This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Second-order topological insulator and fragile topology in topological circuitry simulation

Ce Shang [email protected] King Abdullah University of Science and Technology (KAUST), Physical Science and Engineering Division (PSE), Thuwal 23955-6900, Saudi Arabia.    Xiaoning Zang King Abdullah University of Science and Technology (KAUST), Physical Science and Engineering Division (PSE), Thuwal 23955-6900, Saudi Arabia.    Wenlong Gao Paderborn University, Department of Physics, Warburger Str. 100, 33098 Paderborn, Germany    Udo Schwingenschlögl King Abdullah University of Science and Technology (KAUST), Physical Science and Engineering Division (PSE), Thuwal 23955-6900, Saudi Arabia.    Aurélien Manchon [email protected] [email protected] CINaM, Aix-Marseille University, CNRS, Marseille, France. King Abdullah University of Science and Technology (KAUST), Physical Science and Engineering Division (PSE), Thuwal 23955-6900, Saudi Arabia. King Abdullah University of Science and Technology (KAUST), Computer, Electrical and Mathematical Science and Engineering Division (CEMSE), Thuwal 23955-6900, Saudi Arabia.
Abstract

Second-order topological insulators (SOTIs) are the topological phases of matter in dd dimensions that manifest (d2)(d-2)-dimensional localized modes at the intersection of the edges. We show that SOTIs can be designed via stacked Chern insulators with opposite chiralities connected by interlayer coupling. To characterize the bulk-corner correspondence, we establish a Jacobian-transformed nested Wilson loop method and an edge theory that are applicable to a wider class of higher-order topological systems. The corresponding topological invariant admits a filling anomaly of the corner modes with fractional charges. The system manifests a fragile topological phase characterized by the absence of a Wannier gap in the Wilson loop spectrum. Furthermore, we argue that the proposed approach can be generalized to multilayers. Our work offers perspectives for exploring and understanding higher-order topological phenomena.

preprint: APS/123-QEDpreprint: APS/123-QED

Introduction.—Identifying and classifying topological phases of matter plays an important role in modern physics. Bulk-boundary correspondence and emergent particles with quantized charge reflect the most fundamental property of a topologically protected system [1]. A topological insulator behaves as an insulator in its interior (bulk) in dd dimensions, and (d1)(d-1)-dimensional metallic surface states reside in the gap of the bulk states [2, 3]. Forming a novel class of topological insulators, higher-order topological insulators (HOTIs) support gapless states in a subspace of a dimension lower than d1d-1 [4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14]. In two-dimensional (2D) space, a second-order topological insulator (SOTI) combines two essential ingredients: (i) fractional corner charges in the presence of certain symmetries; (ii) topologically robust zero-dimensional (0D) corner states as adiabatic deformations of the boundary with preserved symmetry [10, 11].

While wavefunctions and spectra may be similar or even identical, the topological classification can be evaluated by integration of the Berry curvature over a closed manifold [15, 16] or by the winding of the Wilson loop (holonomy) [17, 18, 19]. To characterize the boundary topology, a Wannier Hamiltonian is defined as the product of U(2N){{U}}(2N) Berry connection along a noncontractible Wilson loop in the Brillouin zone (BZ), and its eigenvalues are encoded with the U(1){{U}}(1) phase of the Wannier center in an effective one-dimensional (1D) form [17]. Crucially, the Wannier centers of occupied bands refer to Wannier bands carrying a topological invariant determining the bulk-corner correspondence. These can be investigated by a nested Wilson loop formalism [7, 20, 21, 22, 23].

HOTIs have been simulated in a plethora of platforms, including solid state [24], photonics [25], and acoustics systems [26, 27, 28]. In the present work, we propose to employ a bilayer topological circuitry, a remarkably flexible and tunable platform [29, 30], to realize HOTIs. The layer degree of freedom splits the space of wavefunctions into a finite number of independent copies. This constitutes a direct product relationship of the spinors with the ‘layer’ and ‘sublattice’ parts, and the top and bottom states can be coupled by the interlayer hopping.

In this Letter, we put forward a theoretical design of SOTIs by stacking Chern insulators with opposite chiralities coupled by interlayer hopping. We start from a bilayer Haldane model [31] with the time-reversal symmetry 𝒯\mathcal{T} and opposite Chern numbers [32, 33] (1D Dirac-type gapless helical edge states). By introducing a 𝒯\mathcal{T}-breaking interlayer hopping, the edge modes become gapped and 0D corner modes are obtained. However, the corner modes are protected by the product of 𝒯\mathcal{T} and the C2C_{2} rotation symmetry, leading to a fragile topological phase [34, 22]. To characterize the bulk-corner correspondence, we establish a Jacobian-transformed nested Wilson loop method by performing a Jacobian transformation of the Cartesian coordinates that reparametrizes the BZ. The corresponding topological invariant admits a filling anomaly of the corner modes with fractional charges |e|/2|e|/2 [35]. We utilize the edge theory to explain the underlying physics in an intuitive manner. We also claim that the stacked Chern insulator approach can be generalized to multilayers, where only an even number of layers supports corner modes. The gap of the 1D band decays exponentially as a function of the number of layers due to finite size effects [36]. Our work offers new perspectives for exploring and understanding higher-order topological phenomena.

Model.— As illustrated in Fig. 1, we investigate a honeycomb lattice layout realized by an electrical circuit [37, 38], specifically a bilayer-bipartite lattice with AA and BB sites in both of top (α\alpha) and bottom (β\beta) layers. The corresponding tight-binding Hamiltonian can be written as

H=ijt1ciτ0cj+ijit2vijciτzcj+itciσxci,H=\sum\limits_{\left\langle{ij}\right\rangle}{{t_{1}}c_{i}^{\dagger}{\tau_{0}}{c_{j}}}+\sum\limits_{\left\langle{\left\langle{ij}\right\rangle}\right\rangle}{i{t_{2}}{v_{ij}}c_{i}^{\dagger}{\tau_{z}}}{c_{j}}+\sum\limits_{i}{{t_{\bot}}c_{i}^{\dagger}{\sigma_{x}}}{c_{i}}, (1)

where ci=(ci,α,ci,β)Tc_{i}^{\dagger}={({c_{i,\alpha}^{\dagger},c_{i,\beta}^{\dagger}})^{T}} is the two-component creation operator acting at the site ii of the α\alpha and β\beta layers. The Pauli matrices σ(τ)x,σ(τ)y,σ(τ)z\sigma(\tau)_{x},\sigma(\tau)_{y},\sigma(\tau)_{z} and the identity matrix σ(τ)0\sigma(\tau)_{0} refer to the layer (sublattice) degree of freedom. t1t_{1} denotes the intralayer hopping between the nearest neighbor sites. t2t_{2} emulates the spin-orbit coupling between next nearest neighbor sites, where vij=di×dj/|di×dj|{v_{ij}}={\textbf{d}_{i}}\times{\textbf{d}_{j}}/|{\textbf{d}_{i}}\times{\textbf{d}_{j}}| depends on the orientations of the mediating bonds, where di{\textbf{d}_{i}} and dj{\textbf{d}_{j}} are unit vectors connecting site jj and ii. tt_{\bot} denotes the interlayer hopping between site ii in different layers.

Refer to caption
Refer to caption
(a)(b)
Figure 1: (a) Schematic representation of an electrical circuit consisting of capacitors representing the interlayer hopping t1t_{1} (black) and interlayer hopping tt_{\bot} (red), and operational amplifiers representing the spin-orbit interaction t2t_{2} (blue/green) [38]. (b) The energy spectrum EE along the high symmetry path for t1=1t_{1}=1, t2=0.1t_{2}=0.1 and t=0.2t_{\bot}=0.2.

By Fourier transformation of the Hamiltonian (1) into momentum space, we obtain the Bloch matrix

(k)=d1(k)Γ1+d2(k)Γ2+d3(k)Γ3tΓ34,{\cal{H}}(\textbf{k})={d_{1}}(\textbf{k}){\Gamma_{1}}+{d_{2}}(\textbf{k}){\Gamma_{2}}+{d_{3}}(\textbf{k}){\Gamma_{3}}-{t_{\bot}}{\Gamma_{34}}, (2)

with

d1(k)=t1[1+2cos(3aky/2)cos(3akx/2)],d2(k)=2t1sin(3aky/2)cos(3akx/2),d3(k)=t2[4cos(3aky/2)sin(3aky/2)2sin(3akx)],\begin{split}{d_{1}}(\textbf{k})=&{t_{1}}[1+2\cos(3a{k_{y}}/2)\cos(\sqrt{3}a{k_{x}}/2)],\\ {d_{2}}(\textbf{k})=&2{t_{1}}\sin(3a{k_{y}}/2)\cos(\sqrt{3}a{k_{x}}/2),\\ {d_{3}}(\textbf{k})=&{t_{2}}[4\cos(3a{k_{y}}/2)\sin(\sqrt{3}a{k_{y}}/2)-2\sin(\sqrt{3}a{k_{x}})],\end{split} (3)

where Γ(1,2,3,4,5)=(σ0τx,σ0τy,σzτz,σyτz,σxτz){\Gamma_{(1,2,3,4,5)}}=({\sigma_{0}}\otimes{\tau_{x}},{\sigma_{0}}\otimes{\tau_{y}},{\sigma_{z}}\otimes{\tau_{z}},{\sigma_{y}}\otimes{\tau_{z}},{\sigma_{x}}\otimes{\tau_{z}}) and Γab=[Γa,Γb]/2i\Gamma_{ab}=[{\Gamma_{a}},{\Gamma_{b}}]/2i are the Clifford-algebra matrices. Due to common pseudospin 1/21/2 representation in momentum space, we can regard the BZ to be that of a 2D hexagonal lattice with spin. By diagonalizing Eq. (2), we obtain the band structure shown in Fig. 1. Dimensionless units are used throughout this Letter.

Topological invariant.—Emergence of topological corner states is the physical manifestation of a SOTI, with the topological property determined by the second Stiefel-Whitney number w2w_{2} [20, 21]. w2w_{2} is derived from the parity of the occupied Bloch wave functions at the time-reversal invariant momenta [23] 𝚪i{𝐌1,𝐌2,𝐌3,𝚪}{{{\bf{\Gamma}}}_{i}}\in\{{{\bf{M}}_{1}},{{\bf{M}}_{2}},{{\bf{M}}_{3}},{\bf{\Gamma}}\} [Fig. 2] as

(1)w2=i=14(1)[Nocc(𝚪𝐢)/2],{(-1)^{{w_{2}}}}=\prod\limits_{i=1}^{4}{{{(-1)}^{[N_{\rm{occ}}^{-}({\bf{\Gamma}_{i}})/2]}}}, (4)

where [Nocc(𝚪i)/2][N_{\rm{occ}}^{-}({{\bf{\Gamma}}_{i}})/2] counts the number of occupied bands with odd parity.

Refer to caption
Refer to caption
Refer to caption
(a)(b)(c)
Figure 2: (a) Mapping from a honeycomb BZ to a square BZ and its corresponding with time-reversal invariant momenta, labled 𝚪\bf{\Gamma}^{\prime} and 𝐌\bf{M}^{\prime}. 𝒲x(y),𝐤{\cal{W}}_{x^{\prime}(y^{\prime}),{\bf{k}}^{\prime}} is the Wilson loop along kx(y)k_{x^{\prime}(y^{\prime})} starting at 𝐤=[kx,ky]{\bf{k}^{\prime}}=[k_{x^{\prime}},k_{y^{\prime}}]. Wannier bands (b) vyv_{y^{\prime}} and (c) vxv_{x^{\prime}} constructed by Wilson loop diagonalization.

w2w_{2} also can be determined by nested Wilson loop calculation [7]. We introduce a Jacobian-transformed nested Wilson loop to conveniently confirm gauge-invariance. The primitive vectors of the honeycomb lattice in real and momentum space do not point along the (x,y)(x,y) and (kx,ky)(k_{x},k_{y}), respectively, as shown in Fig. 2, because the hexagonal primitive cell is topologically equivalent to a rectangular primitive cell [39]. Hence, the transition matrix =diag[3/2,3/2]{\mathcal{M}}=\rm{diag}[\sqrt{3}/2,3/2] performs the Jacobian transformation [kx,ky]=[kx,ky][{k_{x^{\prime}}},{k_{y^{\prime}}}]={\mathcal{M}}[{k_{x}},{k_{y}}], mapping the honeycomb BZ to a T2=[0,2π]×[0,2π]{T^{2}}=[0,2\pi]\times[0,2\pi] torus. The corresponding real space coordinates read [x,y]=1[x,y][x^{\prime},y^{\prime}]={{\mathcal{M}}^{-1}}[x,y]. Two manifolds with the same Euler characteristic can be continuously deformed into each other. Thus, the Wilson loop is redefined in the standardized square primitive cell as a path-ordered product of the exponential of Berry connections,

𝒲(kx+2π,ky)(kx,ky)=limNFN1FN2F1F0,{{\cal{W}}_{({k_{x^{\prime}}}+2\pi,{k_{y^{\prime}}})\leftarrow({k_{x^{\prime}}},{k_{y^{\prime}}})}}=\mathop{\lim}\limits_{N\to\infty}{F_{N-1}}{F_{N-2}}\cdots{F_{1}}{F_{0}}, (5)

where [F]mn=um(2π(+1)N,ky)|un(2πN,ky){[{F_{\ell}}]_{mn}}=\langle{u_{m}}(\frac{{2\pi(\ell+1)}}{N},{k_{y^{\prime}}}){\rm{|}}{u_{n}}(\frac{{2\pi\ell}}{N},{k_{y^{\prime}}})\rangle, |un\left|u_{n}\right\rangle and |um\left|u_{m}\right\rangle denote occupied Bloch functions, NN is the number of discrete points along the kxk_{x^{\prime}} direction, and m,n=1,,Noccm,n=1,\cdots,N_{\rm{occ}}. The Wilson loop eigenvalues indicates Wannier centers eivx(ky){e^{i{v_{x^{\prime}}}({k_{y^{\prime}}})}} at given kyk_{y^{\prime}}. According to Figs. 2 and 2, the Wilson loop spectra have no Wannier gap and belong to the orthogonal group SO(2)SO(2) [34, 40], with the homotopy group π1[SO(2)]={\pi_{1}}[SO(2)]=\mathbb{Z} guaranteeing the existence of a Stiefel-Whitney class invariant. This situation is identified as a fragile topological phase [41]. The Stiefel-Whitney number can be derived from the nested Wilson loop

𝒲~(ky+2π)(ky)=limNF~N1F~N2F~1F~0,{\tilde{\cal W}_{({k_{y^{\prime}}}+2\pi)\leftarrow({k_{y^{\prime}}})}}=\mathop{\lim}\limits_{N\to\infty}{\tilde{F}_{N-1}}{\tilde{F}_{N-2}}\cdots{\tilde{F}_{1}}{\tilde{F}_{0}}, (6)

where [F~]mn=vx,m(2π(+1)N)|vx,n(2πN){[{{\tilde{F}}_{\ell}}]_{mn}}=\langle{v_{x^{\prime},m}}(\frac{{2\pi(\ell+1)}}{N}){\rm{|}}{v_{x^{\prime},n}}(\frac{{2\pi\ell}}{N})\rangle. The determinant of the nested Wilson loop operator is det(W~)=(1)w2\det(\tilde{W})={(-1)^{{w_{2}}}}, and we can assert a nontrivial topological phase.

Edge spectrum and topological corner states.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
(a)(b)(c)(d)
Figure 3: Band spectra in a quasi-1D bilayer strip consisting of 10 unit cells (a) without and (b) with interlayer hopping (t=0.2t_{\bot}=0.2), where, t1=1t_{1}=1 and t2=0.1t_{2}=0.1. (c/d) Wavefunction for (a/b). The insets show the corresponding edge spectra.

We now investigate the edge spectrum and the physical properties of topological corner states. In a zigzag terminated ribbon with open boundary condition, without tt_{\bot}, a pair of helical gapless edge modes forms a Dirac-like band crossing projected by 𝒯\mathcal{T} symmetry, see Fig. 3. Figure 3 shows for a hexagonal-shaped flake with open boundary condition that the wavefuntion is localized at the edges, and the spectrum given as inset is quasicontinuous as expected for quantum spin Hall edge states. With tt_{\bot}, the 𝒯\mathcal{T} symmetry is broken and the edge modes become gapped, see Fig. 3. Due to the hexagonal geometry, the spectrum given as inset in Fig. 3 shows six zero modes. The spatial distribution of the wavefunction is evaluated by the local density of states ρ(𝐫)=π1Im[i0+𝕀H(𝐫)]1\rho({\bf{r}})=-{\pi^{-1}}{{\rm{Im}}[i{0^{+}}{\mathbb{I}}-H({\bf{r}})]^{-1}} at zero energy, where 0+0^{+} is a small broadening. When gapped edge modes encounter one another at the corners, the C6C_{6} symmetry enforces 6-fold degenerate localized corner states [Fig. 3]. The fractional charges in the six symmetric sectors are given by Q6=e2w2Q_{6}=\frac{e}{2}{w_{2}} [35].

Edge theory.— In order to gain intuitive understanding of SOTI phase, we construct the low-energy edge theory. The effective Hamiltonian is obtained by the lowest order expansion in the t=0t_{\bot}=0 limit with respect to the high symmetry points K:

bulk=vFkxΓ1+vFkyΓ2+mHΓ3,{{\cal H}_{bulk}}=\hbar{v_{F}}{k_{x}}{\Gamma_{1}}+\hbar{v_{F}}{k_{y}}{\Gamma_{2}}+m_{\rm{H}}{\Gamma_{3}}, (7)

where vF=3t1a/2v_{F}=3t_{1}a/2 is the Fermi velocity and mHm_{\rm{H}} is the Haldane mass. Upon replacement kx,yix,y{k_{x,y}}\to-i{\partial_{x,y}}, the real-space Dirac Hamiltonian reads

Hbulk=ivFxΓ1ivFyΓ2+mHΓ3{{H}_{bulk}}=-i\hbar{v_{F}}{\partial_{x}}{\Gamma_{1}}-i\hbar{v_{F}}{\partial_{y}}{\Gamma_{2}}+m_{\rm{H}}{\Gamma_{3}} (8)

in the 2D semi-space x(,)x\in({{-}}\infty,\infty) and y(,0]y\in({{-}}\infty,0]. Manifestly, we have 𝒯Hbulk𝒯1=Hbulk\mathcal{T}{H}_{bulk}{\mathcal{T}^{-1}}={H}_{bulk}. Solving the Dirac equation HbulkΨ(𝐫)=EΨ(𝐫){H}_{bulk}\Psi({\bf{r}})=E\Psi({\bf{r}}), the spinor Ψ(𝐫)\Psi({\bf{r}}) must satisfy a boundary condition that ensures a vanishing normal component of the particle current density,

Jy(𝐫)=Ψ(𝐫)|[y,Hbulk]|Ψ(𝐫)/i=0.{J_{y}}({\bf{r}})=\langle{\Psi}({\bf{r}})|[y,{H_{bulk}}]|{\Psi({\bf{r}})}\rangle/i\hbar=0. (9)

A suitable boundary condition is

Ψ(x,y=0)=Ψ(x,y=0),=Γ23,\Psi(x,y=0)={\mathcal{B}}\Psi(x,y=0),\quad{\mathcal{B}}=\Gamma_{23}, (10)

where 𝒯𝒯1=\mathcal{T}{\mathcal{B}\mathcal{T}^{-1}}={\mathcal{B}} and the 𝒯\mathcal{T} symmetry is conserved at the edges. Solutions with Bloch wave propagation along the edge with wave vector kxk_{x} and decaying exponentially into the bulk with length scale λ\lambda seeking in the form Ψkx(𝐫)=Φeikxx+y/λ{\Psi_{{k_{x}}}}({\bf{r}})=\Phi{e^{i{k_{x}}x+y/\lambda}}. With the constraint Γ2Φ=iΓ3Φ\Gamma_{2}\Phi=-i\Gamma_{3}\Phi resulting from the boundary condition, Eq. (10), we have

vFkxΓ1Φ+(mHvF/λ)Γ3Φ=EΦ.\hbar{v_{F}}k_{x}{\Gamma_{1}}\Phi{{+}}(m_{\rm{H}}-\hbar{v_{F}}/\lambda){\Gamma_{3}}\Phi{{=}}E\Phi. (11)

Setting the decay length λ=vF/mH\lambda=\hbar{v_{F}}/{m_{\rm{H}}}, one can solve Eq. (11) and obtains a set of bound states Φα\Phi_{\alpha} [42]. To turn on tt_{\bot}, a mass term mIm_{\rm{I}} is added to the Hamiltonian. Thus, in the basis Φα\Phi_{\alpha}, we have the 1D effective Hamiltonian

eff=vFkxτz+mIτx{{\cal{H}}_{\rm{eff}}}=\hbar{v_{F}}{k_{x}}{\tau_{z}}+m_{\rm{I}}{\tau_{x}} (12)

with edge energy spectrum E±=±(vFkx)2+mI2{E_{\pm}}=\pm\sqrt{{{\left({\hbar{v_{F}}{k_{x}}}\right)}^{2}}+m_{\rm{I}}^{2}}. Therefore, a 0D corner state must appear at the intersection of two edges with masses of opposite sign, corresponding to the Jackiw-Rebbi topological domain wall mode [43].

Generalization to multilayers. The proposed approach can be generalized to multilayers of Chern insulators, as illustrated in Fig. 4. The single-particle Hamiltonian is formed as direct sum of the Hamiltonians of the individual layers in the Hilbert space m=L1L2Lm{\mathscr{H}_{m}}={\mathscr{H}_{L1}}\oplus{\mathscr{H}_{L2}}\oplus\cdots\oplus{\mathscr{H}_{Lm}}. For a state-independent interlayer coupling, we have

H=mCmCmLm+mnCmCnt𝕀,H=\sum\limits_{m}{C_{m}^{\dagger}{C_{m}}\otimes{\mathscr{H}_{Lm}}}+\sum\limits_{\left\langle{mn}\right\rangle}{C_{m}^{\dagger}{C_{n}}\otimes{t_{\bot}}\mathbb{I}}, (13)

where CmC_{m}^{\dagger} is the creation operator acting on the whole layer mm. We stress that only an even number of helical modes (layers) can be used for construction of a SOTI. For an odd number of layers, there will always be a single helical mode corresponding to a strong topological phase, which cannot be gapped by a local perturbation as illustrated in Fig. 4.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
(a)(b)(c)(d)
Figure 4: (a) Multilayer of Chern insulators with opposite chiralities connected by the interlayer hopping t=0.2t_{\bot}=0.2, where t1=1t_{1}=1 and t2=0.1t_{2}=0.1. (b) Band spectrum of a trilayer strip with 5 unit cells. (c) Dispersion of edge states. (d) Energy gap Δ\Delta as a function of the number of layers.

For an even number of layers, Fig. 4 shows the lowest valence and highest conduction dispersion of edge states for m=2m=2 (blue) and m=40m=40 (red). The energy gap Δ\Delta between the edge states [Fig. 4] decays exponentially for increasing mm due to finite size effects [36]. For mm\to\infty, the gap Δ\Delta becomes tiny and the edge states are no longer well separated. This means scattering between the edge and corner states becomes possible. When the disorder or interaction scattering are introduced, the single-particle elastic scattering between the edge and corner states is no longer forbidden, implying that the corner states lose their topological protection.

Summary.—We propose and analyze a scheme to create SOTIs by stacking Chern insulators with opposite chiralities subject to interlayer coupling. To characterize the bulk-corner correspondence and quantify the topological invariant, we establish a Jacobian-transformed nested Wilson loop method. We demonstrate a fragile topological phase by the absence of the Wannier gap in the Wilson loop spectrum. The topological invariant is found to admit a filling anomaly of the corner modes with fractional charges |e|/2|e|/2. Finally, we show that the introduced approach can be generalized to multilayers. Our proposal can be readily realized by topological circuitry, and developing new bulk materials or van der Waals heterostructures calls for a thrilling challenge.

The authors acknowledge financial support from the King Abdullah University of Science and Technology (KAUST).

References

  • Hasan and Kane [2010] M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).
  • Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • Bansil et al. [2016] A. Bansil, H. Lin, and T. Das, Colloquium: Topological band theory, Rev. Mod. Phys. 88, 021004 (2016).
  • Benalcazar et al. [2017a] W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Electric multipole moments, topological multipole moment pumping, and chiral hinge states in crystalline insulators, Phys. Rev. B 96, 245115 (2017a).
  • Song et al. [2017] Z. Song, Z. Fang, and C. Fang, (d2)(d-2)-dimensional edge states of rotation symmetry protected topological states, Phys. Rev. Lett. 119, 246402 (2017).
  • Langbehn et al. [2017] J. Langbehn, Y. Peng, L. Trifunovic, F. von Oppen, and P. W. Brouwer, Reflection-symmetric second-order topological insulators and superconductors, Phys. Rev. Lett. 119, 246401 (2017).
  • Benalcazar et al. [2017b] W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Quantized electric multipole insulators, Science 357, 61 (2017b).
  • Schindler et al. [2018a] F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. P. Parkin, B. A. Bernevig, and T. Neupert, Higher-order topological insulators, Sci. Adv. 4, eaat0346 (2018a).
  • Ezawa [2018] M. Ezawa, Magnetic second-order topological insulators and semimetals, Phys. Rev. B 97, 155305 (2018).
  • Sheng et al. [2019] X.-L. Sheng, C. Chen, H. Liu, Z. Chen, Z.-M. Yu, Y. X. Zhao, and S. A. Yang, Two-dimensional second-order topological insulator in Graphdiyne, Phys. Rev. Lett. 123, 256402 (2019).
  • Park et al. [2019] M. J. Park, Y. Kim, G. Y. Cho, and S. Lee, Higher-order topological insulator in twisted bilayer Graphene, Phys. Rev. Lett. 123, 216803 (2019).
  • Chen et al. [2020] R. Chen, C.-Z. Chen, J.-H. Gao, B. Zhou, and D.-H. Xu, Higher-order topological insulators in quasicrystals, Phys. Rev. Lett. 124, 036803 (2020).
  • Huang and Liu [2020] B. Huang and W. V. Liu, Floquet higher-order topological insulators with anomalous dynamical polarization, Phys. Rev. Lett. 124, 216601 (2020).
  • Ren et al. [2020] Y. Ren, Z. Qiao, and Q. Niu, Engineering corner states from two-dimensional topological insulators, Phys. Rev. Lett. 124, 166804 (2020).
  • Thouless et al. [1982] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Quantized Hall conductance in a two-dimensional periodic potential, Phys. Rev. Lett. 49, 405 (1982).
  • Xiao et al. [2010] D. Xiao, M.-C. Chang, and Q. Niu, Berry phase effects on electronic properties, Rev. Mod. Phys. 82, 1959 (2010).
  • Yu et al. [2011] R. Yu, X. L. Qi, A. Bernevig, Z. Fang, and X. Dai, Equivalent expression of Z2\rm{Z}_{2} topological invariant for band insulators using the non-Abelian Berry connection, Phys. Rev. B 84, 075119 (2011).
  • Fidkowski et al. [2011] L. Fidkowski, T. S. Jackson, and I. Klich, Model characterization of gapless edge modes of topological insulators using intermediate Brillouin-zone functions, Phys. Rev. Lett. 107, 036601 (2011).
  • Alexandradinata et al. [2014] A. Alexandradinata, X. Dai, and B. A. Bernevig, Wilson-loop characterization of inversion-symmetric topological insulators, Phys. Rev. B 89, 155114 (2014).
  • Ahn et al. [2018] J. Ahn, D. Kim, Y. Kim, and B.-J. Yang, Band topology and linking structure of nodal line semimetals with Z2{Z}_{2} monopole charges, Phys. Rev. Lett. 121, 106403 (2018).
  • Wang et al. [2019] Z. Wang, B. J. Wieder, J. Li, B. Yan, and B. A. Bernevig, Higher-order topology, monopole nodal lines, and the origin of large fermi arcs in transition metal dichalcogenides XTe2\rm{X}\rm{T}\rm{e}_{2} (X=Mo,W\rm{X}=\rm{Mo},\rm{W}), Phys. Rev. Lett. 123, 186401 (2019).
  • Wieder et al. [2020] B. J. Wieder, Z. Wang, J. Cano, X. Dai, L. M. Schoop, B. Bradlyn, and B. A. Bernevig, Strong and fragile topological Dirac semimetals with higher-order Fermi arcs, Nat. Commun. 11, 627 (2020).
  • Lee et al. [2020] E. Lee, R. Kim, J. Ahn, and B.-J. Yang, Two-dimensional higher-order topology in monolayer Graphdiyne, npj Quantum Materials 5, 1 (2020).
  • Schindler et al. [2018b] F. Schindler, Z. Wang, M. G. Vergniory, A. M. Cook, A. Murani, S. Sengupta, A. Y. Kasumov, R. Deblock, S. Jeon, I. Drozdov, H. Bouchiat, S. Guéron, A. Yazdani, B. A. Bernevig, and T. Neupert, Higher-order topology in Bismuth, Nat. Phys. 14, 918 (2018b).
  • Mittal et al. [2019] S. Mittal, V. V. Orre, G. Zhu, M. A. Gorlach, A. Poddubny, and M. Hafezi, Photonic quadrupole topological phases, Nat. Photon. 13, 692 (2019).
  • Xue et al. [2018] H. Xue, Y. Yang, F. Gao, Y. Chong, and B. Zhang, Acoustic higher-order topological insulator on a Kagome lattice, Nat. Mat. 18, 108 (2018).
  • Xue et al. [2019] H. Xue, Y. Yang, G. Liu, F. Gao, Y. Chong, and B. Zhang, Realization of an acoustic third-order topological insulator, Phys. Rev. Lett. 122, 244301 (2019).
  • Qi et al. [2020] Y. Qi, C. Qiu, M. Xiao, H. He, M. Ke, and Z. Liu, Acoustic realization of quadrupole topological insulators, Phys. Rev. Lett. 124, 206601 (2020).
  • Lee et al. [2018] C. H. Lee, S. Imhof, C. Berger, F. Bayer, J. Brehm, L. W. Molenkamp, T. Kiessling, and R. Thomale, Topolectrical circuits, Commun. Phys. 1, 39 (2018).
  • Imhof et al. [2018] S. Imhof, C. Berger, F. Bayer, J. Brehm, L. W. Molenkamp, T. Kiessling, F. Schindler, C. H. Lee, M. Greiter, T. Neupert, and R. Thomale, Topolectrical-circuit realization of topological corner modes, Nat. Phys. 14, 925 (2018).
  • Haldane [1988] F. D. M. Haldane, Model for a quantum Hall effect without landau levels: Condensed-matter realization of the ”parity anomaly”, Phys. Rev. Lett. 61, 2015 (1988).
  • Kane and Mele [2005a] C. L. Kane and E. J. Mele, Quantum spin Hall effect in graphene, Phys. Rev. Lett. 95, 226801 (2005a).
  • Kane and Mele [2005b] C. L. Kane and E. J. Mele, Z2{Z}_{2} topological order and the quantum spin Hall effect, Phys. Rev. Lett. 95, 146802 (2005b).
  • Bzdušek and Sigrist [2017] T. c. v. Bzdušek and M. Sigrist, Robust doubly charged nodal lines and nodal surfaces in centrosymmetric systems, Phys. Rev. B 96, 155105 (2017).
  • Benalcazar et al. [2019] W. A. Benalcazar, T. Li, and T. L. Hughes, Quantization of fractional corner charge in Cn{C}_{n}-symmetric higher-order topological crystalline insulators, Phys. Rev. B 99, 245151 (2019).
  • Zhou et al. [2008] B. Zhou, H.-Z. Lu, R.-L. Chu, S.-Q. Shen, and Q. Niu, Finite size effects on helical edge states in a quantum spin-Hall system, Phys. Rev. Lett. 101, 246807 (2008).
  • Ezawa [2019] M. Ezawa, Electric circuits for non-Hermitian Chern insulators, Phys. Rev. B 100, 081401 (2019).
  • Hofmann et al. [2019] T. Hofmann, T. Helbig, C. H. Lee, M. Greiter, and R. Thomale, Chiral voltage propagation and calibration in a topolectrical Chern circuit, Phys. Rev. Lett. 122, 247702 (2019).
  • Vanderbilt and King-Smith [1993] D. Vanderbilt and R. D. King-Smith, Electric polarization as a bulk quantity and its relation to surface charge, Phys. Rev. B 48, 4442 (1993).
  • Kooi et al. [2019] S. H. Kooi, G. van Miert, and C. Ortix, Classification of crystalline insulators without symmetry indicators: Atomic and fragile topological phases in twofold rotation symmetric systems, Phys. Rev. B 100, 115160 (2019).
  • Po et al. [2018] H. C. Po, H. Watanabe, and A. Vishwanath, Fragile topology and Wannier obstructions, Phys. Rev. Lett. 121, 126402 (2018).
  • Yan [2019] Z. Yan, Majorana corner and hinge modes in second-order topological insulator/superconductor heterostructures, Phys. Rev. B 100, 205406 (2019).
  • Jackiw and Rebbi [1976] R. Jackiw and C. Rebbi, Solitons with fermion number, Phys. Rev. D 13, 3398 (1976).