• Photonics Research
  • Vol. 13, Issue 5, 1321 (2025)
Yi-Ling Zhang1, Li-Wei Wang2, Yang Liu1, Zhao-Xian Chen3, and Jian-Hua Jiang1,4,5,*
Author Affiliations
  • 1School of Physical Science and Technology, and Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow University, Suzhou 215006, China
  • 2School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China
  • 3National Laboratory of Solid State Microstructures, Collaborative Innovation Center of Advanced Microstructures, and College of Engineering and Applied Sciences, Nanjing University, Nanjing 210093, China
  • 4School of Biomedical Engineering, Division of Life Sciences and Medicine, University of Science and Technology of China, Hefei 230026, China
  • 5Suzhou Institute for Advanced Research, University of Science and Technology of China, Suzhou 215123, China
  • show less
    DOI: 10.1364/PRJ.550282 Cite this Article Set citation alerts
    Yi-Ling Zhang, Li-Wei Wang, Yang Liu, Zhao-Xian Chen, Jian-Hua Jiang, "Floquet hybrid skin-topological effects in checkerboard lattices with large Chern numbers," Photonics Res. 13, 1321 (2025) Copy Citation Text show less

    Abstract

    Non-Hermitian topology provides an emergent research frontier for studying unconventional topological phenomena and developing new materials and applications. Here, we study the non-Hermitian Chern bands and the associated non-Hermitian skin effects in Floquet checkerboard lattices with synthetic gauge fluxes. Such lattices can be realized in a network of coupled resonator optical waveguides in two dimensions or in an array of evanescently coupled helical optical waveguides in three dimensions. Without invoking nonreciprocal couplings, the system exhibits versatile non-Hermitian topological phases that support various skin-topological effects. Remarkably, the non-Hermitian skin effect can be engineered by changing the symmetry, revealing rich non-Hermitian topological bulk-boundary correspondences. Our system offers excellent controllability and experimental feasibility, making it appealing for exploring diverse non-Hermitian topological phenomena in photonics.

    1. INTRODUCTION

    Non-Hermitian physics provides an effective framework for investigating non-equilibrium and open quantum systems as well as their classical counterparts, wherein dissipation, energy gain, and nonreciprocal couplings give rise to physical effects that are unattainable in conventional Hermitian systems. Rich non-Hermitian phenomena such as unidirectional invisibility [1], non-Hermitian skin effects (NHSEs) [26], exceptional points, lines, and surfaces [715], non-Bloch bands [1619], non-Hermitian topology [2023], and non-Hermitian dynamics [2427] were found in the past decades. A particularly interesting frontier lies at the interface between topological physics and non-Hermitian physics. For instance, the NHSEs can be characterized by the spectral topology in the complex plane for finite non-Hermitian systems [2,28]. Moreover, in non-Hermitian topological systems, the non-Hermitian skin modes coexist with topological edge modes, which thus extends the bulk-edge correspondence to a more general picture [29]. The interplay between the NHSEs and the topological edge modes can also bring unconventional effects such as the skin-topological effect [3033] wherein the corner skin modes originate from the NHSEs of the topological edge states.

    On the other hand, Floquet lattice systems provide an efficient pathway toward engineering desired Hamiltonians and realizing the underlying physics [3438]. In particular, the periodic modulation in Floquet systems introduces artificial gauge fields and enables access to various topological phases and the NHSEs without nonreciprocal couplings. Thus, Floquet non-Hermitian systems provide a rich realm for exploring dynamic and topological phenomena [3943]. With experimental feasibility in photonic and acoustic systems, skin-topological effects are especially accessible in Floquet non-Hermitian systems [30,32,44]. However, the existing studies on skin-topological effects focus on the relatively simple regime where the spatial symmetry dictates the phenomenon and there is only one branch of edge states [33]. Much remains unknown when the Chern number (and hence the number of the branches of edge states) is varied and spatial symmetry is controllable, which may offer a valuable platform and unconventional mechanisms for steering light by the interplay between topology and non-Hermicity [45].

    In this work, we study the interplay between the NHSE and the topological physics in checkerboard lattices where we find controllable Floquet skin-topological effects with varying Chern numbers. At the heart of our theory is the creation of a Floquet non-Hermitian checkerboard lattice model wherein rich, controllable non-Hermitian topological phases induced only by gain and loss can be realized. For concreteness, we design an array of evanescently coupled helical optical waveguides in three dimensions to realize the model. Nevertheless, the underlying physics is more general and is not restricted to photonic systems. For instance, the same physics can be realized also in cold atom systems where the synthetic gauge flux can be realized via other mechanisms [4648]. We find that our checkerboard lattices can efficiently generate non-Hermitian bands with large Chern numbers. The induced multiple chiral edge states have rich non-Hermitian skin effects due to the fact that each branch of the chiral edge states has its own non-Hermitian properties. The total skin-topological effect comes from the NHSEs of the multiple chiral edge states as well as their interaction. We find that the scattering between these non-Hermitian topological edge channels at the corner boundaries can induce an unconventional mechanism for the corner skin effect, which is distinct from the previous mechanisms solely associated with the nonreciprocal properties of the non-Hermitian edge states. With a simple and controllable design, our system is found to support non-Hermitian hybrid skin-topological effects, exhibiting rich corner skin effects that can be engineered by varying the lattice symmetry.

    The structure of this paper is organized as follows. In Section 2.A, we introduce the non-Hermitian effective model of photonic checkerboard lattices and the topological phases, which provide the groundwork for realizing the NHSE. In Section 2.B, we investigate the mixed skin-topological effects in the non-Hermitian effective model with C4 symmetry. We analyze the origin of corner skin modes from a dynamical perspective. The locations of different corner skin modes are influenced by various non-equilibrium interactions of topological edge states. In Section 2.C, we explore the non-Hermitian effective model with C2 symmetry. Corner skin modes still exist, and we present the dynamic analysis process, finding that the locations of corner skin modes depend not only on the signs of group velocity and the gain/loss of topological edge states, but also on the contributions from the topological edge states. Finally, we conclude in Section 3.

    2. RESULTS

    A. Model and Chern Insulator Phases

    We introduce interstitial sites between the main sites of the two-dimensional square potential or Su-Schrieffer-Heeger (SSH) lattice, as shown in Fig. 1. In each unit cell (marked by the dotted-dashed box), an interstitial site (e) is added in the intermediate region between the four sublattice sites (a, b, c, d), forming a coupling path between each pair of sublattice sites. There is an inhomogeneous dissipation distribution and an imbalanced on-site potential in each unit cell. In this case, the z-dependent Hamiltonian in this model is given byH(z)=HSSH+Hloss-potential(z)+Hinter-SSH(z),where HSSH and Hloss-potential(z) respectively describe the main sites and their dissipation and on-site potential distributions, whereas Hinter-SSH is the coupling between the interstitial sites and main sites of sublattice. We take HSSH=i,jt1(ai,jbi,j+di,jci,j+bi,jdi,j+ai,jci,j)+i,jt2(ai+1,jbi,j+di+1,jci,j+bi,j+1di,j+ai,j+1ci,j)+h.c.,where t1 and t2 represent the intra-cell and inter-cell coupling strengths in the 2D SSH lattice, respectively. ξi,j(ξi,j)(ξ{a,b,c,d}) creates (annihilates) the operator at the site (i,j) and h.c. denotes the Hermitian conjugate. The coupling interaction between the two sub-systems is characterized by Hinter-SSH(z)=i,jm=14gm(z)(ei,jξi,j). The clockwise rotation involves time-dependent nearest-neighbor interactions [with couplings g1(z),g2(z),g3(z),g4(z)], as shown in Fig. 1. The on-site potential at the sublattice sites (a, b, c, d), with dissipation γ, is given by Hloss-potential(z)=i,j(Vi,j(z)iγ)ξi,jξi,j.We consider a sinusoidal modulation Vi,j(z)=sin(2πz/T+2πm/4), with a chiral phase shift of 2π/4 between the interstitial sites in the intermediate region of each main site, where m1,2,3,4 and T denotes the Floquet period along the z direction. In the periodical Floquet systems, the static effective Hamiltonian Heff is defined as Heff=i/Tlog(U),where U=Texp(i0TH(z)dz) describes the evolution operator U for one period T, and T is the time-ordering. In the high driving frequency regime, Heff can be a perturbation expansion in powers of 1/T (see details in Appendix A).

    Floquet potentials in a checkerboard lattice (left panel). A unit cell (marked by the dotted box) has five sites (labeled as a, b, c, d, e). The four sites, with loss, for the on-site potential ΔV are modulated in sinusoidal form (right panel). Such Floquet NHSE systems offer rich non- Hermitian topological phase and states in a geometrically static arrangement.

    Figure 1.Floquet potentials in a checkerboard lattice (left panel). A unit cell (marked by the dotted box) has five sites (labeled as a, b, c, d, e). The four sites, with loss, for the on-site potential ΔV are modulated in sinusoidal form (right panel). Such Floquet NHSE systems offer rich non- Hermitian topological phase and states in a geometrically static arrangement.

    Next, we analyze whether the edge states are topological in the effective model. Generally, topological invariants are used to characterize them. Notably, unlike the Hermitian case, for complex non-Hermitian systems in two dimensions, topological invariants can be calculated in real space. Detailed calculations can be found in Appendix B. To verify the experimental feasibility of the results, and to ensure that the model exhibits a large Chern number at a specific coupling strength, we consider the loss coefficient to be at the value of (0,1). Using this method, we compute the real-space topological invariants of the loss γ(0,1) as a function of g1(g3) and g2(g4), resulting in the phase diagrams for gaps I/IV and II/III, as illustrated in Figs. 2(a) and 2(b), respectively. For coupling strengths gm(0,1.5), the system is gapless, making the calculation of the Chern number infeasible. As the coupling strength increases, topological phases with a Chern number C=1 emerge in gaps I and IV. However, with further increases in coupling strength, these phases eventually disappear. In gaps II and III, a larger nontrivial Chern number C=2 is observed, corresponding to two pairs of topological chiral edge states. The results for both gaps are consistent with the projected band structure.

    The topological phase of the non-Hermitian effective model. (a), (b) The real-space Chern numbers calculated for gaps I/IV and gaps II/III as functions of g1(g3) and g2(g4) under loss γ∈(0,1), respectively. The Chern numbers are distinguished by different colors. The other parameters are set as θm=(m−1)π/2, where m=1,2,3,4.

    Figure 2.The topological phase of the non-Hermitian effective model. (a), (b) The real-space Chern numbers calculated for gaps I/IV and gaps II/III as functions of g1(g3) and g2(g4) under loss γ(0,1), respectively. The Chern numbers are distinguished by different colors. The other parameters are set as θm=(m1)π/2, where m=1,2,3,4.

    B. High Chern Number Topological Edge States and Hybrid Skin-Topological Modes under C4 Symmetry

    The non-Hermitian properties of a kagome lattice under both C3 symmetry and broken C3 symmetry conditions are analyzed in Ref. [33], revealing corner modes induced by non-Hermitian skin effects arising from topological edge states. Furthermore, by analyzing the dynamic behavior and interactions of edge states at the corners, the study demonstrates how the skin effect localizes these states, accurately predicting hybrid topological-skin modes. This analytical approach is also applicable to our upcoming investigation of the non-Hermitian properties of the model under C4 symmetry. However, unlike the previous research, which only analyzes a pair of edge states at the corners, we extend our study to multiple pairs of edge states at the corners, and similarly, accurately predict the emergence of hybrid topological-skin modes.

    The model possesses C4 symmetry and exhibits non-Hermitian properties, allowing easy engineering of the coupling strengths as g1,2,3,4=0.5, with the loss coefficient set to γ=0.6. Then, based on the phase diagrams (Fig. 2), one infers that two pairs of chiral edge states exist within band gaps II and III, as clearly evidenced in the real part of projected band structure depicted in Fig. 3(a). Moreover, the projected band structure exhibits as gapless, with edge states marked as dark purple and light purple in Fig. 3(a). This situation is not further analyzed in this paper. We use green (orange) to represent the eigenstates localized at the upper (lower) edge. This labeling facilitates the subsequent analysis of the hybrid skin-topological modes. The corresponding complex energy spectra for the non-Hermitian effective model are shown in Fig. 3(b), from top to bottom, corresponding to three different boundary conditions (xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC). PBC (OBC) represents the periodic (open) boundary condition in the x or y direction. A schematic of the regular rectangular supercell with xOBC/yOBC is presented in Fig. 3(c). The cancellation of nonreciprocity arising from the symmetry of the local flux, along with a balance between gain and loss within the bulk, as explained in Ref. [30], leads to an almost unchanged complex energy spectrum of the bulk continuum. This invariance is illustrated in Fig. 3(b) under different boundary conditions.

    The non-Hermitian properties of this effective model under the C4 rotation symmetry. (a) The real part of the projected band structure. The Chern numbers for band gaps II and III are labeled. The edge states within band gaps II and III are shown in green and orange, respectively, depending on their localization at the upper and lower edges. (b) The complex energy spectra with xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC, respectively, from top to bottom. The olive dots represent the degenerate edge states. The hybrid skin-topological corner modes are indicated in red. (c) The schematic depiction of the quadrilateral supercell with xOBC/yOBC. The black dots represent omitted unit cells. The number of sites for xPBC/yOBC and xOBC/yOBC is 60 and 3380, respectively. The parameters are t1=1, t2=2, g1,2,3,4=0.5, and γ=0.6.

    Figure 3.The non-Hermitian properties of this effective model under the C4 rotation symmetry. (a) The real part of the projected band structure. The Chern numbers for band gaps II and III are labeled. The edge states within band gaps II and III are shown in green and orange, respectively, depending on their localization at the upper and lower edges. (b) The complex energy spectra with xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC, respectively, from top to bottom. The olive dots represent the degenerate edge states. The hybrid skin-topological corner modes are indicated in red. (c) The schematic depiction of the quadrilateral supercell with xOBC/yOBC. The black dots represent omitted unit cells. The number of sites for xPBC/yOBC and xOBC/yOBC is 60 and 3380, respectively. The parameters are t1=1, t2=2, g1,2,3,4=0.5, and γ=0.6.

    Before specifically analyzing the hybrid skin-topological modes, we need to focus on the group velocity of the edge states and the imaginary part of the energy. The group velocity of the chiral edge states can be calculated from the specific wave vector k in the projected band structure, i.e., v=Re(E(k))k. The sign of the group velocity indicates the direction in which the wave packet moves along the boundary. Furthermore, the magnitude of the imaginary part of the energy of the edge states causes an imbalance in the lifetime of the wave packet, meaning the wave packet will either amplify or decay as it propagates along the boundary.

    Let us analyze a specific region, A. In Fig. 3(b), for the same real energy, the degenerate states with higher imaginary energy, marked in olive color, are composed of the upper edge state ψgreen,I and the lower edge state ψorange,I, while the degenerate states with lower imaginary energy are composed of the upper edge state ψgreen,II and the lower edge state ψorange,II. Meanwhile, the upper (lower) edge states marked in green (orange) have a negative (positive) group velocity, indicating that the wave packet propagates in the negative direction along that boundary, with the rightward direction along the lower boundary defined as the positive direction. Furthermore, applying OBC to the rectangular supercell, the upper and left boundaries correspond to the upper edge states, while the lower and right boundaries correspond to the lower edge states [see Fig. 3(c)]. It is worth noting that, when comparing the complex energy spectra of xPBC/yOBC and xOBC/yOBC in Fig. 3(b), at the same real energy, the imaginary energy in region A satisfies the following relation: Im(Egreen/orange,II,xPBC/yOBC)<. Im(Ered-up/-down,xOBC/yOBC)<Im(Egreen/orange,I,xPBC/yOBC). Moreover, all the gap states marked in red are localized at the four corners of the system, as shown in Fig. 4.

    Non-Hermitian C4 symmetric effective model with corner skin modes and configurations. (a) The top of the schematic for corner skin modes existing in the configurations of region A. Blue (red) arrows represent chiral edge states exhibiting relative attenuation (amplification), accumulating waves in their propagation (opposite) directions. Glowing purple spheres indicate potential localized corner skin modes. The bottom of (a) shows corner skin modes with higher imaginary energy in the parallelogram supercell. The height of the columns in the supercell is proportional to the amplitude at each atomic position. (b)–(d) Similar to the explanation for (a).

    Figure 4.Non-Hermitian C4 symmetric effective model with corner skin modes and configurations. (a) The top of the schematic for corner skin modes existing in the configurations of region A. Blue (red) arrows represent chiral edge states exhibiting relative attenuation (amplification), accumulating waves in their propagation (opposite) directions. Glowing purple spheres indicate potential localized corner skin modes. The bottom of (a) shows corner skin modes with higher imaginary energy in the parallelogram supercell. The height of the columns in the supercell is proportional to the amplitude at each atomic position. (b)–(d) Similar to the explanation for (a).

    Based on the above analysis, we can derive the directions of wave packet propagation on the four boundaries and their relative amplification or attenuation during propagation. As shown at the top of Fig. 4, the configurations of the hybrid skin-topological modes for regions A and B are displayed, where arrows indicate the direction of wave packet propagation, with blue (red) representing relative amplification (attenuation) along their respective directions. Further analyzing the configuration of the hybrid skin-topological modes with higher imaginary energy in region A, at the top of Fig. 4(a), we observe two types of wave packets exhibiting directional amplification or attenuation along the boundaries. For example, one dynamic type is as follows: the wave packet ψgreen,I propagates leftward along the upper edge, has a longer lifetime, and eventually collapses to the upper-left corner over time. In contrast, the wave packet ψgreen,II moves downward along the left edge, with a shorter lifetime and exponential decay. This dynamic type will localize at the corner, labeled as 2 (for ψgreen,I and ψgreen,II), and this type of dynamics will eventually result in localized corner skin modes at all four corners, as shown by the glowing purple spheres in the top of Fig. 4(a). In contrast, another dynamic type involves the wave packet ψgreen,II propagating leftward along the upper edge, with a shorter lifetime and faster decay. Meanwhile, ψgreen,I moves downward along the left edge, has a longer lifetime, and eventually collapses to the lower-left corner during dynamic evolution. This shows that the interaction between these two wave packets at the upper-left corner does not lead to localization, so we do not consider this dynamic type.

    A similar analysis has been conducted for the other corner skin modes, as shown at the top of Figs. 4(b)–4(d), where the corresponding corner skin mode configurations are displayed. It can be observed that the configurations are the same because the Chern number in gaps II and III is +2 for both. The bottoms of Figs. 4(a) and 4(b) [4(c) and 4(d)] show the hybrid topological skin modes with higher (lower) imaginary energy, respectively, where the height of the columns on the plot is proportional to the amplitude strength at each atomic site.

    C. High Chern Number Topological Edge States and Hybrid Skin-Topological Modes under C2 Symmetry

    To demonstrate the feasibility of the system and to extend the discussion to a richer hybrid skin effect, we will investigate the non-Hermitian effective model with C2 symmetry, where the coupling strengths are g1,3=1 and g2,4=2, and the dissipation coefficient is γ=1. As shown in Fig. 5(a), the real part of the projected band structure reveals that two pairs of edge states still exist in band gaps II and III, marked in the same way as before, while there is one pair of edge states in band gaps I and IV. Figure 5(b) corresponds to the complex energy spectra under three boundary conditions: xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC, from the top to bottom. Notably, under the xPBC/yOBC and xOBC/yOBC boundary conditions, distinct crossing points are observed within both band gaps II and III. At these crossing points, the imaginary part of the energy reverses on either side. Consequently, these crossings divide band gap II into regions A and B, and band gap III into regions C and D. In regions A and D, the real energy is the same, and the imaginary energy of both the upper and lower supercell states lies between the chiral edge states. In regions B and C, only the imaginary energy of the upper supercell state is positioned between the chiral edge states, while the imaginary energy of the lower supercell state is located below the edge states. This slight shift in the edge states may be attributed to the C2 symmetry of the structure. However, it does not affect the validity of our subsequent analysis. And this clear distinction in imaginary energy distribution across different regions aids in understanding the properties of the corner skin modes that will be discussed later.

    The non-Hermitian properties of the effective model under C2 symmetry. (a) The real part of the projected band structure with xPBC/yOBC. The edge states within band gaps II and III are labeled the same as in Fig. 3(a). (b) The complex energy spectra with xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC, respectively, from the top to bottom. The olive dots represent the degenerate edge states. The hybrid skin-topological corner modes are indicated in red. The number of sites for xPBC/yOBC and xOBC/yOBC is 60 and 3380, respectively. The parameters are t1=1, t2=2, g1,3=1, g2,4=2 and γ=1.

    Figure 5.The non-Hermitian properties of the effective model under C2 symmetry. (a) The real part of the projected band structure with xPBC/yOBC. The edge states within band gaps II and III are labeled the same as in Fig. 3(a). (b) The complex energy spectra with xPBC/yPBC, xPBC/yOBC, and xOBC/yOBC, respectively, from the top to bottom. The olive dots represent the degenerate edge states. The hybrid skin-topological corner modes are indicated in red. The number of sites for xPBC/yOBC and xOBC/yOBC is 60 and 3380, respectively. The parameters are t1=1, t2=2, g1,3=1, g2,4=2 and γ=1.

    More importantly, the dynamical behavior of chiral edge states is influenced not only by the imaginary energy and the associated group velocity but also by the weight of the chiral edge states in specific corner skin modes, denoted as cj (j=1,2,3,4), where j represents the j-th edge state. This factor must be considered, as in the coupling interaction g1,3=1 and g2,4=2, the Chern number C=2 emerges in gaps II and III, indicating the presence of two pairs of topological edge states rather than a single pair. Moreover, the C2 symmetry of the model alters the three factors that affect the dynamical behavior of the chiral edge states, resulting in a more diverse set of hybrid skin-topological modes. Thus, we need to redefine the previous analysis and provide a detailed derivation (see details in Appendix C).

    We redefine Pjcjexp(Ejtj) (j=1,2,3,4), where the imaginary part of the energy Ej, time tj, and coefficient cj are used to calculate the impact of wave functions during the dynamic process along different edges. In this definition, the imaginary part of Ej represents the difference between the edge state and the imaginary energy of the corner skin mode (as reference energy). The time tj is determined by the group velocity of the edge state, representing the time required for a wave packet to propagate a unit lattice distance along the edge. Notably, for a wave packet with a longer lifetime, it will gradually collapse to the corner when interacting with another wave packet of a shorter lifetime, and the product operation of Pj gives the magnitude of the wave function distribution at that corner. Assuming the initial wave function distribution at a corner is Pj=cj, it will amplify or decay relative to the motion along their respective boundaries. By comparing the wave function distributions at different corners, we can assess the manifestation of the hybrid topological skin effect at each corner.

    Based on the existing analytical methods, we also need to consider the group velocity of the chiral edge states, which is the same as in the case with C4 symmetry considered previously. For convenience in subsequent analysis, the complex energy spectrum of xPBC/yOBC in Fig. 5(b) shows that, for the same real energy, the degenerate states (marked in olive) with higher imaginary energy in region A and lower imaginary energy in region B are composed of the upper edge state ψgreen,I and the lower edge state ψorange,I, while the degenerate states with lower imaginary energy in region A and higher imaginary energy in region B are composed of the upper edge state ψgreen,II and the lower edge state ψorange,II. The degenerate states in regions C and D are composed of edge states with opposite labels compared to those in regions A and B.

    Figure 6 shows the corresponding configurations of corner skin modes in different regions. The configurations on the left are derived from the analysis on the right, reflecting the transition from C4 symmetry to C2 symmetry. We will now focus on the analysis of region A, with the corresponding configuration shown in Fig. 6(a). For instance, we choose the corner skin mode with higher imaginary energy [Im(E)=0.8332]. In this case, the configuration could either be Type 1 or Type 2. First, we consider the Type 1 configuration. Suppose the initial wave packet ψgreen,I has a longer lifetime and propagates leftward along the upper edge. At this point, Pb3=1.0616, and after some time, the wave packet localizes at the upper-left corner, yielding Pb1=cgreen,Iexp((Egreen,IExOBC/yOBC)tgreen,I)=1.0801. Simultaneously, the wave packet ψgreen,II, with a shorter lifetime, propagates downward along the left edge and decays, yielding Pb4=Pb1×Pgreen,II=1.0552. Next, the wave packet ψorange,I, with a longer lifetime, moves rightward along the lower edge and eventually localizes at the lower-right corner, leading to Pb2=Pb1×Pb4=1.1402. Clearly, the Type 1 configuration corresponds to the corner skin mode. Similarly, we can calculate the Type 2 configuration, and find that it also fits the corner skin mode. Finally, by comparing the P-values of the wave packets at each corner for both configurations [as shown at the bottom of Fig. 7(a)], we find that Pb1 and Pb2 for Type 1 are greater than Pb3 and Pb4 for Type 2. Therefore, the configuration in this mode is Type 1, as shown at the top of Fig. 7(a), with the corresponding corner skin mode at the bottom. The configuration analysis for other regions follows a similar procedure (for more details, see Fig. 6).

    (a), (b) Configuration diagram of the supercell, with the left side showing the predicted corner skin modes at all corners, and the right side showing the predicted corner skin modes in regions A, D, and B, C under the C2 symmetry.

    Figure 6.(a), (b) Configuration diagram of the supercell, with the left side showing the predicted corner skin modes at all corners, and the right side showing the predicted corner skin modes in regions A, D, and B, C under the C2 symmetry.

    Non-Hermitian C2 symmetric effective model with corner skin modes and configurations. (a) The top of the schematic for corner skin modes existing in the configurations of region A. The bottom of (a) shows corner skin modes with higher imaginary energy in the parallelogram supercell. The bottom right of (a) shows the P values of the corner skin modes at the four corners, where Pb1, Pb2, Pb3, and Pb4 correspond to the positions shown at the top. (b)–(h) Similar to the explanation for (a).

    Figure 7.Non-Hermitian C2 symmetric effective model with corner skin modes and configurations. (a) The top of the schematic for corner skin modes existing in the configurations of region A. The bottom of (a) shows corner skin modes with higher imaginary energy in the parallelogram supercell. The bottom right of (a) shows the P values of the corner skin modes at the four corners, where Pb1, Pb2, Pb3, and Pb4 correspond to the positions shown at the top. (b)–(h) Similar to the explanation for (a).

    Through similar analysis, Figs. 7(a)–7(d) and 7(e)–7(h) show the corresponding configurations and corner skin effects for regions A, D, and B, C, respectively. The top of Fig. 7 presents the corresponding configuration diagrams, while the bottom displays the observed corner skin effect patterns. Notably, for the same real energy, the imaginary energy of the corner skin modes in Figs. 7(f) and 7(h) is lower than that of the edge states under xPBC/yOBC boundary conditions. In this case, we still need to refer to the imaginary energy of the supercell, but this does not affect the value of Pj. Similarly, we can calculate Pb1, Pb2, Pb3, and Pb4, and derive the corresponding configurations, as shown in the top of Figs. 7(f) and 7(h). Here, the introduced concept Pjcjexp(Ejtj) provides a more universally applicable method for analyzing multiple hybrid skin-topological modes, enriching the study of symmetry models.

    3. CONCLUSION AND DISCUSSION

    We establish a non-Hermitian Chern insulator model based on checkerboard lattices. Despite its simplicity, the model encompasses rich non-Hermitian topological phases. Of special interest here are the non-Hermitian topological band gaps with large Chern numbers that support multiple chiral edge states. In finite systems, a corner boundary can trigger the scattering among these chiral edge states without causing back scattering, making the bulk-boundary correspondence more versatile. Interestingly, we find that the corner-induced scattering leads to a higher-order non-Hermitian skin effect where the waves are trapped at the corner boundaries. Depending on the system’s symmetry and parameters, the corner skin modes can appear at all the four corners or only at a pair of diagonal corners. This phenomenon is different from the previously found hybrid skin-topological effect where the system has only one branch of non-Hermitian chiral edge states and the corner skin effect is solely due to the gain and loss of the chiral edge states as determined often by the symmetry of the system. Instead, here, we reveal through the dynamical analysis that the corner-induced scattering plays an important role in the higher order non-Hermitian skin effect, which is unique for non-Hermitian topological phases with large Chern numbers.

    We emphasize that the checkerboard lattice model proposed in this work can be realized in a number of systems. First, as studied here, a photonic Floquet lattice consisting of coupled optical waveguides can realize the model, which has been realized in experiments [35]. Second, in cold atom systems where the artificial gauge fields can be generated through the scheme of optical flux lattices [46], the checkerboard lattice model can also be realized if the dissipation is engineered as in a recent work [48].

    We remark that this work not only adds to the fundamental research in non-Hermitian topological physics by enriching the mechanisms for non-Hermitian topological bulk-boundary correspondence, but also facilitates the development of non-Hermitian topological photonics, which may enable the realization of unconventional photonic materials and applications.

    Acknowledgment

    Acknowledgment. We are grateful to Dr. Weiwei Zhu for valuable discussions.

    APPENDIX A: ANALYZE THE EFFECTIVE HAMILTONIAN

    In Floquet systems (i.e., periodically driven systems), gauge transformations may also involve redefinition in the time domain. When analyzing the Floquet effective Hamiltonian, an appropriate gauge transformation can eliminate the time dependence, transforming it into a static equivalent description, thus simplifying the physical interpretation. We will consider artificial potential engineering as much as possible to obtain the desired coupled phases. One can consider such a Hamiltonian [Eq. (1)], and the time-dependent Schrödinger equation is given by iz|ψ(z)=H(z)|ψ(z).

    Here, we apply a unitary gauge transformation to the evolved state |ψ(z)=u(z)|ϕ(z), where u(z)=i,jexp(iΦ(z)z)·Ki,jKi,j(K{a,b,c,d,e}). Plugging |ψ(z) into Eq. (A1), we get the transformed Schrödinger equation iz|ϕ(z)=H(z)|ϕ(z),where H(z)=u(z)H(z)u(z)+iu(z)zu(z), and we can obtain H(z)=i,jt1(ai,jbi,j+di,jci,j+bi,jci,j+ai,jdi,j)+i,jt2(ai+1,jbi,j+di+1,jci,j+bi,j+1ci,j+ai,j+1di,j)+i,j(G1(z)(ei,jai,j)+G2(z)(ei,jbi,j))+i,j(G3(z)(ei,jci,j)+G4(z)(ei,jdi,j)+iγ(ξi,jξi,j))+h.c.The modulation parameters are Gm(z)=gmexp(iAsin(ωz  +θm))(m{1,2,3,4}). Here, θm=0z(θi,j,eθi,j,ξ)dz(ξ{a,b,c,d,e}). The discrete Fourier component of Hamiltonian Hn is Hn=i,jt1(ai,jbi,j+di,jci,j+bi,jci,j+ai,jdi,j)+i,jt2(ai+1,jbi,j+di+1,jci,j+bi,j+1ci,j+ai,j+1di,j)+i,jgmeinθm((ei,jai,j)+(ei,jbi,j)+(ei,jci,j)+(ei,jdi,j))+i,jiγ(ξi,jξi,j)+h.c.

    In the high frequency expansion of Heff, Heff=H0+(H1,H1)ω+O(1ω2)=i,jt1(ai,jbi,j+di,jci,j+bi,jci,j+ai,jdi,j)+i,jt2(ai+1,jbi,j+di+1,jci,j+bi,j+1ci,j+ai,j+1di,j)+1ωi,j(g1eiθ1(ei,jai,j)+g2eiθ2(ei,jbi,j))+1ωi,j(g3eiθ3(ei,jci,j)+g4eiθ4(ei,jdi,j))+i,jiγ(ξi,jξi,j)+h.c.

    In this case, the driving frequency ω can be neglected.

    APPENDIX B: NON-HERMITIAN TOPOLOGICAL PHASE

    So far, there are two methods to characterize the topological properties of non-Hermitian systems using topological invariants: one involves calculating the non-Bloch Chern number by integrating the Berry curvature over the generalized Brillouin zone, and the other involves computing the Chern number in real space on a finite lattice. In one-dimensional cases, the generalized Brillouin zone can be quickly determined, but for two-dimensional cases, it is generally challenging to obtain. Here, we consider the second method, which involves using the Kitaev formula to compute the Chern number [33,49]. Consider a finite lattice with a circle of radius r, which is divided into three regions: A, B, and C (see the inset of Fig. 8). We can get Creal-space=12πijAkBlC(PjkPklPljPjlPlkPkj). Here, j, k, and l represent the positions of these three regions, respectively. And P^=En<Ef|ψR(n)ψL(n)| is the projection operator, En is the n-th eigenenergy, and Ef is the Fermi energy. Finally, the real-space Chern numbers of the model can be calculated (see Fig. 2).

    (a), (b) The real-space Chern number of non-Hermitian checkerboard lattices as a function of the Fermi energy, where (a) represents C4 symmetry and (b) represents C2 symmetry. The inset shows the schematic diagram of the supercell used to calculate the real-space Chern number.

    Figure 8.(a), (b) The real-space Chern number of non-Hermitian checkerboard lattices as a function of the Fermi energy, where (a) represents C4 symmetry and (b) represents C2 symmetry. The inset shows the schematic diagram of the supercell used to calculate the real-space Chern number.

    Specifically, we further calculate the real-space Chern number C as a function of the Fermi energy Ef for the non-Hermitian checkerboard lattices with C4 and C2 symmetry, as shown in Figs. 8(a) and 8(b). The parameters used here are the same as those in the main text. It is clearly observed that within these energy gap regions, the computed Chern numbers are integer values. For example, when considering the C4 symmetry, the real-space Chern number is +2 in gaps II and III, indicating the presence of two pairs of topological edge states.

    APPENDIX C: WAVE PACKETS DURING DYNAMIC EVOLUTION OF THE EIGENSTATE

    Here, the final state |ψ(t,n) can be obtained from any given moment as |ψ(t,n)=U(t)|ψ(t0,n),where U(t)eiHOBCt represents the time evolution operator. By exciting an initial state |ψ(t0,n) using a Hamiltonian HOBC under OBC, the final state |ψ(t,n) is obtained through time evolution. Next, we can perform a Fourier transform of the real-space state |ψ(t,n) into the general k-space |ψ(t,k): |ψ(t,k)=n=1N|ψ(t,n)eikxn.

    Thus, the final state |ψ(t,k) with contributing wave vectors k during the evolution process can be decomposed into the eigenstates of the non-Bloch Hamiltonian, and the component coefficients cj(t,k) can be calculated: |ψ(t,k)=jcj(t,k)|ϕj,R(k),cj(t,k)=ϕj,L(k)|ψ(t,k),where j represents the degrees of freedom of the non-Bloch Hamiltonian; the right and left eigenvectors of the non-Bloch Hamiltonian are |ϕj,R(k) and ϕj,L(k)|, respectively, and they satisfy the biorthogonality condition ϕi,L(k)|ϕj,R(k)=δi,j.

    Therefore, at any moment, Eq. (C1) can be rewritten by analyzing the contributions to the eigenstates of the Hamiltonian HOBC quantities for the final state: |ψ(t,k)=jϕj,L(k)|ψ(t0,k)exp(iEjt)|ϕj,R(k)=jcj(t,k)eiRe(Ej)teIm(Ej)t|ϕj,R(k).

    At this point, Ej=Re(Ej)+iIm(Ej) is a complex energy. When considering the magnitude of the imaginary energy, we can define Pcjexp(Im(Ej)t), which means that the magnitude of the P value determines the directional amplification or decay effect during the dynamic evolution of the eigenstates.

    References

    [1] L. Feng, Y.-L. Xu, W. Fegadolli. Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies. Nat. Mater., 12, 108-113(2013).

    [2] N. Okuma, K. Kawabata, K. Shiozaki. Topological origin of non-Hermitian skin effects. Phys. Rev. Lett., 124, 086801(2020).

    [3] K. Kawabata, M. Sato, K. Shiozaki. Higher-order non-Hermitian skin effect. Phys. Rev. B, 102, 205118(2020).

    [4] L. Li, C. H. Lee, S. Mu. Critical non-Hermitian skin effect. Nat. Commun., 11, 5491(2020).

    [5] X. Zhu, H. Wang, S. K. Gupta. Photonic non-Hermitian skin effect and non-Bloch bulk-boundary correspondence. Phys. Rev. Res., 2, 013280(2020).

    [6] X.-R. Ma, K. Cao, X.-R. Wang. Non-Hermitian chiral skin effect. Phys. Rev. Res., 6, 013213(2024).

    [7] C. M. Bender, S. Boettcher. Real spectra in non-Hermitian Hamiltonians having PT symmetry. Phys. Rev. Lett., 80, 5243-5246(1998).

    [8] R. El-Ganainy, K. G. Makris, M. Khajavikhan. Non-Hermitian physics and PT symmetry. Nat. Phys., 14, 11-19(2018).

    [9] T. Suchanek, K. Kroy, S. A. M. Loos. Time-reversal and parity-time symmetry breaking in non-Hermitian field theories. Phys. Rev. E, 108, 064123(2023).

    [10] K. Jones-Smith, H. Mathur. Non-Hermitian quantum Hamiltonians with PT symmetry. Phys. Rev. A, 82, 042101(2010).

    [11] L. Feng, R. El-Ganainy, L. Ge. Non-Hermitian photonics based on parity-time symmetry. Nat. Photonics, 11, 752-762(2017).

    [12] M.-A. Miri, A. Alu. Exceptional points in optics and photonics. Science, 363, eaar7709(2019).

    [13] I. Mandal, E. J. Bergholtz. Symmetry and higher-order exceptional points. Phys. Rev. Lett., 127, 186601(2021).

    [14] Y.-C. Tzeng, C.-Y. Ju, G.-Y. Chen. Hunting for the non-Hermitian exceptional points with fidelity susceptibility. Phys. Rev. Res., 3, 013015(2021).

    [15] S. Ramezanpour, A. Bogdanov. Tuning exceptional points with Kerr nonlinearity. Phys. Rev. A, 103, 043510(2021).

    [16] K. Yokomizo, S. Murakami. Non-Bloch band theory of non-Hermitian systems. Phys. Rev. Lett., 123, 066404(2019).

    [17] S. Yao, Z. Wang. Edge states and topological invariants of non-Hermitian systems. Phys. Rev. Lett., 121, 086803(2018).

    [18] H.-Y. Wang, F. Song, Z. Wang. Amoeba formulation of non-Bloch band theory in arbitrary dimensions. Phys. Rev. X, 14, 021011(2024).

    [19] K. Yokomizo, S. Murakami. Non-Bloch band theory in bosonic Bogoliubov–de Gennes systems. Phys. Rev. B, 103, 165123(2021).

    [20] K. Ding, C. Fang, G. Ma. Non-Hermitian topology and exceptional-point geometries. Nat. Rev. Phys., 4, 745-760(2022).

    [21] K. Kawabata, K. Shiozaki, M. Ueda. Symmetry and topology in non-Hermitian physics. Phys. Rev. X, 9, 041015(2019).

    [22] A. Banerjee, A. Bandyopadhyay, R. Sarkar. Non-Hermitian topology and flat bands via an exact real-space decimation scheme. Phys. Rev. B, 110, 085431(2024).

    [23] H. Gao, H. Xue, Z. Gu. Non-Hermitian route to higher-order topology in an acoustic crystal. Nat. Commun., 12, 1888(2021).

    [24] H. Li, S. Wan. Dynamic skin effects in non-Hermitian systems. Phys. Rev. B, 106, L241112(2022).

    [25] W.-T. Xue, Y.-M. Hu, F. Song. Non-Hermitian edge burst. Phys. Rev. Lett., 128, 120401(2022).

    [26] Z. Li, L.-W. Wang, X. Wang. Observation of dynamic non-Hermitian skin effects. Nat. Commun., 15, 6544(2024).

    [27] L. Xiao, W.-T. Xue, F. Song. Observation of non-Hermitian edge burst in quantum dynamics. Phys. Rev. Lett., 133, 070801(2024).

    [28] K. Zhang, Z. Yang, C. Fang. Correspondence between winding numbers and skin modes in non-Hermitian systems. Phys. Rev. Lett., 125, 126402(2020).

    [29] Z. Yang, K. Zhang, C. Fang. Non-Hermitian bulk-boundary correspondence and auxiliary generalized Brillouin zone theory. Phys. Rev. Lett., 125, 226402(2020).

    [30] Y. Li, C. Liang, C. Wang. Gain-loss-induced hybrid skin-topological effect. Phys. Rev. Lett., 128, 223903(2022).

    [31] J. Sun, C.-A. Li, S. Feng. Hybrid higher-order skin-topological effect in hyperbolic lattices. Phys. Rev. B, 108, 075122(2023).

    [32] Y. Sun, X. Hou, T. Wan. Photonic Floquet skin-topological effect. Phys. Rev. Lett., 132, 063804(2024).

    [33] L.-W. Wang, Z.-K. Lin, J.-H. Jiang. Non-Hermitian topological phases and skin effects in kagome lattices. Phys. Rev. B, 108, 195126(2023).

    [34] E. Lustig, S. Weimann, Y. Plotnik. Photonic topological insulator in synthetic dimensions. Nature, 567, 356-360(2019).

    [35] M. Rechtsman, J. Zeuner, Y. Plotnik. Photonic Floquet topological insulators. Nature, 496, 196-200(2013).

    [36] K. Wintersperger, C. Braun, F. Ãoenal. Realization of an anomalous Floquet topological system with ultracold atoms. Nat. Phys., 16, 1058-1063(2020).

    [37] L. Zhou, J. Gong. Non-Hermitian Floquet topological phases with arbitrarily many real-quasienergy edge states. Phys. Rev. B, 98, 205417(2018).

    [38] B. Höckendorf, A. Alvermann, H. Fehske. Non-Hermitian boundary state engineering in anomalous Floquet topological insulators. Phys. Rev. Lett., 123, 190403(2019).

    [39] H. Wu, B.-Q. Wang, J.-H. An. Floquet second-order topological insulators in non-Hermitian systems. Phys. Rev. B, 103, L041115(2021).

    [40] H. Wu, J.-H. An. Floquet topological phases of non-Hermitian systems. Phys. Rev. B, 102, 041119(2020).

    [41] H. Wu, J.-H. An. Non-Hermitian Weyl semimetal and its Floquet engineering. Phys. Rev. B, 105, L121113(2022).

    [42] Z. Lin, W. Song, L.-W. Wang. Observation of topological transition in Floquet non-Hermitian skin effects in silicon photonics. Phys. Rev. Lett., 133, 073803(2024).

    [43] T. T. Koutserimpas, R. Fleury. Nonreciprocal gain in non-Hermitian time-Floquet systems. Phys. Rev. Lett., 120, 087401(2018).

    [44] H. Fan, H. Gao, T. Liu. Reconfigurable topological modes in acoustic non-Hermitian crystals. Phys. Rev. B, 107, L201108(2023).

    [45] S. Xia, D. Kaltsas, D. Song. Nonlinear tuning of PT symmetry and non-Hermitian topological states. Science, 372, 72-76(2021).

    [46] N. R. Cooper. Optical flux lattices for ultracold atomic gases. Phys. Rev. Lett., 106, 175301(2011).

    [47] J.-H. Jiang. Tunable topological Weyl semimetal from simple-cubic lattices with staggered fluxes. Phys. Rev. A, 85, 033640(2012).

    [48] Z. Ren, D. Liu, E. Zhao. Chiral control of quantum states in non-Hermitian spin-orbit-coupled fermions. Nat. Phys., 18, 385-389(2022).

    [49] W. Zhu, J. Gong. Photonic corner skin modes in non-Hermitian photonic crystals. Phys. Rev. B, 108, 035406(2023).

    Yi-Ling Zhang, Li-Wei Wang, Yang Liu, Zhao-Xian Chen, Jian-Hua Jiang, "Floquet hybrid skin-topological effects in checkerboard lattices with large Chern numbers," Photonics Res. 13, 1321 (2025)
    Download Citation