the topological interface remain essentially unchanged, the interface configuration is changed. The interface geometry supporting the valley-Hall dispersion relation is mirror-symmetric, where the momentum space patterns along the and paths are consistent. In contrast, the interface geometry responsible for flat dispersion relations lacks this regularity. This inconsistency of the dispersion relations along the and paths explains the difference.
To reveal the flatness of the dispersion relation in the 2D crystal structure, we calculate the density of states (DOS). The right panel of Fig. 3(f) shows the DOS values for the and zigzag topological interface dispersion relations. The interface has a very high DOS value around 14.47 GHz, indicating better electromagnetic wave localization and a flatter band. In contrast, the interface has a lower DOS value, suggesting that it does not exhibit a flat band dispersion. Similarly, the left panel of Fig. 3(f) shows the normalized transmission spectra for different interfaces, which reveals a sharp transmission peak for the interface in the frequency range of 14.45–14.50 GHz, while the interface shows a high transmission across a broad frequency range of 12.40–17.26 GHz. This further highlights the difference between valley-Hall edge states and flat band dispersions. Additionally, we present the calculated group velocities of the edge states and in Fig. 3(g), which clearly demonstrates that the anisotropy interface can induce a slow light waveguide.
Closer examination of the flat band dispersion curve is shown in Fig. 3(h), which involves a local zoom into the energy bands depicted in Fig. 3(e). The amplified dispersion curve exhibits a monotonically increasing trend in the range of . In addition, the slope of the dispersion curve nearly remains constant within the range of . The slow light properties of the waveguide modes on the flat band dispersion curve are further characterized quantitatively by using the group velocity , group refractive index , and group velocity dispersion . As the dispersion curve closely shows a straight line within the range of , the group velocity dispersion is approximately equal to zero . To this end, we calculate the intrinsic fields at the frequencies of 14.462 GHz, 14.482 GHz, and 15.502 GHz, indicated by the three blue dots in Fig. 3(h), as well as the electric field intensity distribution perpendicular to the domain wall interface. Figure 3(i) shows that the electric field intensity distributions at these three eigenfrequencies are almost completely overlapped within the PC waveguide channel. The results intuitively demonstrate that the slow light exhibits near-zero group velocity dispersion within the frequency range of 14.462–15.502 GHz.
To demonstrate the stability of the flat bands, we introduce cavity and disorder defects into the domain walls. The results are depicted in Fig. 4(a), which shows that electromagnetic waves can still propagate along the interfaces, implying the robustness of the transmission. In addition, in order to confirm that anisotropy impacts the topological edge states, we investigate the evolution of topological states with different ellipticities , as shown in Fig. 4(b). When both sides of the domain wall interface have the same anisotropy strength , it is found that the bulk gap of the supercell increases with the increasing of , but the frequency ranges of the edge states decrease. Lower panels of Fig. 4(b) show the calculated band dispersion diagrams with , 3, and 5.5. Only as reaches a certain level will the anisotropy induced edge states show a flat dispersion. Moreover, a larger results in a narrower dispersion frequency range and a flatter dispersion. We demonstrate a new method to generate stable flat bands via lattice anisotropy induced by low symmetry. Note that the anisotropy is limited to a maximum of 5.5 by the constraint of the structural physical size. In Fig. 4(c), we set on one side of the domain interface wall with the constant anisotropy strength . The results show that the bulk gap and band dispersion evolution for different have the same trend as those in Fig. 4(b), except for minor changes in the edge mode dispersion.
Figure 4.Robustness and evolution of topological states with anisotropy. (a) Simulated electric field distributions of disorder defect (left panel) and cavity defect (right panel). (b) Evolution of the edge modes with varying anisotropy strength , where both sides of the domain wall have the same . (c) Evolution of the edge modes with varying anisotropy strength where one side of the domain wall has a fixed . Lower panels of (b) and (c) show the calculated band dispersion diagrams with , 3, and 5.5. Bulk states, edge states, and band gap are marked as gray shaded regions, green shaded regions, and white shaded regions, respectively.
The anisotropy interface has induced slow light flattened dispersion and a wide edge band gap, offering an opportunity to study higher-order topological corner states. We construct a 60° zigzag edge corner structure, as depicted in Fig. 5(a). The frequencies of the eigenmodes are presented in Fig. 5(b), where two types of higher-order topological corner states are proved within the edge band gap. The profiles for the two corner modes at the frequencies of 13.33 GHz and 14.29 GHz are displayed in Fig. 5(c). Notably, the Zak phases of the two gapped 1D edge states formed at the corner structures are quantized by mirror symmetries, providing an alternative perspective on understanding topological corner states. The Dirac masses of the one-dimensional edge states exhibit opposite signs for the edges along the and directions, as shown in Fig. 5(a), resulting in in-gap Jackiw–Rebbi soliton states localized at the corners where these edges intersect [50]. We measure the transmission by using four different antenna-probe methods to visualize two corner states and a flat band. A sample used in the experiment is shown in the left panel of Fig. 5(d). The transmission spectra are presented in the right panel of Fig. 5(d), featuring three sharp peaks around 13.33 GHz, 14.29 GHz, and 14.45 GHz. These peaks clearly indicate the presence of the edge and corner modes within the bulk band gap, which is in good agreement with the simulations. Additionally, the appearance of two corner states with distinct local field distributions is attributed to the localization of the Jackiw–Rebbi soliton state at different lattice positions in the corner structure. The result is due to the reduced symmetry. As depicted in Fig. 5(a), there are three distinct zigzag domain walls at the corner structure induced by anisotropy. The energy of the Jackiw–Rebbi soliton state is localized within the angular lattice for type-I higher-order states, influenced solely by the coupling structure of boundary . In contrast, the electromagnetic wave energy is localized at the lattice position adjacent to the corner for type-II higher-order states, influenced by the coupling of boundaries and .
Figure 5.Higher-order topology in corner structure. (a) Schematic image of the 60° zigzag domain walls corner structure. The blue lines show the interface of the domain walls. (b) Eigen modes of the parallelograms supercell corresponding to (a). Inset shows the schematic of the supercell structure. (c) Numerically simulated electric field distribution of the two corner modes at 13.33 GHz and 14.29 GHz. (d) Left panel shows the schematic diagram of experimental set-up and the fabricated corner configuration sample. The green dot represents the source, and four yellow dots indicate different detection positions (one, two, three, and four dots for edge, corner I, corner II, and bulk antenna-probes, respectively). Right panel shows the measured transmission spectra for four different types of antenna-probes.
It is crucial to note that our system can bear close resemblance to symmetry-mismatched heterointerfaces in electronic systems [51–53], and these symmetry-mismatched heterointerfaces exemplify a novel academic concept of utilizing lattice symmetry breaking at the heterojunction interface to achieve a Berry curvature dipole moment within the band structure. Anisotropy introduces an additional degree of freedom; it can support topological flattened edge dispersion and multi-type higher-order states on anisotropic crystal structures with reduced symmetry, enabling the investigation of novel topological phenomena. It is worth mentioning that the exploration of topological similarities and differences induced by anisotropy requires a well-established physical mechanism that explicitly considers irregular geometries. This remains an open question and necessitates further research.
3. CONCLUSIONS
In conclusion, our work reveals crystal anisotropy on topological edge states and higher-order topological states in the symmetric triangular lattice. Anisotropy is induced through the deformation of the interface, and flat band dispersion can be achieved. Additionally, our research demonstrates the presence of multi-type higher-order states in anisotropy-induced corner structures. We have theoretically and experimentally observed that the edge states can be from chiral gapless to gapped flattened dispersion, all while maintaining topological protection. Anisotropy emerges as a pivotal factor in the engineering of topological states. These results offer valuable insights and contribute significantly to the exploration of photonics topological phenomena, highlighting the advancement of our understanding in this field.
APPENDIX A: EVOLUTION OF THE ANISOTROPY STRENGTH AS THE STRUCTURE ELLIPTICITY
Figures 6(a)–6(d) show the evolution of the dispersion relation corresponding to the and paths with different ellipticities. The dispersion relation is modified along the different directions. We note that the difference in dispersion of the and paths is more pronounced with increasing ellipticity factor , which demonstrates that the anisotropy strength of the crystal structure is affected by the ellipticity of the elliptical scatterer. The in-plane anisotropy remains as long as the symmetry of the scatterer is reduced from to in the triangular photonic lattice.
Figure 6.Schematic diagram of dispersion relation with different ellipticities. (a)–(d) Change in projection band corresponding to the and paths with increasing dimensionless shape deviation factor from 1 to 4.5.
The mirror symmetry is broken by rotating the elliptical metal pillars at an angle , as shown in Fig. 1(a). We calculate the lowest two energy bands at the high-symmetry points of the Brillouin zone, and the distinct phase diagrams of the elliptical PC are presented in Figs. 7(a)–7(c). Our findings reveal that pseudo-Kramer degeneracy does not occur simultaneously at the , , and points, which highlight the anisotropy in the three mirror symmetry directions [, , and are the three lattice basis vector directions, as shown in Fig. 1(b)] of the hexagonal lattice.
Figure 7.Anisotropic phase diagrams. (a)–(c) Phase diagrams of lowest two energy bands at the high-symmetry points , , and as a function of , respectively.
APPENDIX C: EFFECT OF STRUCTURE ELLIPTICITY ON THE POSITION OF DEGENERATE POINTS
We consider first a 2D photonic crystal consisting of parallel-plate metal waveguides arranged by a triangular lattice of uniform circular scatterers in air and connected to the top and bottom metal plates. Each meta-atom is a block made of metallic copper, which is impenetrable to electromagnetic waves. The air channels surrounding them form a graphene-like honeycomb lattice; such photonics graphene hosts Dirac cones at the and points, around which the effective Dirac model applies. However, as the circular scatterers are deformed into ellipses, which appear as non-equivalent triangular-shaped rods or asymmetric AB sublattices, the crystal symmetry is reduced from to or from to . Strong symmetry anisotropic scatterers are induced, and the Dirac points move in the hexagonal Brillouin zone. This gives a constant vector potential for a uniformly deformed photonic crystal. Interestingly, the momentum shift can be captured by a relation , where the coefficient depends on the spatial filling ratio of the scatterers. Hence, the vector potential of the Dirac points’ move is influenced by the ellipticity in our proposed structure. The emergence of Dirac points at and is accidental and only occurs when the ellipticity , as shown in Fig. 2(a).
APPENDIX D: REALIZATION OF THE EDGE STATES IN C2v SYMMETRY TRIANGULAR PHOTONIC CRYSTAL
As demonstrated earlier, the topological edge state can appear in symmetry triangular photonic crystals. The anisotropy of the lattice shows its important role in controlling the topological edge dispersion without affecting the realization of the traditional edge state. We numerically simulate the detailed evolution process of the edge state, whose pattern is analogous to the valley-Hall topological edge state, as shown in Fig. 8.
Figure 8.Topological edge states between two PCs with opposite signs of the Dirac mass. (a)–(c) Projected band of edge states corresponding to , , and , respectively.