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

One-dimensional \mathbb{Z}-classified topological crystalline insulator under space-time inversion symmetry

Ling Lin1,2 Email: [email protected]    Yongguan Ke1,2 Email: [email protected]    Chaohong Lee1,2 Email: [email protected], [email protected] 1Institute of Quantum Precision Measurement, State Key Laboratory of Radio Frequency Heterogeneous Integration, College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China 2Quantum Science Center of Guangdong-Hong Kong-Macao Greater Bay Area (Guangdong), Shenzhen 518045, China
Abstract

We explore a large family of one-dimensional (1D) topological crystalline insulators (TCIs) classified by \mathbb{Z} invariants under space-time inversion symmetry. This finding stands in marked contrast to the conventional classification of 1D band topology protected by inversion symmetry and characterized by 2\mathbb{Z}_{2}-quantized polarization (Berry-Zak phase). By considering the nontrivial relative polarization among sublattices (orbitals), we introduce the inversion winding number as a topological invariant for characterizing and categorizing band topology. The inversion winding number reliably captures a novel bulk-edge correspondence, where gapped edge states are related to the inversion winding number of the sandwiching bands. Leveraging real-space analysis, we discover disorder-induced topological Anderson insulators and propose to experimentally discern band topology through relative polarization of edge states or bulk states. Our comprehensive findings present a paradigmatic illustration for the ongoing investigation and classification of band topology in TCIs.

Introduction. Crystalline symmetry plays an important role in classifying topological insulators [1, 2, 3, 4, 5, 6], complementary to the three intrinsic symmetries (i.e. chiral, time-reversal, and particle-hole symmetries). Tremendous topological states protected by crystalline symmetries, known as topological crystalline insulators (TCIs), have been explored [1, 2, 3, 4, 5, 6, 7]. In recent years, systematic search for TCIs has been achieved by analyzing high-symmetry points in momentum space and chemical orbitals in real space [2, 3]. TCIs exhibit nontrivial topological properties such as corner states [8, 9], quantized multipole moments [9, 10], non-Abelian topological charges [6, 11, 12, 13], and period-multiplied Bloch oscillations [14, 15]. These properties show that crystalline symmetries introduce richer topological structures, and may have promising applications in spintronics, quantum computing, and novel electronic devices.

In one dimension (1D), the inversion symmetry is essential for realizing TCI [16, 16, 17, 18, 19, 20, 21]. Given a gapped band, the polarization pp related to the Berry-Zak phase is widely used as a 2\mathbb{Z}_{2}-quantized invariant to characterize different 1D TCIs [22, 23, 24, 25, 20, 26], though it may fail to predict the bulk-edge correspondence [27]. For composite multiple bands, one can utilize the inversion eigenvalues at high-symmetry points or the Wilson loop to identify a non-negative integer invariant \mathbb{Z}^{\geq} [16, 17, 10, 28]. In the presence of chiral symmetry, it is well known that the band topology in 1D transcends 2\mathbb{Z}_{2} classification and falls into the \mathbb{Z} classification, which can be characterized by the chiral winding number [29, 30]. Notably, it has been discovered that the chiral winding number can be equivalently expressed as the polarization difference between the two sublattices (or orbitals) [31, 32, 33], offering a more fundamentally physical perspective on comprehending 1D chiral-symmetric topological insulators. The number of in-gap zero modes is proportional to the chiral winding number, meaning that nontrivial relative polarization between sublattices can give rise to nontrivial boundary states. Therefore, in the absence of chiral symmetry, a natural question arises: Can nontrivial relative inter-sublattice polarization exist as a result of crystalline symmetry and further categorize the band topology in 1D?

In this Letter, we present a theory demonstrating that, under space-time inversion symmetry, exotic topological structures can exist in a large family of 1D TCIs. Space-time inversion symmetry, referring to the combination of inversion symmetry and time-reversal symmetry, is expressed by

𝒯H(k)(𝒯)1=H(k),\displaystyle{\cal I\cal T}H\left(k\right){\left({{\cal I\cal T}}\right)^{-1}}=H\left(k\right), (1)

where kk is the quasi momentum, \mathcal{I} denotes the space-inversion operator and 𝒯{\cal T} represents the time-reversal operator. The space-time inversion symmetry imposes constraints on the Bloch Hamiltonian and its eigenvectors locally within the momentum space. This phenomenon brings internal topological structures to the Bloch states, enabling a substantial classification of band topology within 1D TCIs. Our work gives insights into systematic exploration of TCIs with hidden topological structures in higher dimensions.

Three-band model and inversion winding number. We consider a simple but general three-band model with space-time inversion symmetry (1). Its Hamiltonian matrix and inversion operator are given by

H(k)=(ΔkgkhkgkΔkgkhkgkΔk),=(001010100),H\left(k\right)=\left({\begin{array}[]{*{20}{c}}{{\Delta_{k}}}&{{g_{k}}}&{{h_{k}}}\\ {{g_{k}}^{*}}&{-{\Delta_{k}}}&{{g_{k}}}\\ {{h_{k}}^{*}}&{{g_{k}}^{*}}&{{\Delta_{k}}}\end{array}}\right),\;\;{\cal I}=\left({\begin{array}[]{*{20}{c}}{0}&{0}&1\\ {0}&1&{0}\\ 1&{0}&{0}\end{array}}\right), (2)

with gk,hkg_{k},h_{k}\in\mathbb{C} being complex parameters and Δk\Delta_{k}\in\mathbb{R} being real parameter. The time-reversal operation on the Bloch Hamiltonian here is the complex conjugation operation: 𝒯H(k)𝒯1=H(k){\cal T}H(k){\cal T}^{-1}={H(k)}^{*}. Clearly, this model has no chiral symmetry. Due to the space-time inversion symmetry, the eigenvector of H(k)H(k) can be written as a general form

𝒖k=(αkeiξkβkeiξk2αk)T{\bm{u}_{k}}={\left({\begin{array}[]{*{20}{c}}{{\alpha_{k}}{e^{-i{\xi_{k}}}}}&{{\beta_{k}}{e^{-i\frac{{{\xi_{k}}}}{2}}}}&{{\alpha_{k}}^{*}}\end{array}}\right)^{\rm{T}}} (3)

with ξk[0,2π)\xi_{k}\in[0,2\pi), αk\alpha_{k}\in\mathbb{C} and βk\beta_{k}\in\mathbb{R}. For simplicity, we have omitted the band index and assumed that all bands are gapped. Under the inversion transformation, we have (𝒖k)=eiξk𝒖k{\left({{\cal I}{\bm{u}_{k}}}\right)^{*}}={e^{i{\xi_{k}}}}{\bm{u}_{k}}. In particular, the first and third sublattices are related by a space-time inversion operation, and the phase factor associated with ξk\xi_{k} cannot be gauged out. This leads to the possible presence of nontrivial topological structures among the sublattices.

It is important to note that, in 1D chiral-symmetric topological insulators, the polarization difference between two sublattices can lead to nontrivial polarizations and charge accumulations at the edge [31, 33, 34]. This phenomenon is associated with the chiral winding number belonging to the \mathbb{Z} class. Thus, in the 1D TCIs with space-time inversion symmetry, we define a gauge-invariant winding number via the sublattice polarization difference between the first and the third sublattice,

ν=1πi02π𝑑k[uk|k(P^3P^1)ukuk|P^1+P^3|uk],\displaystyle\nu=\frac{1}{{\pi i}}\int_{0}^{2\pi}{dk\;\left[{\frac{{\langle{u_{k}}|\frac{\partial}{{\partial k}}\left({{{\hat{P}}_{3}}-{{\hat{P}}_{1}}}\right){u_{k}}\rangle}}{{\langle{u_{k}}|{{\hat{P}}_{1}}+{{\hat{P}}_{3}}|{u_{k}}\rangle}}}\right]}, (4)

where P^σ=k|k,σk,σ|{{\hat{P}}_{\sigma}}=\sum\nolimits_{k}{|k,\sigma\rangle\langle k,\sigma|} is the projector onto the σ=1,2,3\sigma=1,2,3 sublattice with |k,σ|k,\sigma\rangle being the basis of the momentum space, and uk|P^1+P^3|uk{\langle{u_{k}}|{{\hat{P}}_{1}}+{{\hat{P}}_{3}}|{u_{k}}\rangle} denotes the normalization factor. Due to the inversion symmetry, one can find uk|P^1|uk=uk|P^3|uk{\langle{u_{k}}|{{\hat{P}}_{1}}|{u_{k}}\rangle}={\langle{u_{k}}|{{\hat{P}}_{3}}|{u_{k}}\rangle}. This gauge-invariant quantity is quantized to integers ν\nu\in\mathbb{Z} and it is directly related to the quantized polarization such that p=ν/2mod1p=\nu/2\mod 1 [35]. It is worth noting that the inversion winding number shares some similarities with the mirror Chern number, where the difference of the Chern number between the two sectors having different mirror eigenvalues constitutes the mirror Chern number [36, 37]. In our study, the inversion winding number is obtained via projection onto the two inversion-symmetric sublattices.

Below we discuss the topological stability of the defined inversion winding number (4). It can be found that the inversion winding number becomes ill-defined when αk=0\alpha_{k}=0 in Eq. (3). To change the value of the inversion winding number, the system must pass this singular point. To find in what condition that αk=0\alpha_{k}=0, we solve the eigenvalue equation H(k)uk=EkukH(k)u_{k}=E_{k}u_{k} by assuming αk=0\alpha_{k}=0. We find that gk=0g_{k}=0 is necessary for αk=0\alpha_{k}=0. It can be concluded that, to change the inversion winding number, one must either break the space-time inversion symmetry, close the spectral gap, or introduce certain terms to ensure that gk=0g_{k}=0. Generally speaking, the topological protection in TCIs manifests that the target system cannot be adiabatically connected to other distinct TCI phases without closing the spectral gap and breaking the underlying symmetry. In this sense, the inversion winding number here seems unstable, as it may possibly change through tuning gk=0g_{k}=0. Nevertheless, the inversion winding number is related to the polarization via p=ν/2mod1p=\nu/2\mod 1. Strictly speaking, the parity of the inversion winding number is a 2\mathbb{Z}_{2} topological invariant. Although the inversion winding number becomes ill-defined when αk=0\alpha_{k}=0, it is still possible to introduce perturbations to ensure that gk0g_{k}\neq 0, thereby avoiding this singular point.

Fortunately, the condition gk0g_{k}\neq 0 for avoiding the singular point naturally arises in most realistic systems. As the parameter gkjjtj,jσ,σeik(jj){g_{k}}\propto\sum\nolimits_{j\leq j^{\prime}}{t_{j,j^{\prime}}^{\sigma,\sigma^{\prime}}{e^{ik\left({j-j^{\prime}}\right)}}} describes the coupling strength between σ={1,3}\sigma=\{1,3\} and σ=2\sigma^{\prime}=2 sublattices in momentum space, one can find that gk=0g_{k}=0 physically requires some very special tunneling terms in real space. Here, tj,jσ,σt_{j,j^{\prime}}^{\sigma,\sigma^{\prime}} represents the tunneling strength from the σ\sigma sublattice in the jjth cell to the σ\sigma^{\prime} sublattice in the jj^{\prime}th cell in real space. For example, if we set gk=0g_{k}=0 for all kB.Z.k\in\rm{B.Z.}, some certain lattices become completely isolated, resulting in the system being in the atomic limit. Therefore, we can assert that by excluding those specific conditions to ensure that gk0g_{k}\neq 0, the inversion winding number remains topologically stable. This argument remains valid for a large family of systems, offering a more comprehensive classification for space-time inversion-invariant systems. Later, we will provide some evidence to support this point.

Refer to caption
Figure 1: (a) Energy spectrum under the OBC as a function of η/t1\eta/t_{1}. The color of each line corresponds to the inverse participation ration of the eigenstate: IPR=j,σ|ψn|j,σ|4\mathrm{IPR}=\sum_{j,\sigma}|\langle\psi_{n}|j,\sigma\rangle|^{4}, which reflects the localization property of an eigenstate. Grey-dashed lines roughly indicate the gap-closing point between the first and second bands and divide into four regions (i-iv). (b1-b3) Inversion winding number for the three bands (n=1,2,3n=1,2,3) as functions of η/t1\eta/t_{1} under PBC. (c1-c3) Polarization of the three bands under PBC. Other parameters are chosen as t2/t1=1/2t_{2}/t_{1}=1/2.

Prototypical model and bulk-edge correspondence. To show the validity of the defined inversion winding number and demonstrate its \mathbb{Z} classification, we choose gk=t1g_{k}=-t_{1}, hk=t2eikηe2ikh_{k}=-t_{2}e^{-ik}-\eta e^{-2ik} and Δk=0\Delta_{k}=0, which corresponds to a system with long-range tunneling in real space. The real-space formulation of this model can be found in Supllementary Material. By varying η/t1\eta/t_{1}, the spectrum under open boundary condition (OBC) is shown in Fig. 1 (a), comprising three bands labeled as n=1,2,3n=1,2,3 from the bottom to the top. In both the first and second band gaps, several in-gap states appear. These in-gap states correspond to edge states that are exponentially localized near the two boundaries. To better elucidate the results, we concentrate on the lowest gap and designate four regions (i-iv) separated by gap-closing points and delve into its topological properties. Intriguingly, there are three different phases. We find that the in-gap edge states exhibit a four-fold degeneracy in regions (i) and (iv), and a two-fold degeneracy in region (ii). In region (iii), there is no in-gap edge state. This suggests that these three regions are three distinct phases.

Subsequently, we calculate the inversion winding number (4) under the periodic boundary condition (PBC), see Figs. 1 (b1-b3). Remarkably, the inversion winding number identifies more distinct phases and its variation coincides with the locations of gap-closing points. Under the OBC, the number of edge states closely relates to the inversion winding number of the adjacent bands. For instance, the numbers of in-gap edge states in the first and second gaps are respectively proportional to the inversion winding numbers of the first (n=1n=1) and third (n=3n=3) bands: Nedge#1=2|ν1|N_{{\rm{edge}}}^{\#1}=2|{\nu_{1}}| and Nedge#2=2|ν3|N_{{\rm{edge}}}^{\#2}=2|{\nu_{3}}|, where Nedge#mN_{{\rm{edge}}}^{\#m} and νn\nu_{n} respectively denote the number of edge states in the mmth gap and the inversion winding number of the nnth band. Regarding the second band (n=2n=2), its inversion winding number is the sum of the first and third bands: ν2=ν1+ν3\nu_{2}=\nu_{1}+\nu_{3}, and the change of this inversion winding number requires the closing of either the first or the second band gap. It can be further deduced that the nontrivial topology of the second band contributes to the emergence of in-gap edge states in both band gaps.

Refer to caption
Figure 2: Energy spectrum versus the real-valued parameter r[0,1]r\in[0,1]. Here, r=1r=1 corresponds to the PBC, while r=0r=0 corresponds to the OBC. The inversion winding number (νn\nu_{n}) of each band (n=1,2,3n=1,2,3) under PBC and the number of edge states (Nedge#mN_{\rm{edge}}^{\#m}) in each gap (m=1,2m=1,2) under OBC are explicitly labelled accordingly, which clearly manifests the bulk-edge correspondence.

To elucidate the bulk-edge correspondence more profoundly, we consider a continuous connection between the PBC and the OBC. This transition can be effectively modeled by introducing a real-valued parameter r[0,1]r\in[0,1] to the tunneling: tj,jσ,σrtj,jσ,σt_{j,j^{\prime}}^{\sigma,\sigma^{\prime}}\to rt_{j,j^{\prime}}^{\sigma,\sigma^{\prime}} when j,jj,j^{\prime} crosses the boundary under PBC. Thereby, r=1r=1 corresponds to the PBC, while r=0r=0 corresponds to the OBC, which allows the emergency of edge states. As depicted in Fig. 2 (a-b), edge states arise from the two adjacent bands, with their number related to the inversion winding number. The number of edge states arising from the second band is equal to the number of edge states arising from the other two bands. These findings not only reveal a more extensive classification of the band topology protected by the space-time inversion symmetry, but also well demonstrate the nature of bulk-edge correspondence for the inversion winding number. More discussions about the bulk-edge correspondence can be found in the Supplementary Material [35].

For comparison, we calculate the polarization for each band, as shown in Figs. 1 (c1-c3). Generally, the emergence of in-gap edge states is expected as a result of non-zero band polarization, which can be calculated using the Berry-phase formalism under PBC [38, 39]: pn=1/2πB.Z.𝑑kukn|iuknmod1p_{n}=1/2\pi\int_{{\rm{B}}.{\rm{Z}}.}{dk\;\langle{u^{n}_{k}}|i\partial{u^{n}_{k}}\rangle}\mod 1, with nn being the band index. The polarization provides insights into the center of Wannier function within a unit cell [40]. Owing to the space-time inversion symmetry, the polarization for each gapped band is quantized to either pn=0p_{n}=0 or pn=1/2p_{n}=1/2 [41], which has been widely employed to distinguish the band topology. However, as shown by the numerical results, the polarization is unable to explain the presence of the four-fold degenerate edge states observed above. Therefore, unlike the defined inversion winding number, the polarization fails to identify the rich band topology.

Robustness of inversion winding number. Having established the momentum-space formula of inversion winding number, we then delve into exploring its real-space counterpart to assess the robustness of band topology against disorder that preserves the space-time inversion symmetry. In analogy to the chiral-symmetric system [33, 42], the real-space formula of the inversion winding number can be expressed as a Bott index form,

ν=12πIm{Tr[log(𝒳1𝒳3)]},\nu=-\frac{1}{{2\pi}}{\mathop{\rm Im}\nolimits}\left\{{{\rm{Tr}}\left[{{\rm{log}}\left({{{\cal X}_{1}}^{\dagger}{{\cal X}_{3}}}\right)}\right]}\right\}, (5)

in which 𝒳σ=𝚿exp(i2πLx^)P^σ𝚿{{\cal X}_{\sigma}}={{\bf{\Psi}}^{\dagger}}\exp\left({i\frac{{2\pi}}{L}\hat{x}}\right){{\hat{P}}_{\sigma}}{\bf{\Psi}} is the projected-position matrix of the σ=1,3\sigma=1,3 sublattices and 𝚿=(,|ψj,){\bf{\Psi}}=\left({\cdots,|{\psi_{j}}\rangle,\cdots}\right) denotes a set of eigenstates on the target band. Eq. (5) is equivalent to the momentum-space formula in the presence of translation symmetry [35].

The real-space formula enables us to efficiently calculate the inversion winding number in the presence of inversion-invariant disorder. Here, we introduce disorders to the nearest-neighbor tunneling between the first and third sublattice: H^dis=Wjδx|x,1x,3|+H.c.{{\hat{H}}_{dis}}=-W\sum\nolimits_{j}{{\delta_{x}}|x,1\rangle\langle x,3|}+{\rm{H.c.}}, wherein δx[1/2,1/2]\delta_{x}\in[-1/2,1/2] represents a uniformly distributed random variable and WW is the strength of disorder. Here, |x,σ,x=1,2,,L|x,\sigma\rangle,x=1,2,\cdots,L denotes the position basis. To preserve the space-time inversion symmetry, we enforce ^H^dis^1=(H^dis)\hat{\mathcal{I}}{{\hat{H}}_{dis}}\hat{\mathcal{I}}^{-1}={\left({{{\hat{H}}_{dis}}}\right)^{*}}. Meanwhile, this term only entails the coupling between the sublattices σ=1\sigma=1 and σ=3\sigma=3. As mentioned earlier, the inversion winding number retains its well-defined nature as long as t10t_{1}\neq 0 in this system.

Refer to caption
Figure 3: Phase diagram of the lowest band in the presence of disorder: (a) the inversion winding number and (b) the polarization. Dashed lines indicate the phase boundary determined by the half-height between two phases. Number of cells is L=200L=200 and other parameters are the same as Fig. 1. All results are averaged over 10310^{3} random realizations.

Subsequently, by utilizing Eq. (5), we calculate the η\eta-WW phase diagram of inversion winding number for the lowest band under PBC, see Fig. 3 (a). Remarkably, the real-space formula yields a quantized value in every random realization. For weak disorder strength (W/t11W/t_{1}\ll 1), the first band gap remains finite and the fluctuations of the inversion winding number are nearly zero, as shown in Figs. 3 (c) and (d). As the disorder strength intensifies, we observe the emergence of a disorder-induced topological transition. Particularly, there is the disorder-induced topological phase (ν=0ν=1\nu=0\to\nu=1), which is identified as the topological Anderson insulating phase [43, 44, 45, 32, 46, 47, 48]. Significant statistical fluctuations in the inversion winding number are observed around all phase transitions. Away from the phase transition points, the inversion winding number remains remarkably stable in broad areas for different random realizations. More detailed discussions, including the evidence for bulk-edge correspondence, can be found in the Supplementary Material [35].

For comparison, we calculate the polarization through the projected position operator methods [49, 10, 50], as shown in Fig. 3 (b). Due to the inversion symmetry, the polarization is also quantized. The polarization pp is connected to the inversion winding number ν\nu through the relation p=ν/2mod1p=\nu/2\mod 1. As expected, the polarization fails to distinguish the phase of ν=0\nu=0 and 2.

Refer to caption
Figure 4: Simulation of experimentally measuring the inversion winding number. (a) Measuring through the density distribution of edge states in the first gap. (a1-a2) shows two examples of the edge states (only the left end is shown for simplicity). (b) Measuring through the time-evolution method with the total evolution time being T=30/t1T=30/t_{1}. The system’s size is chosen as L=200L=200. Other parameters are the same as Fig. 1.

Real-space characterization of band topology. In above, we have found that the inversion winding number is associated with the polarization difference between two sublattices. This offers the feasibility for experimentally characterizing different topological phases. Under the OBC, the inversion winding number can be approximately extracted from bulk states or edge states via

ν=(1)χ2Lntargetψn|X^3X^1|ψnψn|P^3+P^1|ψn\nu=(-1)^{\chi}\frac{2}{L}\sum\limits_{n\in{\rm{target}}}{\frac{{\langle{\psi_{n}}|{{\hat{X}}_{3}}-{{\hat{X}}_{1}}|{\psi_{n}}\rangle}}{{\langle{\psi_{n}}|{{\hat{P}}_{3}}+{{\hat{P}}_{1}}|{\psi_{n}}\rangle}}} (6)

in which X^σ=x^P^σ\hat{X}_{\sigma}=\hat{x}\hat{P}_{\sigma} is the position operator projected to the σ={1,2,3}\sigma=\{1,2,3\} sublattice. Eq. (6) is a summation over bulk states in a certain band (χ=1\chi=1) or edge states in a certain gap (χ=0\chi=0). It implies that one can measure the inversion winding number through the bulk or the edge states. In experiments, the edge states can be probed via versatile approaches and then the information of inversion winding number can be directly extracted. As for the bulk state, it is also feasible to measure the winding number through time-evolution [51, 32, 52, 53, 54, 55, 56]: νlimT1T0T𝑑τO(τ)\nu\approx\mathrm{lim}_{T\to\infty}\frac{1}{T}\int_{0}^{T}{d\tau\;O\left(\tau\right)} with O(τ)=Ψ(τ)|X^3X^1|Ψ(τ)Ψ(τ)|P^3+P^1|Ψ(τ)O\left(\tau\right)=\frac{{\langle\Psi\left(\tau\right)|{{\hat{X}}_{3}}-{{\hat{X}}_{1}}|\Psi\left(\tau\right)\rangle}}{{\langle\Psi\left(\tau\right)|{{\hat{P}}_{3}}+{{\hat{P}}_{1}}|\Psi\left(\tau\right)\rangle}} and |Ψ(τ)|\Psi\left(\tau\right)\rangle being a superposition of bulk states in a certain band. In Fig. 4, we present our numerical simulation for the experimental measurement through edge states or bulk states [35]. Remarkably, both methods produces a very prominent results compared with the theoretical prediction. The tiny difference between them can be attributed to the finite-size effect and the relatively small spectral gap, especially for bulk measurement. This again highlights the bulk-edge correspondence principle in this system: the appearance of in-gap edge state is a result of the nontrivial sublattice polarization of a band. The nontrivial inversion winding number has prominent physical consequences. The above results can be verified in various systems, such as, electrical circuits [57, 58], acoustic systems [59, 60] and other photonic or atomic systems [61, 26, 62, 63, 64, 64, 65, 21, 55, 11, 66].

Summary and discussion. We uncover that, under the space-time inversion symmetry, the band topology can be further classified beyond the conventional 2\mathbb{Z}_{2} classification in a large family of 1D three-band models. Based upon the sublattice polarization difference from the space-time inversion symmetry, we introduce a gauge-invariant winding number: the inversion winding number. The bulk topology and the bulk-edge correspondence can be well characterized by the inversion winding number ν\nu\in\mathbb{Z}. In addition, we give a real-space formula for the inversion winding number. We also find that the inversion winding number remains effective even in the presence of strong symmetry-invariant disorders. With the real-space formula for the inversion winding number, one can experimentally identify distinct topological phases in TCIs.

This work is supported by the National Key Research and Development Program of China (Grant No. 2022YFA1404104), the National Natural Science Foundation of China (Grant No. 12025509, Grant No. 12247134 and Grant No. 12275365), and the Natural Science Foundation of Guangdong Province (Grant No.2023A1515012099).

References

  • Fu [2011] L. Fu, Topological crystalline insulators, Phys. Rev. Lett. 106, 106802 (2011).
  • Bradlyn et al. [2017] B. Bradlyn, L. Elcoro, J. Cano, M. G. Vergniory, Z. Wang, C. Felser, M. I. Aroyo, and B. A. Bernevig, Topological quantum chemistry, Nature 547, 298 (2017).
  • Po et al. [2017] H. C. Po, A. Vishwanath, and H. Watanabe, Symmetry-based indicators of band topology in the 230 space groups, Nature Communications 8, 50 (2017).
  • Fu and Kane [2007] L. Fu and C. L. Kane, Topological insulators with inversion symmetry, Phys. Rev. B 76, 045302 (2007).
  • Cornfeld and Chapman [2019] E. Cornfeld and A. Chapman, Classification of crystalline topological insulators and superconductors with point group symmetries, Phys. Rev. B 99, 075105 (2019).
  • Wu et al. [2019] Q. Wu, A. A. Soluyanov, and T. Bzdušek, Non-Abelian band topology in noninteracting metals, Science 365, 1273 (2019).
  • Ando and Fu [2015] Y. Ando and L. Fu, Topological crystalline insulators and topological superconductors: From concepts to materials, Annual Review of Condensed Matter Physics 6, 361 (2015).
  • Liu and Wakabayashi [2017] F. Liu and K. Wakabayashi, Novel topological phase with a zero berry curvature, Phys. Rev. Lett. 118, 076803 (2017).
  • Benalcazar et al. [2017a] W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Quantized electric multipole insulators, Science 357, 61 (2017a).
  • Benalcazar et al. [2017b] 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 (2017b).
  • Guo et al. [2021] Q. Guo, T. Jiang, R.-Y. Zhang, L. Zhang, Z.-Q. Zhang, B. Yang, S. Zhang, and C. T. Chan, Experimental observation of non-Abelian topological charges and edge states, Nature 594, 195 (2021).
  • Li and Hu [2023] T. Li and H. Hu, Floquet non-Abelian topological insulator and multifold bulk-edge correspondence, Nature Communications 14, 6418 (2023).
  • Wang et al. [2024] Q.-D. Wang, Y.-Q. Zhu, S.-L. Zhu, and Z. Zheng, Synthetic non-Abelian topological charges in ultracold atomic gases, Phys. Rev. A 110, 023321 (2024).
  • Höller and Alexandradinata [2018] J. Höller and A. Alexandradinata, Topological bloch oscillations, Phys. Rev. B 98, 024310 (2018).
  • Di Liberto et al. [2020] M. Di Liberto, N. Goldman, and G. Palumbo, Non-abelian bloch oscillations in higher-order topological insulators, Nature communications 11, 5942 (2020).
  • Hughes et al. [2011] T. L. Hughes, E. Prodan, and B. A. Bernevig, Inversion-symmetric topological insulators, Phys. Rev. B 83, 245132 (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).
  • Shiozaki and Sato [2014] K. Shiozaki and M. Sato, Topology of crystalline insulators and superconductors, Phys. Rev. B 90, 165114 (2014).
  • Xiao et al. [2014] M. Xiao, Z. Q. Zhang, and C. T. Chan, Surface impedance and bulk band geometric phases in one-dimensional systems, Phys. Rev. X 4, 021017 (2014).
  • Hwang et al. [2019] Y. Hwang, J. Ahn, and B.-J. Yang, Fragile topology protected by inversion symmetry: Diagnosis, bulk-boundary correspondence, and wilson loop, Phys. Rev. B 100, 205126 (2019).
  • Vaidya et al. [2023] S. Vaidya, A. Ghorashi, T. Christensen, M. C. Rechtsman, and W. A. Benalcazar, Topological phases of photonic crystals under crystalline symmetries, Phys. Rev. B 108, 085116 (2023).
  • Zak [1980] J. Zak, Symmetry specification of bands in solids, Phys. Rev. Lett. 45, 1025 (1980).
  • Zak [1981] J. Zak, Band representations and symmetry types of bands in solids, Phys. Rev. B 23, 2824 (1981).
  • Zak [1982] J. Zak, Band center—a conserved quantity in solids, Phys. Rev. Lett. 48, 359 (1982).
  • Zak [1989] J. Zak, Berry’s phase for energy bands in solids, Phys. Rev. Lett. 62, 2747 (1989).
  • Song et al. [2018] B. Song, L. Zhang, C. He, T. F. J. Poon, E. Hajiyev, S. Zhang, X.-J. Liu, and G.-B. Jo, Observation of symmetry-protected topological band with ultracold fermions, Science Advances 4, eaao4748 (2018).
  • Rhim et al. [2018] J.-W. Rhim, J. H. Bardarson, and R.-J. Slager, Unified bulk-boundary correspondence for band insulators, Phys. Rev. B 97, 115143 (2018).
  • Takahashi et al. [2020] R. Takahashi, Y. Tanaka, and S. Murakami, Bulk-edge and bulk-hinge correspondence in inversion-symmetric insulators, Phys. Rev. Res. 2, 013300 (2020).
  • Chiu et al. [2016] C.-K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, Classification of topological quantum matter with symmetries, Rev. Mod. Phys. 88, 035005 (2016).
  • Ryu et al. [2010] S. Ryu, A. P. Schnyder, A. Furusaki, and A. W. W. Ludwig, Topological insulators and superconductors: tenfold way and dimensional hierarchy, New J. Phys. 12, 065010 (2010).
  • Mondragon-Shem et al. [2014] I. Mondragon-Shem, T. L. Hughes, J. Song, and E. Prodan, Topological Criticality in the Chiral-Symmetric AIII Class at Strong Disorder, Phys. Rev. Lett. 113, 046802 (2014).
  • Meier et al. [2018] E. J. Meier, F. A. An, A. Dauphin, M. Maffei, P. Massignan, T. L. Hughes, and B. Gadway, Observation of the topological Anderson insulator in disordered atomic wires, Science 362, 929 (2018).
  • Lin et al. [2021] L. Lin, Y. Ke, and C. Lee, Real-space representation of the winding number for a one-dimensional chiral-symmetric topological insulator, Phys. Rev. B 103, 224208 (2021).
  • Lin and Lee [2024] L. Lin and C. Lee, Probing chiral-symmetric higher-order topological insulators with multipole winding number (2024), arXiv:2401.03699 [cond-mat.mes-hall] .
  • [35] See Supplementary Material for details of: (S1) band structure and inversion winding number; (S2) derivation of the real-space formula; (S3) discussions on the bulk-edge correspondence; (S4) the quantization relations of topological invariants; (S5) discussion on the real-space model; (S6) discussion on the two- and four-band systems, which includes Refs. [xxx].
  • Teo et al. [2008] J. C. Y. Teo, L. Fu, and C. L. Kane, Surface states and topological invariants in three-dimensional topological insulators: Application to Bi1xSbx{\text{Bi}}_{1-x}{\text{Sb}}_{x}Phys. Rev. B 78, 045426 (2008).
  • Hsieh et al. [2012] T. H. Hsieh, H. Lin, J. Liu, W. Duan, A. Bansil, and L. Fu, Topological crystalline insulators in the SnTe material class, Nature Communications 3, 982 (2012).
  • King-Smith and Vanderbilt [1993] R. D. King-Smith and D. Vanderbilt, Theory of polarization of crystalline solids, Phys. Rev. B 47, 1651 (1993).
  • Resta [1994] R. Resta, Macroscopic polarization in crystalline dielectrics: the geometric phase approach, Rev. Mod. Phys. 66, 899 (1994).
  • Marzari et al. [2012] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Maximally localized Wannier functions: Theory and applications, Rev. Mod. Phys. 84, 1419 (2012).
  • Ahn et al. [2019] J. Ahn, S. Park, D. Kim, Y. Kim, and B.-J. Yang, Stiefel–whitney classes and topological phases in band theory, Chinese Physics B 28, 117101 (2019).
  • Benalcazar and Cerjan [2022] W. A. Benalcazar and A. Cerjan, Chiral-symmetric higher-order topological phases of matter, Phys. Rev. Lett. 128, 127601 (2022).
  • Li et al. [2009] J. Li, R.-L. Chu, J. K. Jain, and S.-Q. Shen, Topological Anderson Insulator, Phys. Rev. Lett. 102, 136806 (2009).
  • Altland et al. [2014] A. Altland, D. Bagrets, L. Fritz, A. Kamenev, and H. Schmiedt, Quantum Criticality of Quasi-One-Dimensional Topological Anderson Insulators, Phys. Rev. Lett. 112, 206602 (2014).
  • Wu et al. [2016] B. Wu, J. Song, J. Zhou, and H. Jiang, Disorder effects in topological states: Brief review of the recent developments*, Chinese Physics B 25, 117311 (2016).
  • Lin et al. [2022] Q. Lin, T. Li, L. Xiao, K. Wang, W. Yi, and P. Xue, Observation of non-Hermitian topological Anderson insulator in quantum dynamics, Nature Communications 13, 3229 (2022).
  • Tang et al. [2022] L.-Z. Tang, S.-N. Liu, G.-Q. Zhang, and D.-W. Zhang, Topological Anderson insulators with different bulk states in quasiperiodic chains, Phys. Rev. A 105, 063327 (2022).
  • Ren et al. [2024] M. Ren, Y. Yu, B. Wu, X. Qi, Y. Wang, X. Yao, J. Ren, Z. Guo, H. Jiang, H. Chen, X.-J. Liu, Z. Chen, and Y. Sun, Realization of Gapped and Ungapped Photonic Topological Anderson Insulators, Phys. Rev. Lett. 132, 066602 (2024).
  • Resta [1998] R. Resta, Quantum-mechanical position operator in extended systems, Phys. Rev. Lett. 80, 1800 (1998).
  • Lin et al. [2023] L. Lin, Y. Ke, and C. Lee, Topological invariants for interacting systems: From twisted boundary conditions to center-of-mass momentum, Phys. Rev. B 107, 125161 (2023).
  • Cardano et al. [2017] F. Cardano, A. D’Errico, A. Dauphin, M. Maffei, B. Piccirillo, C. de Lisio, G. De Filippis, V. Cataudella, E. Santamato, L. Marrucci, et al., Detection of zak phases and topological invariants in a chiral quantum walk of twisted photons, Nature communications 8, 15516 (2017).
  • Wang et al. [2018] X. Wang, L. Xiao, X. Qiu, K. Wang, W. Yi, and P. Xue, Detecting topological invariants and revealing topological phase transitions in discrete-time photonic quantum walks, Phys. Rev. A 98, 013835 (2018).
  • Xie et al. [2019] D. Xie, W. Gou, T. Xiao, B. Gadway, and B. Yan, Topological characterizations of an extended su–schrieffer–heeger model, npj Quantum Information 5, 55 (2019).
  • Wang et al. [2019a] Y. Wang, Y.-H. Lu, F. Mei, J. Gao, Z.-M. Li, H. Tang, S.-L. Zhu, S. Jia, and X.-M. Jin, Direct observation of topology from single-photon dynamics, Phys. Rev. Lett. 122, 193903 (2019a).
  • Cai et al. [2019] W. Cai, J. Han, F. Mei, Y. Xu, Y. Ma, X. Li, H. Wang, Y. P. Song, Z.-Y. Xue, Z.-q. Yin, S. Jia, and L. Sun, Observation of topological magnon insulator states in a superconducting circuit, Phys. Rev. Lett. 123, 080501 (2019).
  • Xie et al. [2020] D. Xie, T.-S. Deng, T. Xiao, W. Gou, T. Chen, W. Yi, and B. Yan, Topological quantum walks in momentum space with a bose-einstein condensate, Phys. Rev. Lett. 124, 050502 (2020).
  • Olekhno et al. [2022] N. A. Olekhno, A. D. Rozenblit, V. I. Kachin, A. A. Dmitriev, O. I. Burmistrov, P. S. Seregin, D. V. Zhirihin, and M. A. Gorlach, Experimental realization of topological corner states in long-range-coupled electrical circuits, Phys. Rev. B 105, L081107 (2022).
  • Li et al. [2023a] Y. Li, J.-H. Zhang, F. Mei, B. Xie, M.-H. Lu, J. Ma, L. Xiao, and S. Jia, Large-chiral-number corner modes in \mathbb{Z}-class higher-order topolectrical circuits, Phys. Rev. Appl. 20, 064042 (2023a).
  • Wang et al. [2023] D. Wang, Y. Deng, J. Ji, M. Oudich, W. A. Benalcazar, G. Ma, and Y. Jing, Realization of a \mathbb{Z}-classified chiral-symmetric higher-order topological insulator in a coupling-inverted acoustic crystal, Phys. Rev. Lett. 131, 157201 (2023).
  • Li et al. [2023b] Y. Li, H. Qiu, Q. Zhang, and C. Qiu, Acoustic higher-order topological insulators protected by multipole chiral numbers, Phys. Rev. B 108, 205135 (2023b).
  • Poddubny et al. [2014] A. Poddubny, A. Miroshnichenko, A. Slobozhanyuk, and Y. Kivshar, Topological majorana states in zigzag chains of plasmonic nanoparticles, ACS Photonics 1, 101 (2014).
  • Zhang et al. [2018] D.-W. Zhang, Y.-Q. Zhu, Y. X. Zhao, H. Yan, and S.-L. Zhu, Topological quantum matter with cold atoms, Advances in Physics 67, 253 (2018).
  • Kruk et al. [2019] S. Kruk, A. Poddubny, D. Smirnova, L. Wang, A. Slobozhanyuk, A. Shorokhov, I. Kravchenko, B. Luther-Davies, and Y. Kivshar, Nonlinear light generation in topological nanostructures, Nature Nanotechnology 14, 126 (2019).
  • Ozawa et al. [2019] T. Ozawa, H. M. Price, A. Amo, N. Goldman, M. Hafezi, L. Lu, M. C. Rechtsman, D. Schuster, J. Simon, O. Zilberberg, and I. Carusotto, Topological photonics, Rev. Mod. Phys. 91, 015006 (2019).
  • Wang et al. [2019b] H.-X. Wang, G.-Y. Guo, and J.-H. Jiang, Band topology in classical waves: Wilson-loop approach to topological numbers and fragile topology, New Journal of Physics 21, 093029 (2019b).
  • Jiao et al. [2021] Z.-Q. Jiao, S. Longhi, X.-W. Wang, J. Gao, W.-H. Zhou, Y. Wang, Y.-X. Fu, L. Wang, R.-J. Ren, L.-F. Qiao, and X.-M. Jin, Experimentally detecting quantized zak phases without chiral symmetry in photonic lattices, Phys. Rev. Lett. 127, 147401 (2021).