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

thanks: These authors contribute equally to this work.thanks: These authors contribute equally to this work.

Higher-order Topological Anderson Insulators

Yan-Bin Yang1    Kai Li1    L.-M. Duan1    Yong Xu1,2 yongxuphy@tsinghua.edu.cn 1Center for Quantum Information, IIIS, Tsinghua University, Beijing 100084, People’s Republic of China 2Shanghai Qi Zhi Institute, Shanghai 200030, People’s Republic of China
Abstract

We study disorder effects in a two-dimensional system with chiral symmetry and find that disorder can induce a quadrupole topological insulating phase (a higher-order topological phase with quadrupole moments) from a topologically trivial phase. Their topological properties manifest in a topological invariant defined based on effective boundary Hamiltonians, the quadrupole moment and zero-energy corner modes. We find gapped and gapless topological phases and a Griffiths regime. In the gapless topological phase, all the states are localized, while in the Griffiths regime, the states at zero energy become multifractal. We further apply the self-consistent Born approximation to show that the induced topological phase arises from disorder renormalized masses. We finally introduce a practical experimental scheme with topolectrical circuits where the predicted topological phenomena can be observed by impedance measurements. Our work opens the door to studying higher-order topological Anderson insulators and their localization properties.

I Introduction

Traditional topological phases usually feature the bulk-boundary correspondence that (n1)(n-1)-dimensional gapless boundary states exist for an nn-dimensional topological system. Recently, topological phases have been generalized to the case where there exist (nm)(n-m)-dimensional (instead of n1n-1) gapless boundary states with 1<mn1<m\leq n for an nn-dimensional system Taylor2017Science ; Fritz2012PRL ; ZhangFan2013PRL . In the past few years, the higher-order topological phenomena have drawn tremendous attention, and various higher-order topological states have been discovered Taylor2017Science ; Fritz2012PRL ; ZhangFan2013PRL ; Slager2015PRB ; FangChen2017PRL ; Brouwer2017PRL ; Bernevig2018SciAdv ; Brouwer2019PRX ; Roy2019PRB ; SYang2019PRL ; Vincent2020PRL ; Haiping2020PRL ; Yanbin2020PRR ; Tiwari2020PRL , such as quadrupole topological phases with zero-energy corner modes Taylor2017Science and its type-II cousin Yanbin2020PRR and second-order topological insulators with chiral hinge modes Bernevig2018SciAdv . It has also been shown that higher-order topological insulators (HOTIs) are robust against weak disorder Hatsugai2019PRB ; Jiang2019CPB ; Franca2019PRB ; CALi2020PRB ; Agarwala2020PRR ; XRWang2020 .

Disorder plays an important role in quantum transport, such as Anderson localization and metal-insulator transitions Evers2008RMP . In the context of first-order topological phases, it has been shown that they are usually stable against weak symmetry preserving disorder. But disorder is not always detrimental to first-order topological phases. Ref. Sheng2009PRL theoretically predicted that disorder can drive a topological phase transition from a metallic trivial phase to a quantum spin Hall insulator; topological insulators induced by disorder are called topological Anderson insulators (TAIs) Sheng2009PRL ; Beenakker2009PRL . Since their discovery, there has been great interest and advancement in the study of TAIs Jiang2009PRB ; Franz2010PRL ; Altland2014PRL ; Prodan2014PRL ; Rafael2015PRL ; SZhang2017PRL . In addition, disorder can drive a transition from a Weyl semimetal to a 3D quantum anomalous Hall state Xie2015PRL . Remarkably, the TAI has been experimentally observed in a photonic waveguide array Szameit2018Nat and disordered cold atomic wire Gadway2018Science .

Disorder, topology and symmetry are closely connected, which can be seen from classification theories. For example, random matrix theories are classified based on three internal symmetries, explaining universal transport properties of disordered physical systems Altland1997PRB ; Beenakker1997RMP ; HaakeBook . Similarly, the classification of topological phases is made according to these internal symmetries Ludwig2010NJP . Among these symmetries, chiral symmetry plays an important role in disordered systems and many peculiar properties have been found, such as the divergence of density of states (DOS) and localization length at energy E=0E=0 Dyson ; Cohen1976PRB ; Eggarter1978PRB ; Eilmes1998EPJB ; Brouwer1998PRL . In 2D, first-order topological phases are not allowed in a system with only chiral symmetry. Yet, it has been reported that a second-order topological phase can exist in a 2D system with chiral symmetry Okugawa2019PRB ; Qibo2020PRB and thus provides an ideal platform to study the interplay between disorder and topology.

Here we study the interplay between disorder and higher-order topology in a 2D system with chiral symmetry. We prove that the quantization of the quadrupole moment is maintained by chiral symmetry irrespective of crystalline symmetries, indicating that the quadrupole topological insulator can exist in a system with chiral symmetry without the requirement of any crystalline symmetry. This also gives us an opportunity to explore the effects of off-diagonal disorder respecting chiral symmetry. We theoretically predict the existence of a disorder induced HOTI [dubbed higher-order topological Anderson insulator (HOTAI)] with zero-energy corner modes, which arises through the localization-delocalization-localization phase transition. We further apply the self-consistent Born approximation (SCBA) to show that the induced phase appears due to the disorder renormalized masses. Besides, we find gapped and gapless HOTAIs and a Griffiths regime. In the gapless regime, all the states are localized, while in the Griffiths regime, the states at zero energy become multifractal. In addition, we study the disorder effects on a HOTI and show the existence of gapped and gapless topological phases and a Griffiths regime. Finally, we propose an experimental scheme using topolectrical circuits to realize and detect the HOTAI.

II Model Hamiltonian

We start by considering the following higher-order Hamiltonian

H^=𝐫[c^𝐫h0c^𝐫+(c^𝐫hxc^𝐫+𝐞x+c^𝐫hyc^𝐫+𝐞y+H.c.)],\hat{H}=\sum_{\bf r}[\hat{c}^{\dagger}_{\bf r}h_{0}\hat{c}_{\bf r}+(\hat{c}^{\dagger}_{\bf r}h_{x}\hat{c}_{{\bf r}+{\bf e}_{x}}+\hat{c}^{\dagger}_{\bf r}h_{y}\hat{c}_{{\bf r}+{\bf e}_{y}}+H.c.)], (1)

where c^𝐫=(c^𝐫1c^𝐫2c^𝐫3c^𝐫4)\hat{c}^{\dagger}_{\bf r}=\left(\begin{array}[]{cccc}\hat{c}^{\dagger}_{{\bf r}1}&\hat{c}^{\dagger}_{{\bf r}2}&\hat{c}^{\dagger}_{{\bf r}3}&\hat{c}^{\dagger}_{{\bf r}4}\\ \end{array}\right) with c^𝐫ν\hat{c}^{\dagger}_{{\bf r}\nu} (c^𝐫ν\hat{c}_{{\bf r}\nu}) being a creation (annihilation) operator at the ν\nuth site in a unit cell described by 𝐫=(x,y){\bf r}=(x,y) with xx and yy being integers (suppose that the lattice constants are equal to one), 𝐞x=(1,0){\bf e}_{x}=(1,0) and 𝐞y=(0,1){\bf e}_{y}=(0,1). Here

h0=(0im𝐫yim𝐫x0im𝐫y00im¯𝐫xim𝐫x00im¯𝐫y0im¯𝐫xim¯𝐫y0)h_{0}=\left(\begin{array}[]{cccc}0&-im_{{\bf r}}^{y}&-im_{{\bf r}}^{x}&0\\ im_{{\bf r}}^{y}&0&0&i\bar{m}_{{\bf r}}^{x}\\ im_{{\bf r}}^{x}&0&0&-i\bar{m}_{{\bf r}}^{y}\\ 0&-i\bar{m}_{{\bf r}}^{x}&i\bar{m}_{{\bf r}}^{y}&0\\ \end{array}\right) (2)

depicts the intra-cell hopping, and

hx=(00000000t𝐫x0000t¯𝐫x00)andhy=(0000t𝐫y000000000t¯𝐫y0)\displaystyle h_{x}=\left(\begin{array}[]{cccc}0&0&0&0\\ 0&0&0&0\\ t_{{\bf r}}^{x}&0&0&0\\ 0&-\bar{t}_{{\bf r}}^{x}&0&0\\ \end{array}\right)\text{and}~{}h_{y}=\left(\begin{array}[]{cccc}0&0&0&0\\ t_{{\bf r}}^{y}&0&0&0\\ 0&0&0&0\\ 0&0&\bar{t}_{{\bf r}}^{y}&0\\ \end{array}\right) (3)

describe the inter-cell hopping along xx and yy, respectively [also see Fig. 1(a) for the hopping parameters]. The system parameters m𝐫xm_{{\bf r}}^{x}, m¯𝐫x\bar{m}_{{\bf r}}^{x}, m𝐫ym_{{\bf r}}^{y}, m¯𝐫y\bar{m}_{{\bf r}}^{y}, t𝐫xt_{{\bf r}}^{x}, t¯𝐫x\bar{t}_{{\bf r}}^{x}, t𝐫yt_{{\bf r}}^{y} and t¯𝐫y\bar{t}_{{\bf r}}^{y} all take real values. For simplicity without loss of generality, we take the inter-cell hopping magnitude as energy units so that t𝐫x=t¯𝐫x=t𝐫y=t¯𝐫y=1t_{\bf r}^{x}=\bar{t}_{\bf r}^{x}=t_{\bf r}^{y}=\bar{t}_{\bf r}^{y}=1. In this case, hx=σσzh_{x}=\sigma_{-}\otimes\sigma_{z} and hy=σ0σh_{y}=\sigma_{0}\otimes\sigma_{-} with σ=[0 0;1 0]\sigma_{-}=[0\,0;1\,0] and σ0\sigma_{0} being a 2×22\times 2 identity matrix. For a clean system with m𝐫x=m¯𝐫x=m𝐫y=m¯𝐫y=mm_{\bf r}^{x}=\bar{m}_{\bf r}^{x}=m_{\bf r}^{y}=\bar{m}_{\bf r}^{y}=m, the system respects a generalized C4C_{4} symmetry as detailed in Appendix A.

To show that the Hamiltonian (1) describes a higher-order phase supporting zero-energy corner modes in the clean case with m𝐫x=m¯𝐫x=mxm_{\bf r}^{x}=\bar{m}_{\bf r}^{x}=m_{x} and m𝐫y=m¯𝐫y=mym_{\bf r}^{y}=\bar{m}_{\bf r}^{y}=m_{y}, we write the Hamiltonian in momentum space as

H^=𝐤c^𝐤H0(𝐤)c^𝐤.\hat{H}=\sum_{\bf k}\hat{c}_{\bf k}^{\dagger}H_{0}({\bf k})\hat{c}_{\bf k}. (4)

Here

H0(𝐤)=Hx(kx,mx)σz+σ0Hy(ky,my),H_{0}({\bf k})=H_{x}(k_{x},m_{x})\otimes\sigma_{z}+\sigma_{0}\otimes H_{y}(k_{y},m_{y}), (5)

where Hν(kν,mν)=coskνσx+(mν+sinkν)σyH_{\nu}(k_{\nu},m_{\nu})=\cos k_{\nu}\sigma_{x}+(m_{\nu}+\sin k_{\nu})\sigma_{y} (ν=x,y)(\nu=x,y) with σν\sigma_{\nu} (ν=x,y,z)(\nu=x,y,z) being the Pauli matrices and σ0\sigma_{0} being a 2×22\times 2 identity matrix. To see the presence of zero-energy corner modes in the system, we recast the Hamiltonian (5) to a form in continuous real space by replacing sinkν\sin k_{\nu} by iν-i\partial_{\nu} and coskν\cos k_{\nu} by 1+ν2/21+\partial_{\nu}^{2}/2 (ν=x,y\nu=x,y) so that Hν(kν)H¯νH_{\nu}(k_{\nu})\rightarrow\bar{H}_{\nu}. Considering semi-infinite boundaries along xx and yy, if |ux|u_{x}\rangle and |uy|u_{y}\rangle are zero-energy edge modes of H¯x\bar{H}_{x} and H¯y\bar{H}_{y}, respectively, |ux|uy|u_{x}\rangle\otimes|u_{y}\rangle is a zero-energy mode of H¯0\bar{H}_{0} localized at a corner.

Since the system contains only the nearest-neighbor hopping, it respects chiral symmetry, i.e., ΠHΠ1=H{\Pi}H{\Pi}^{-1}=-H, where HH is the first-quantization Hamiltonian and Π{\Pi} is a unitary matrix. But this system breaks the time-reversal symmetry and thus the particle-hole symmetry, because h0h_{0} is complex. In contrast, if we generalize the Benalcazar-Bernevig-Hughes (BBH) model Taylor2017Science to the disordered case, it still respects the time-reversal, particle-hole and chiral symmetries. However, these two models are connected through a local transformation and thus have similar topological and localization properties as proved in Appendix B. The equivalence also tells us that our system supports zero-energy corner modes and has quantized quadrupole moments Cho2018arXiv ; Wheeler2018arXiv protected by reflection symmetries. But with disorder breaking the reflection symmetry, one may wonder whether the quadrupole moment is still quantized. Here we prove the quantization of the quadrupole moment maintained by chiral symmetry (see Appendix C), indicating that chiral symmetry can protect a quadrupole topological insulator. We remark that in three dimensions (3D) chiral symmetry maintains the quantization of the octupole moment as proved in Appendix C, indicating that chiral symmetry can protect the third-order topological insulator with zero-energy corner modes in 3D.

Refer to caption
Figure 1: (Color online) (a) Schematics of our model (1). (b) The phase diagram with respect to the disorder strength WW mapped out based on the topological invariant PP, the quadrupole moments qxyq_{xy}, and the bulk energy gap shown in Fig. 2(a). We observe distinct phases including gapped/gapless HOTAI, Griffiths phase, and trivial I/II phases, separated by vertical dashed lines. Here mx=my=1.1m_{x}=m_{y}=1.1.

To study the disorder effects, we consider the disorder in the intra-cell hopping, that is, m𝐫ν=mν+WνV𝐫νm_{\bf r}^{\nu}=m_{\nu}+W^{\nu}V_{\bf r}^{\nu} and m¯𝐫ν=mν+W¯νV¯𝐫ν\bar{m}_{\bf r}^{\nu}={m}_{\nu}+\bar{W}^{\nu}\bar{V}_{\bf r}^{\nu} with ν=x,y\nu=x,y, where V𝐫ν{V}_{\bf r}^{\nu} and V¯𝐫ν\bar{V}_{\bf r}^{\nu} are uniformly randomly distributed in [0.5,0.5][-0.5,0.5] without correlation. Here Wν{W}^{\nu} and W¯ν\bar{W}^{\nu} represent the disorder strength. For simplicity, we take Wx=W¯x=Wy=W¯y=W{W}^{x}=\bar{W}^{x}={W}^{y}=\bar{W}^{y}=W. Because of the random character, we perform the average over 200-2000 sample configurations for numerical calculation.

Refer to caption
Figure 2: (Color online) (a) The bulk energy gap versus WW. (b) The quadrupole moment (red circles) versus the system size LL when W=2.6W=2.6, which is fitted by a power-law function plotted as the black line. The zero energy LDOS obtained under open boundary conditions for (c) W=2.6W=2.6 and (d) W=3.4W=3.4. The zoomed-in view of LDOS around one corner is shown in the insets. Here mx=my=1.1m_{x}=m_{y}=1.1.

III Higher-order topological Anderson insulators

We map out the phase diagram in Fig. 1(b), showing remarkably the presence of a disorder-induced higher-order topological phase transition. To characterize the phase transition, we evaluate the polarization pxp_{x} (pyp_{y}) of effective boundary Hamiltonians at a yy-normal (xx-normal) boundary at half filling. In one dimension, the polarization is equivalent to the Berry phase in a translation invariant system, which can be used as a topological invariant Resta . In fact, the polarization as a topological invariant can be evaluated in real space for a system without translational symmetries based on Resta’s formula Resta ; Prodan2014PRL . For the quadrupole topological phase, we define a topological invariant based on pxp_{x} and pyp_{y} as

P=4|pxpy|.P=4|p_{x}p_{y}|. (6)

When P=1P=1, the system is in a higher-order topologically nontrivial phase, and when P=0P=0, it is in a trivial phase (see Appendix D for its justification for a clean system).

We now generalize it to the disordered case. Specifically, we evaluate the average polarization of the effective boundary Hamiltonian at the yy-normal boundary (similarly for xx-normal one) by px=1Nin=1Ni|px,n|p_{x}=\frac{1}{N_{i}}\sum_{n=1}^{N_{i}}|p_{x,n}|, where px,n=ImlogΨn|e2πix^/Lx|Ψn/(2π)p_{x,n}=\text{Im}\text{log}\langle\Psi_{n}|e^{2\pi i\hat{x}/L_{x}}|\Psi_{n}\rangle/(2\pi) Resta with x^=xxn^x\hat{x}=\sum_{x}x\hat{n}_{x}, n^x\hat{n}_{x} being the particle number operator at the site xx, and LxL_{x} being the length of the system along xx (we also deduct the atomic positive charge contribution). Here |Ψn|\Psi_{n}\rangle is the ground state at half filling of the boundary Hamiltonian Hn=G2n(E=0)1H_{n}=-G_{2n}(E=0)^{-1} with G2nG_{2n} being the 2n2nth boundary Green’s function obtained by Dai2008PRB ; Oppen2017PRB

Gn=(EhnVn1Gn1Vn1)1,G_{n}=(E-h_{n}-V_{n-1}G_{n-1}V_{n-1}^{\dagger})^{-1}, (7)

where hnh_{n} is the Hamiltonian for the nnth layer and Vn1V_{n-1} is the coupling between the (n1)(n-1)th and nnth layer. We note that px,np_{x,n} is quantized to be either 0 or 0.50.5 for each iteration since HnH_{n} also preserves chiral symmetry. The polarization is evaluated at even steps of Green’s function given that there are two different layers in the clean limit. In the disordered case, the intra-cell hopping parts in hnh_{n} and VnV_{n} are randomly generated for each iteration (see Appendix D). The topological invariant PP is finally determined.

In Fig. 1(b), we plot the topological invariant PP as the disorder magnitude increases. We see that PP suddenly jumps to 11 when W2.1W\approx 2.1, indicating the occurrence of a topological phase transition. PP remains quantized to be 11 until W>3.5W>3.5, where it begins decreasing continuously. This regime corresponds to the Griffiths phase where topologically trivial and nontrivial sample configurations coexist (see Appendix E). When W>6W>6, PP vanishes, showing that the system reenters into a trivial phase.

To further identify that the induced topological phase is a quadrupole topological phase, we calculate the quadrupole moment, which can be used as a topological invariant since its quantization is protected by chiral symmetry (see Appendix C). Figure 1(b) shows that the quadrupole moment qualitatively agrees with the results of PP. Yet, conspicuous discrepancy can be observed. The quadrupole moment qxyq_{xy} over many samples is not quantized to 0.50.5 in the regime where P=1P=1 [qxy=0.5q_{xy}=0.5 for most disorder configurations and qxy=0q_{xy}=0 for other configurations (see Fig. C1 in Appendix C)] and the Griffiths regime is much larger. We attribute this to the finite-size effects, given that for the quadrupole moment, we can only perform a computation for a system with its size up to 8080, while to determine PP, we consider a system with its size up to 500500 and iterations up to 10310^{3}. To be more quantitative, we plot 0.5qxy0.5-q_{xy} as the system size LL increases when W=2.6W=2.6 in Fig. 2(b), showing a power law decay and thus suggesting that qxyq_{xy} approaches 0.50.5 in the thermodynamic limit.

Refer to caption
Figure 3: (Color online) (a) The normalized localization length Λx\Lambda_{x} at E=0E=0 versus WW for several distinct LyL_{y}. The scaling of Λx\Lambda_{x} at different energies for (b) W=3.2W=3.2 and (c) W=4.6W=4.6. (d) The LSR and IPR versus WW for the eigenstates around zero energy in a system with size L=500L=500. (e) The fractal dimension D2D_{2} with respect to WW for the eigenstates around zero energy. The vertical dashed lines separate different phases. (f) The topological invariant PP in the (W,m0)(W,m_{0}) plane with m0=mx=mym_{0}=m_{x}=m_{y}. The red dotted line indicates the topological phase boundary determined by the SCBA. In (a-e), mx=my=1.1m_{x}=m_{y}=1.1.

The higher-order topological phase transition occurs as the bulk energy gap closes at W2.1W\approx 2.1 and reopens, as shown in Fig. 2(a). In fact, the transition is associated with the divergence of the localization length at W2.1W\approx 2.1 [see Fig. 3(a)]. When WW is further increased, the energy gap closes again and remains closed due to the strong disorder scattering, leading to the gapless HOTAI. Even in the gapless regime, the topological invariant PP can still be quantized as shown in Fig. 1(b). In fact, in this phase, all the states are localized corresponding to an Anderson insulator (see the following discussion).

To further confirm that the TAI is a higher-order topological state, in Fig. 2(c-d), we display the local density of states (LDOS) at E=0E=0 for two typical values of WW corresponding to a gapped and gapless topological phase, respectively, clearly showing the presence of zero-energy states localized at corners. The evidence above definitely suggests the existence of HOTAIs.

IV Localization properties

Refer to caption
Figure 4: (Color online) The LSR versus energy EE for (a) W=1W=1, (b) W=2.8W=2.8, (c) W=3.2W=3.2 and (d) W=4.6W=4.6 in the trivial-I, gapped HOTAI, gapless HOTAI and Griffiths phases, respectively. The blue, red and yellow lines describe results for the system size Lx=Ly=L=300,400,500L_{x}=L_{y}=L=300,400,500, respectively. The inset in (a) displays the LSR as a function of EE for W=1.5W=1.5. The inset in (b) shows the normalized localization length Λx\Lambda_{x} with respect to LyL_{y} at different energies. Here mx=my=1.1m_{x}=m_{y}=1.1.

We now study the localization properties of energy bands in different phases by evaluating their localization length, adjacent level-spacing ratio (LSR), inverse participation ratio (IPR) and fractal dimensions. The LSR is defined as

r(E)=[1NE2imin(δi,δi+1)/max(δi,δi+1)],r(E)=[\frac{1}{N_{E}-2}\sum_{i}\text{min}(\delta_{i},\delta_{i+1})/\text{max}(\delta_{i},\delta_{i+1})], (8)

where δi=EiEi1\delta_{i}=E_{i}-E_{i-1} with EiE_{i} being the iith eigenenergy sorted in an ascending order and i\sum_{i} denotes the sum over an energy bin around the energy EE with NEN_{E} energy levels counted. For localized states, r0.386r\approx 0.386 corresponding to the Poisson statistics and for extended states of symmetric real Hamiltonians, r0.53r\approx 0.53 corresponding to the Gaussian orthogonal ensemble (GOE) Huse2007PRB .

The localization property can also be characterized by the real space IPR defined as

I(E)=[1NEi𝐫(ν=14|ΨEi,𝐫ν|2)2].I(E)=[\frac{1}{N_{E}}\sum_{i}\sum_{\bf r}(\sum_{\nu=1}^{4}|\Psi_{E_{i},{\bf r}\nu}|^{2})^{2}]. (9)

This quantity evaluates how much a state in an energy bin around energy EE is spatially localized. For an extended state in 2D, I1/L2I\propto 1/L^{2} with LL being the size of a system, which goes zero in the thermodynamic limit; for a state localized in a single unit cell, it is one. It is well known that at the critical point between localized and delocalized phases, the state exhibits multifractal behavior with fractal dimensions D2D_{2} defined through I1/LD2I\propto 1/L^{D_{2}} Castellani1986 . Clearly, D2=2D_{2}=2 and D2=0D_{2}=0 indicates that a state is extended and localized, respectively, in the thermodynamic limit; intermediate values of D2D_{2} suggests the multifractal state.

In Fig. 3(a), we plot the normalized localization length Λx=λx/Ly\Lambda_{x}=\lambda_{x}/L_{y} (similarly for λy/Lx\lambda_{y}/L_{x}) with respect to the disorder strength WW at E=0E=0 for distinct LyL_{y}, where λν\lambda_{\nu} (ν=x,y\nu=x,y) is the localization length along ν\nu calculated by the transfer matrix method Kramer1983 . In the gapless HOTAI and trivial-II phases, we see the decrease of Λx\Lambda_{x} as LyL_{y} is increased, suggesting that the states at E=0E=0 are localized. The decline can also be clearly seen in Fig. 3(b) where Λx\Lambda_{x} versus LyL_{y} is plotted for W=3.2W=3.2 for distinct energies. In fact, all states are localized in these two phases as detailed in the following discussion. This shows that even in the higher-order case, the topology can be carried by localized bulk states. Being localized for the states in these regimes is also evidenced by their relatively large IPR and the LSR approaching 0.3860.386 [see Fig. 3(d)]. In these regimes, the fractal dimension D2D_{2} becomes negative or approaches zero [see Fig. 3(e)], further indicating that the states around zero energy are localized. We note that the negative D2D_{2} arises from finite-size effects. It indicates that the IPR rises with increasing the system size, suggesting that the states are localized (see Appendix F for the finite-size analysis).

Figure 3(a) also demonstrates the existence of a regime (corresponding to the Griffiths regime) where Λx\Lambda_{x} at E=0E=0 remains almost unchanged as LyL_{y} increases, suggesting a multifractal phase in this regime. The multifractal phase resides between two localized phases, which is very different from the conventional wisdom that a multifractal phase lives at the critical point between delocalized and localized phases. In fact, only the states at or very near E=0E=0 become multifractal, and all other states remain localized [see Fig. 3(c)]. The multifractal properties are also evidenced by the fractal dimension of the states around zero energy as shown in Fig. 3(e).

In the gapped regime, there are trivial-I and gapped HOTAI phases. In the trivial phase, the states at the band edge around zero energy exhibit the LSR close to 0.3860.386 [see Fig. 3(d)], suggesting the localized property of these states. The localized property is also evidenced by the negative D2D_{2} (in the region around W=1W=1) [see Fig. 3(e)]. We note that near the phase transition points of W=0W=0 and W=2.1W=2.1, the states exhibit delocalized properties due to the large localization length. In the gapped HOTAI, Fig. 3(d) and (e) illustrate that the LSR experiences a drop from around 0.530.53 to 0.3860.386 and D2D_{2} drops from 1.721.72 to negative values, suggesting that the states at the band edge undergo a phase transition from delocalized to localized ones.

The above results indicate that for strong disorder, all states are localized in the gapless HOTAI and the trivial-II phase. Yet in the Griffiths phase, all states are localized except at E=0E=0 where the states become multifractal. For weak disorder, all states can be localized in the topological regime. In the trivial-I phase, the states at the band edge around zero energy are localized.

In the following, we provide more evidence on localization properties. Figure 4 shows the LSR as a function of energy for five different disorder amplitudes. For small WW corresponding to the trivial-I phase [Fig. 4(a) and its inset], the LSR remains around 0.530.53 except at the lower band edge where it exhibits a sudden drop towards 0.3860.386, indicating that the states at the band edge are localized. But we cannot claim the existence of mobility edges in the trivial-I phase given that it is very possible that the delocalized behavior is caused by the finite-size effects, which is very difficult to identify since the localization length is huge for the weak disorder. In the gapped HOTAI, while we cannot conclusively determine that all states are localized when WW is near the transition point, we show that this occurs when WW is larger. For instance, when W=2.8W=2.8, Fig. 4(b) illustrates that the LSR decreases towards 0.3860.386 with the increase of the system size. We also plot the normalized localization length with respect to LyL_{y} for different energies, the fall of which clearly suggests that the states are localized. These indicators show that all the states are localized. Similarly, Fig. 4(c) indicates that all the states are localized in the gapless HOTAI phase. But in the Griffiths regime, all the states are localized except at E=0E=0 where the LSR remains unchanged as the system size is increased [see Fig. 4(d)].

We also compute the density of states (DOS) at E=0E=0 with respect to the disorder strength WW as shown in Fig. 5(a). The DOS is defined as ρ(E)=𝐫ρ(E,𝐫)/(4LxLy)\rho(E)=\sum_{\bf r}\rho(E,{\bf r})/(4L_{x}L_{y}), which is normalized to one, i.e., 𝑑Eρ(E)=1\int dE\rho(E)=1. Here ρ(E,𝐫)=[jδ(EEj)ν=14|ΨEj,𝐫ν|2]\rho(E,{\bf r})=[\sum_{j}\delta(E-E_{j})\sum_{\nu=1}^{4}|\Psi_{E_{j},{\bf r}\nu}|^{2}] describes the LDOS, where ΨEi,𝐫ν\Psi_{E_{i},{\bf r}\nu} denotes the spatial eigenstate of the system with periodic boundaries corresponding to the eigenenergy EiE_{i}, and [][\cdots] denotes the average over different samples. The DOS rises to the maximum in the multifractal phase and then fall in the trivial-II phase. Specifically, we see the development of a very narrow peak of the DOS at E=0E=0 in this regime [Fig. 5(b)].

V Self-consistent Born approximation

We now explain the disorder induced quadrupole topological insulator based on the self-consistent Born approximation (SCBA) Beenakker2009PRL . As introduced in Sec. II, we consider a disordered system by adding the following random intra-cell hopping terms at each unit cell 𝐫{\bf r}

V(𝐫)\displaystyle V({\bf r}) =W(0iV3(𝐫)iV1(𝐫)0iV3(𝐫)00iV2(𝐫)iV1(𝐫)00iV4(𝐫)0iV2(𝐫)iV4(𝐫)0)\displaystyle=W\left(\begin{array}[]{cccc}0&-iV_{3}({\bf r})&-iV_{1}({\bf r})&0\\ iV_{3}({\bf r})&0&0&iV_{2}({\bf r})\\ iV_{1}({\bf r})&0&0&-iV_{4}({\bf r})\\ 0&-iV_{2}({\bf r})&iV_{4}({\bf r})&0\\ \end{array}\right) (10)
=W[V1(𝐫)U1+V2(𝐫)U2+V3(𝐫)U3+V4(𝐫)U4],\displaystyle=W\left[V_{1}^{\prime}({\bf r})U_{1}+V_{2}^{\prime}({\bf r})U_{2}+V_{3}^{\prime}({\bf r})U_{3}+V_{4}^{\prime}({\bf r})U_{4}\right],

where V1=(V1+V2)/2V_{1}^{\prime}=(V_{1}+V_{2})/2, V2=(V1V2)/2V_{2}^{\prime}=(V_{1}-V_{2})/2, V3=(V3+V4)/2V_{3}^{\prime}=(V_{3}+V_{4})/2, V4=(V3V4)/2V_{4}^{\prime}=(V_{3}-V_{4})/2, and U1=σyσzU_{1}=\sigma_{y}\otimes\sigma_{z}, U2=σyσ0U_{2}=\sigma_{y}\otimes\sigma_{0}, U3=σ0σyU_{3}=\sigma_{0}\otimes\sigma_{y}, U4=σzσyU_{4}=\sigma_{z}\otimes\sigma_{y}. Here we have changed the notation in Sec. II by VxV1V^{x}\rightarrow V_{1}, V¯xV2\bar{V}^{x}\rightarrow V_{2}, VyV3V^{y}\rightarrow V_{3} and V¯yV4\bar{V}^{y}\rightarrow V_{4} for convenience. Since we are interested in disorder without correlations, we require

Vi(𝐑)\displaystyle\langle V_{i}^{\prime}({\bf R})\rangle =\displaystyle= 0\displaystyle 0 (11)
Vi(𝐑1)Vj(𝐑2)\displaystyle\langle V_{i}^{\prime}({\bf R}_{1})V_{j}^{\prime}({\bf R}_{2})\rangle =\displaystyle= 124δijδ𝐑1𝐑2\displaystyle\frac{1}{24}\delta_{ij}\delta_{{\bf R}_{1}{\bf R}_{2}} (12)

for i,j=1,2,3,4i,j=1,2,3,4 with \langle\cdots\rangle denoting the average over disorder ensembles.

Refer to caption
Figure 5: (a) The DOS at zero energy ρ(0)\rho(0) versus the disorder strength WW calculated by the kernel polynomial method (KPM) for the system size L=100L=100 and the expansion order Nc=213N_{c}=2^{13}. (b) The DOS ρ(E)\rho(E) versus EE for different WW with the zoomed-in view around zero energy in the inset. The vertical dashed lines separate different phases. Here mx=my=1.1m_{x}=m_{y}=1.1.

Based on the self-consistent Born approximation, the effective Hamiltonian at E=0E=0 is given by Heff(𝐤)=H0(𝐤)+Σ(E=0)H_{\textrm{eff}}({\bf k})=H_{0}({\bf k})+\Sigma(E=0) where the self-energy Σ\Sigma in the presence of disorder can be calculated through the following self-consistent equation

Σ(E)=W296π2BZd2𝐤n=14UnGUn,\Sigma(E)=\frac{W^{2}}{96\pi^{2}}\int_{BZ}d^{2}{\bf k}\sum_{n=1}^{4}U_{n}GU_{n}, (13)

where G=[(E+i0+)IH0(𝐤)Σ(E)]1G=[(E+i0^{+})I-H_{0}({\bf k})-\Sigma(E)]^{-1}. At energy E=0E=0, we find numerically that the self-energy can be expanded as

Σ=iΣ0I+Σxσyσz+Σyσ0σy,\Sigma=i\Sigma_{0}I+\Sigma_{x}\sigma_{y}\otimes\sigma_{z}+\Sigma_{y}\sigma_{0}\otimes\sigma_{y}, (14)

with Σ0,Σx,Σy\Sigma_{0},\Sigma_{x},\Sigma_{y} being real numbers. It is clear to see that the topological masses mxm_{x} and mym_{y} associated with topological properties are renormalized by disorder to new values

mx\displaystyle m_{x}^{\prime} =mx+Σx,\displaystyle=m_{x}+\Sigma_{x}, (15)
my\displaystyle m_{y}^{\prime} =my+Σy.\displaystyle=m_{y}+\Sigma_{y}. (16)

Based on Eq. (13), we first approximate the self-energy Σ\Sigma by taking Σ=0\Sigma=0 in the right-hand side of the equation, yielding

Σν=W248π2BZ𝑑𝐤Iν,\Sigma_{\nu}=-\frac{W^{2}}{48\pi^{2}}\iint_{BZ}d{\bf k}I_{\nu}, (17)

where

Iν\displaystyle I_{\nu} =\displaystyle= mν+sin(kν)F(𝐤)\displaystyle\frac{m_{\nu}+\sin(k_{\nu})}{F({\bf k})} (18)
F(𝐤)\displaystyle F({\bf k}) =\displaystyle= 2+ν=x,y[mν2+2mνsin(kν)]\displaystyle 2+\sum_{\nu=x,y}[m_{\nu}^{2}+2m_{\nu}\sin(k_{\nu})] (19)

with ν=x,y\nu=x,y. When mx>1m_{x}>1 and my>1m_{y}>1, both Σx\Sigma_{x} and Σy\Sigma_{y} are negative due to the positive integrands, leading to a topological phase transition when the disorder strength WW is sufficiently large so that mx<1m_{x}^{\prime}<1 and my<1m_{y}^{\prime}<1. We also numerically solve the Eq. (13) self-consistently to determine Σx\Sigma_{x} and Σy\Sigma_{y} and plot the results in Fig. 3(f). For weak disorder, the results agree very well with the numerical phase boundary.

VI Disorder effects on HOTIs

Refer to caption
Figure 6: (Color online) (a) The phase diagram with respect to WW, where PP and the quadrupole moments are displayed. Here the gapped/gapless HOTI, the Griffiths phase and the trivial-II phase are observed. (b) Λx\Lambda_{x} at E=0E=0 versus WW for different LyL_{y}. (c) The LSR and IPR versus WW for the eigenstates around zero energy of a system with L=200L=200. The vertical dashed lines separate different phases. Here mx=my=0.5m_{x}=m_{y}=0.5.
Refer to caption
Figure 7: The LSR versus energy EE for (a) W=1W=1, (b) W=2W=2, (c) W=3.2W=3.2 and (d) W=5W=5 in gapped HOTI, gapless HOTI and Griffiths phase, respectively. The top insets in (b), (c) and (d) describe the normalized localization length Λx\Lambda_{x} with respect to LyL_{y} at different energies. The bottom inset in (c) plots the LSR as a function of system size LL at E=1E=-1. Here mx=my=0.5m_{x}=m_{y}=0.5.

In this section, we study the effects of disorder on HOTIs. Specifically, we consider mx=my=0.5m_{x}=m_{y}=0.5 corresponding to a HOTI in the clean limit. We find that the topological phase is stable against weak disorder as evidenced by the quantized topological invariant PP in Fig. 6(a). When the disorder strength becomes sufficiently strong, it enters into a Griffiths regime with fractional PP and finally becomes a trivial phase. The strong disorder also closes the energy gap when W>2.6W>2.6. In the gapless HOTI and trivial-II phases, all states are localized, as evidenced by the normalized localization length, LSR and IPR [see Fig. 6(b) and (c)]. In the Griffiths regime, the states at E=0E=0 are multifractal and all other states are localized [see Fig. 6(b)]. In the disordered gapped HOTI, we find that for weak disorder, the states near the band edge are localized as shown by the LSR around 0.3860.386 in Fig. 6(c). For larger disorder, all the states become localized in this phase.

Figure 7 further plots the LSR with respect to EE for four different disorder strength. When the disorder is weak, e.g., W=1W=1, the LSR shows that the states near the band edge are localized in the gapped HOTI [Fig. 7(a)]. Yet, when W=2W=2, the LSR of all the states decreases towards 0.3860.386 with the increase of the system size, reflecting that all the states are localized. The localized property is also signalled by the decline of the normalized localization length with increasing LyL_{y} [Fig. 7(b)]. Similarly, in the gapless HOTI, all the states are localized as shown in Fig. 7(c). In this case, the LSR at E=1E=-1 decreases towards 0.3860.386 as the system size is increased, providing further evidence for localization. In the Griffiths regime, the LSR becomes smaller for larger system sizes except at E=0E=0 where it remains unchanged, suggesting that the states at E=0E=0 are multifractal and all other states are localized. The multifractal property is also reflected by the unchanged property of the normalized localization length as LyL_{y} is increased [Fig. 7(d)].

VII Experimental realization

Refer to caption
Figure 8: (Color online) (a) Schematics of an electric network for realizing our Hamiltonian (1). Here each node in the circuit represents one lattice site in the Hamiltonian, and four nodes form a unit cell as shown in the blue dashed box. The hopping between two neighboring sites HabH_{ab} is simulated by the admittance RabR_{ab} of the electric element connecting them. The different values of HabH_{ab} correspond to different electric elements, including capacitors, inductors, or INICs whose structure is shown in the red box. Each node aa should be grounded through an electric element with an appropriate admittance RaR_{a}. (b),(c) The averaged magnitude of the impedance |Z(x,y)||Z(x,y)| of each unit cell under open boundary conditions for W=2.6W=2.6 and W=3.4W=3.4, respectively. The insets show the zoomed-in view of |Z(x,y)||Z(x,y)| around one corner. Here mx=my=1.1m_{x}=m_{y}=1.1.

The BBH model has been experimentally realized in several metamaterials, such as microwave, phononic, photonic and topolectrical circuit systems Huber2018Nature ; Bahl2018Nature ; Thomale2018NP ; Hafezi2019NP . In fact, some systems, such as silicon ring resonators Hafezi2019NP , have demonstrated the robustness of zero-energy corner modes to certain disorders. The HOTAI can be easily realized in these systems when the off-diagonal hopping disorder is considered in the experimentally realized BBH model. The BBH model has also been implemented in topolectrical circuits, and zero-energy corner modes are probed by measuring two-point impedances Thomale2018NP . One can involve disorder in the system by tuning the capacitance of capacitors and inductance of inductors to realize the HOTAI, as we have proved that this model is equivalent to our model in topological and localization properties (see Appendix B). In the following, we discuss in detail an experimental scheme to realize the Hamiltonian (1) using topolectrical circuits and show that the HOTAI phase can be detected by two-point impedance measurements.

Let us consider an electric network composed of different nodes and electric element connecting nodes, as shown in Fig. 8(a). We denote the input current and voltage of each node aa by IaI_{a} and VaV_{a}, respectively. According to Kirchhoff’s law, the circuit at a frequency of ω\omega should satisfy the relation

Ia(ω)=bRab(Va(ω)Vb(ω))+RaVa(ω),I_{a}(\omega)=\sum_{b}R_{ab}(V_{a}(\omega)-V_{b}(\omega))+R_{a}V_{a}(\omega), (20)

where RabR_{ab} is the admittance of the corresponding electric element between the node aa and bb, and RaR_{a} is the admittance of the electric element between the node aa and the ground. We can rewrite the above equation into a compact form as

𝐈(ω)=J(ω)𝐕(ω),{\bf I}(\omega)=J(\omega){\bf V}(\omega), (21)

where 𝐈{\bf I} and 𝐕{\bf V} are NN-component column vectors with components IaI_{a} and VaV_{a} for NN nodes, respectively. Here the matrix JJ is the circuit Laplacian. Then we can simulate our Hamiltonian HH with the Laplacian J(ω)J(\omega) at a proper frequency ω\omega through

J(ω)=iH.J(\omega)=iH. (22)

Each node in the circuit represents one lattice site in our Hamiltonian, and each electric element linking two nodes represents the corresponding hopping between the sites, which can be either a capacitor, or an inductor, or a negative impedance converter with current inversion (INIC). For two nodes in the circuit, the electric element between them is determined according to the corresponding matrix element HabH_{ab} between the site aa and bb in our Hamiltonian, as illustrated in Fig. 8(a). Specifically, for two neighboring sites in our Hamiltonian, if HabH_{ab} is a positive (negative) real number, the electric element between aa and bb should be an inductor (a capacitor) with inductance (capacitance) 1ωHab\frac{1}{\omega H_{ab}} (Habω-\frac{H_{ab}}{\omega}). For the case that HabH_{ab} is an imaginary number, we should connect the two nodes using an INIC with resistance 1|Hab|\frac{1}{|H_{ab}|} and proper direction. In addition, we connect every node with the ground by appropriate electric elements to eliminate the extra diagonal terms in the Laplacian.

Similar to the experimental work Thomale2018NP , we utilize the two-point impedance measurement in the circuit to characterize the zero-energy corner modes in the HOTAI phase. The two-point impedance between node aa and bb is defined as

Zab=n|ψn,aψn,b|2jn,Z_{ab}=\sum_{n}\frac{|\psi_{n,a}-\psi_{n,b}|^{2}}{j_{n}}, (23)

where ψn,a\psi_{n,a} is the component for the node aa of the nnth eigenvector of JJ with eigenvalue jnj_{n}. We define the impedance of each unit cell Z(x,y)Z(x,y) as the average two-point impedance between nearest-neighbor nodes within each unit cell as

Z(x,y)=Z12(x,y)+Z24(x,y)+Z43(x,y)+Z31(x,y)4,Z(x,y)=\frac{Z_{12}(x,y)+Z_{24}(x,y)+Z_{43}(x,y)+Z_{31}(x,y)}{4}, (24)

where Zij(x,y)Z_{ij}(x,y) denotes the two-point impedance between the iith node and jjth node of the unit cell (x,y)(x,y). Fig. 8(b) and (c) plot the magnitude of Z(x,y)Z(x,y) averaged over 400400 random samples under open boundary conditions for two values of WW within HOTAI regime for mx=my=1.1m_{x}=m_{y}=1.1, which clearly show the impedance resonance near the corners corresponding to the presence of zero-energy corner modes of the Hamiltonian.

VIII Conclusion

In summary, we have discovered the HOTAI in a 2D disordered system with chiral symmetry. Specifically, we show that a topologically trivial phase can transition into a quadrupole topological phase, when disorder is added. We find gapped and gapless HOTAIs and a Griffiths regime. In the gapless HOTAI, all the states are localized, while in the Griffiths regime, the states at zero energy are multifractal and other states are localized. The Griffiths regime corresponds to a critical regime between two localized phases: a gapless HOTAI and a trivial phase. We also propose an experimental scheme with topolectrical circuits to realize the HOTAI. Our results demonstrate that disorder can induce quadrupole topological insulators with peculiar localization properties from a trivial phase and thus opens a new avenue for studying the role of disorder in higher-order topological phases.

Acknowledgements.
We thank Y.-L. Tao and N. Dai for helpful discussion. Y.B.Y., K.L. and Y.X. are supported by the National Natural Science Foundation of China (11974201), the start-up fund from Tsinghua University and the National Thousand-Young-Talents Program. We acknowledge in addition support from the Frontier Science Center for Quantum Information of the Ministry of Education of China, Tsinghua University Initiative Scientific Research Program, and the National key Research and Development Program of China (2016YFA0301902).

Note added: Recently, we became aware of two related work Shen2020PRL ; Zhang2020arXiv .

Appendix A: Generalized C4C_{4} symmetry

In the clean case, when mx=mym_{x}=m_{y} and tx=tyt_{x}=t_{y}, the Hamiltonian (1) in the main text respects a generalized C4C_{4} symmetry,

UC4H^UC41=H^,U_{C_{4}}\hat{H}U_{C_{4}}^{-1}=\hat{H}, (A1)

where

UC4c^𝐫UC41=S𝐫c^g𝐫,UC4c^𝐫UC41=c^g𝐫S𝐫T,U_{C_{4}}\hat{c}_{\bf r}U_{C_{4}}^{-1}=S_{\bf r}\hat{c}_{g{\bf r}},~{}~{}~{}U_{C_{4}}\hat{c}_{\bf r}^{\dagger}U_{C_{4}}^{-1}=\hat{c}_{g{\bf r}}^{\dagger}S_{\bf r}^{T}, (A2)

with

S𝐫=(00(1)y0(1)y000000(1)y0(1)y00)S_{\bf r}=\left(\begin{array}[]{cccc}0&0&(-1)^{y}&0\\ -(-1)^{y}&0&0&0\\ 0&0&0&(-1)^{y}\\ 0&(-1)^{y}&0&0\\ \end{array}\right) (A3)

and gg being a C4C_{4} rotation operator such that g𝐫=(y,x)g{\bf r}=(-y,x).

In momentum space, let us write H^=𝐤c^𝐤H(𝐤)c^𝐤\hat{H}=\sum_{\bf k}\hat{c}_{\bf k}^{\dagger}H({\bf k})\hat{c}_{\bf k} with c^𝐤=(c^𝐤1c^𝐤2c^𝐤3c^𝐤4)\hat{c}_{\bf k}^{\dagger}=(\begin{array}[]{cccc}\hat{c}_{{\bf k}1}^{\dagger}&\hat{c}_{{\bf k}2}^{\dagger}&\hat{c}_{{\bf k}3}^{\dagger}&\hat{c}_{{\bf k}4}^{\dagger}\end{array}). The generalized C4C_{4} symmetry takes the following form

S1H(𝐤)S1=H(g𝐤),S_{1}^{\dagger}H({\bf k})S_{1}=H(g{\bf k}^{\prime}), (A4)

where 𝐤=(kx,kyπ){\bf k}^{\prime}=(k_{x},k_{y}-\pi) and

S1=(0010100000010100).S_{1}=\left(\begin{array}[]{cccc}0&0&1&0\\ -1&0&0&0\\ 0&0&0&1\\ 0&1&0&0\\ \end{array}\right). (A5)

Appendix B: Equivalence between our model and the disordered BBH model

In this Appendix, we will prove that our model is equivalent to the BBH model in topological and localization properties. The BBH model reads

H~^=𝐫[c^𝐫h~0c^𝐫+(c^𝐫hxc^𝐫+𝐞x+c^𝐫hyc^𝐫+𝐞y+H.c.)],\hat{\tilde{H}}=\sum_{{\bf r}}\left[\hat{c}^{\dagger}_{{\bf r}}\tilde{h}_{0}\hat{c}_{{\bf r}}+\left(\hat{c}^{\dagger}_{{\bf r}}h_{x}\hat{c}_{{\bf r}+{\bf e}_{x}}+\hat{c}^{\dagger}_{{\bf r}}h_{y}\hat{c}_{{\bf r}+{\bf e}_{y}}+H.c.\right)\right], (B1)

where h~0\tilde{h}_{0} is a real matrix expressed as

h~0=(0m𝐫ym𝐫x0m𝐫y00m¯𝐫xm𝐫x00m¯𝐫y0m¯𝐫xm¯𝐫y0).\tilde{h}_{0}=\left(\begin{array}[]{cccc}0&m_{{\bf r}}^{y}&m_{{\bf r}}^{x}&0\\ m_{{\bf r}}^{y}&0&0&-\bar{m}_{{\bf r}}^{x}\\ m_{{\bf r}}^{x}&0&0&\bar{m}_{{\bf r}}^{y}\\ 0&-\bar{m}_{{\bf r}}^{x}&\bar{m}_{{\bf r}}^{y}&0\\ \end{array}\right). (B2)

This model respects the time-reversal, particle-hole and chiral symmetries.

While the two Hamiltonians have different symmetries, they are closely related by a local transformation U𝐫=diag(ix+y1,ix+y,ix+y,ix+y+1)U_{{\bf r}}=\text{diag}(i^{x+y-1},i^{x+y},i^{x+y},i^{x+y+1}), that is, U𝐫h0U𝐫=h~0U_{{\bf r}}^{\dagger}h_{0}U_{{\bf r}}=\tilde{h}_{0}, U𝐫hxU𝐫+𝐞x=hxU_{{\bf r}}^{\dagger}h_{x}U_{{\bf r}+{\bf e}_{x}}={h}_{x} and U𝐫hyU𝐫+𝐞y=hyU_{{\bf r}}^{\dagger}h_{y}U_{{\bf r}+{\bf e}_{y}}={h}_{y}. Specifically, one can transform H^\hat{H} in Eq. (1) to H~^\hat{\tilde{H}} by the transformation c^𝐫U𝐫c^𝐫\hat{c}_{{\bf r}}\rightarrow U_{{\bf r}}\hat{c}_{{\bf r}}. In other words, if ΨEi,𝐫ν\Psi_{E_{i},{\bf r}\nu} is a spatial eigenstate of HH, then Ψ~Ei,𝐫ν=(i)f𝐫νΨEi,𝐫ν\tilde{\Psi}_{E_{i},{\bf r}\nu}=(-i)^{f_{{\bf r}\nu}}\Psi_{E_{i},{\bf r}\nu} with f𝐫1=x+y1f_{{\bf r}1}=x+y-1, f𝐫2=f𝐫3=x+yf_{{\bf r}2}=f_{{\bf r}3}=x+y, and f𝐫4=x+y+1f_{{\bf r}4}=x+y+1 is an eigenstate of H~\tilde{H} corresponding to the same energy EiE_{i}. Here HH and H~\tilde{H} are the first-quantization Hamiltonians of H^\hat{H} and H~^\hat{\tilde{H}}, respectively. Therefore, H^\hat{H} and H~^\hat{\tilde{H}} have the same energy spectrum and density profiles, indicating identical localization properties that they possess. In addition, this local phase transformation does not change the topological property, and thus the two Hamiltonians have the same topology. Under open boundary conditions, the two models are connected by the transformation irrelevant of the system size. Yet, under periodic boundary conditions, the transformation works well only when LxL_{x} and LyL_{y} are integer multiples of four. For topological property, the two models should be equivalent irrelevant of a system size given that the topology does not depend on a specific system size. For localization property, we have also calculated the IPR and LSR of the two Hamiltonians with their sizes being odd and find similar results, showing that their localization properties are irrelevant of the parity of a system size.

Appendix C: Quantization of quadrupole moments by chiral symmetry

In this Appendix, we will prove that the quadrupole moment is protected to be quantized by chiral symmetry and thus can be used as a topological invariant. Note that the quadrupole moment may not characterize the physical quadrupole moment, we here are only interested in the formula as a topological invariant. We consider the quadrupole moment defined by Wheeler2018arXiv ; Cho2018arXiv

qxy\displaystyle q_{xy} [q~xyqxy(0)]mod1\displaystyle\equiv\left[\tilde{q}_{xy}-q_{xy}^{(0)}\right]\mod 1 (C1)
=[12πImlogΨG|e2πiQ^xyΨGqxy(0)]mod1,\displaystyle=\left[\frac{1}{2\pi}\mathrm{Im}\log\langle\Psi_{G}|e^{2\pi i\hat{Q}_{xy}}|\Psi_{G}\rangle-q_{xy}^{(0)}\right]\mod 1,

where Q^xy=j=1ncx^jy^j/(LxLy)\hat{Q}_{xy}=\sum_{j=1}^{n_{c}}\hat{x}_{j}\hat{y}_{j}/(L_{x}L_{y}) with x^j\hat{x}_{j} (y^j\hat{y}_{j}) denoting the xx-position (yy-position) operator for electron jj with nc=2LxLyn_{c}=2L_{x}L_{y} (the number of occupied states in our model) at half filling, and |ΨG|\Psi_{G}\rangle is the many-body ground state of electrons in the system. Here q~xy=12πIm[logΨG|e2πiQ^xyΨG]\tilde{q}_{xy}=\frac{1}{2\pi}\mathrm{Im}[\log\langle\Psi_{G}|e^{2\pi i\hat{Q}_{xy}}|\Psi_{G}\rangle] is the contribution from occupied electrons, and qxy(0)=12j=1naxjyj/(LxLy)q_{xy}^{(0)}=\frac{1}{2}\sum_{j=1}^{n_{a}}x_{j}y_{j}/(L_{x}L_{y}) is the contribution from the background positive charge distribution where (xj,yj)(x_{j},y_{j}) denotes the position of the jjth atomic orbital. Here, nan_{a} is the total number of atomic orbitals so that the single-particle Hamiltonian is a na×nan_{a}\times n_{a} matrix. At half filling, na=2ncn_{a}=2n_{c}.

Let us write the many-body wave function of occupied electrons in real space representation as

ΨG(𝐫1ν1,𝐫2ν2,,𝐫ncνnc)\displaystyle\Psi_{G}({\bf r}_{1}\nu_{1},{\bf r}_{2}\nu_{2},\cdots,{\bf r}_{n_{c}}\nu_{n_{c}})
=\displaystyle= 1nc!|ψ1(𝐫1ν1)ψ2(𝐫1ν1)ψnc(𝐫1ν1)ψ1(𝐫2ν2)ψ2(𝐫2ν2)ψnc(𝐫2ν2)ψ1(𝐫ncνnc)ψ2(𝐫ncνnc)ψnc(𝐫ncνnc)|,\displaystyle\frac{1}{\sqrt{n_{c}!}}\left|\begin{array}[]{cccc}\psi_{1}({\bf r}_{1}\nu_{1})&\psi_{2}({\bf r}_{1}\nu_{1})&\cdots&\psi_{n_{c}}({\bf r}_{1}\nu_{1})\\ \psi_{1}({\bf r}_{2}\nu_{2})&\psi_{2}({\bf r}_{2}\nu_{2})&\cdots&\psi_{n_{c}}({\bf r}_{2}\nu_{2})\\ \vdots&\vdots&\ddots&\vdots\\ \psi_{1}({\bf r}_{n_{c}}\nu_{n_{c}})&\psi_{2}({\bf r}_{n_{c}}\nu_{n_{c}})&\cdots&\psi_{n_{c}}({\bf r}_{n_{c}}\nu_{n_{c}})\\ \end{array}\right|, (C6)

where ψn\psi_{n} represents the nnth occupied eigenstate of a first-quantization Hamiltonian. Then, the quadrupole moment of occupied electrons q~xy\tilde{q}_{xy} can be evaluated through

q~xy=12πImlogΨG|Ψ~G,\tilde{q}_{xy}=\frac{1}{2\pi}\mathrm{Im}\log\langle\Psi_{G}|\tilde{\Psi}_{G}\rangle, (C7)

where

ΨG|Ψ~G=|ψ1|ψ~1ψ1|ψ~2ψ1|ψ~ncψ2|ψ~1ψ2|ψ~2ψ2|ψ~ncψnc|ψ~1ψnc|ψ~2ψnc|ψ~2|,\langle\Psi_{G}|\tilde{\Psi}_{G}\rangle=\left|\begin{array}[]{cccc}\langle\psi_{1}|\tilde{\psi}_{1}\rangle&\langle\psi_{1}|\tilde{\psi}_{2}\rangle&\cdots&\langle\psi_{1}|\tilde{\psi}_{n_{c}}\rangle\\ \langle\psi_{2}|\tilde{\psi}_{1}\rangle&\langle\psi_{2}|\tilde{\psi}_{2}\rangle&\cdots&\langle\psi_{2}|\tilde{\psi}_{n_{c}}\rangle\\ \vdots&\vdots&\ddots&\vdots\\ \langle\psi_{n_{c}}|\tilde{\psi}_{1}\rangle&\langle\psi_{n_{c}}|\tilde{\psi}_{2}\rangle&\cdots&\langle\psi_{n_{c}}|\tilde{\psi}_{2}\rangle\\ \end{array}\right|, (C8)

ψ~n(𝐫ν)=ei2πxy/(LxLy)ψn(𝐫ν)\tilde{\psi}_{n}({\bf r}\nu)=e^{i2\pi xy/(L_{x}L_{y})}\psi_{n}({\bf r}\nu) and ψm|ψ~n=𝐫αψm(𝐫α)ψ~n(𝐫α)\langle\psi_{m}|\tilde{\psi}_{n}\rangle=\sum_{{\bf r}\alpha}\psi_{m}^{*}({\bf r}\alpha)\tilde{\psi}_{n}({\bf r}\alpha). Let us define Uo=(|ψ1,|ψ2,,|ψnc)U_{o}=\left(|\psi_{1}\rangle,|\psi_{2}\rangle,\cdots,|\psi_{n_{c}}\rangle\right) which is a na×ncn_{a}\times n_{c} matrix representing the occupied states of electrons. Then we can express the quadrupole moment of occupied electrons as

q~xy=12πImlogdet(UoD^Uo),\tilde{q}_{xy}=\frac{1}{2\pi}\mathrm{Im}\log\det(U_{o}^{\dagger}\hat{D}U_{o}), (C9)

where we define a na×nan_{a}\times n_{a} diagonal matrix D^=diag{e2πixjyj/(LxLy)}j=1na\hat{D}=\text{diag}\{e^{2\pi ix_{j}y_{j}/(L_{x}L_{y})}\}_{j=1}^{n_{a}} with (xj,yj)(x_{j},y_{j}) denoting the position of jj-th atomic orbital.

For a generic Hamiltonian in real space HH with chiral (sublattice) symmetry, ΠHΠ1=H{\Pi}H{\Pi}^{-1}=-H, if |ψn|\psi_{n}\rangle is an eigenstate of HH corresponding to energy EnE_{n}, Π|ψn{\Pi}|\psi_{n}\rangle is also an eigenstate of HH with energy En-E_{n}, corresponding to an unoccupied state. The set {Π|ψ1,Π|ψ2,,Π|ψnc}\{{\Pi}|\psi_{1}\rangle,{\Pi}|\psi_{2}\rangle,\cdots,{\Pi}|\psi_{n_{c}}\rangle\} therefore constitutes the unoccupied states. We then define Uu=(Π|ψ1,Π|ψ2,,Π|ψnc)=ΠUoU_{u}=\left({\Pi}|\psi_{1}\rangle,{\Pi}|\psi_{2}\rangle,\cdots,{\Pi}|\psi_{n_{c}}\rangle\right)={\Pi}U_{o} representing the unoccupied states of electrons. The quadrupole moment q~xyu\tilde{q}_{xy}^{u} for the unoccupied states is

q~xyu\displaystyle\tilde{q}_{xy}^{u} =12πImlogdet(UuD^Uu)\displaystyle=\frac{1}{2\pi}\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}U_{u}) (C10)
=12πImlogdet(UoΠD^ΠUo).\displaystyle=\frac{1}{2\pi}\mathrm{Im}\log\det(U_{o}^{\dagger}\Pi^{\dagger}\hat{D}\Pi U_{o}). (C11)

Clearly, D^\hat{D} commutes with the chiral (sublattice) symmetry transformation Π\Pi, i.e., [D^,Π]=0\left[\hat{D},\Pi\right]=0,

q~xyu=12πImlogdet(UoD^Uo)=q~xy.\tilde{q}_{xy}^{u}=\frac{1}{2\pi}\mathrm{Im}\log\det(U_{o}^{\dagger}\hat{D}U_{o})=\tilde{q}_{xy}. (C12)

Let us define qxyu[q~xyuqxy(0)]mod1q_{xy}^{u}\equiv\left[\tilde{q}_{xy}^{u}-q_{xy}^{(0)}\right]\mod 1. Then we will have

qxyu=[q~xyqxy(0)]mod1=qxy.q_{xy}^{u}=\left[\tilde{q}_{xy}-q_{xy}^{(0)}\right]\mod 1=q_{xy}. (C13)

Next we will prove that qxy+qxyu=0mod1q_{xy}+q_{xy}^{u}=0\mod 1, i.e., q~xy+q~xyu2qxy(0)=0mod1\tilde{q}_{xy}+\tilde{q}_{xy}^{u}-2q_{xy}^{(0)}=0\mod 1.

Proof.

We define a unitary matrix Ut=(Uo,Uu)U_{t}=\left(U_{o},U_{u}\right). It can be easily seen that

2qxy(0)\displaystyle 2q_{xy}^{(0)} =j=1naxjyj/(LxLy)\displaystyle=\sum_{j=1}^{n_{a}}x_{j}y_{j}/(L_{x}L_{y}) (C14)
=12πImlogdetD^mod1\displaystyle=\frac{1}{2\pi}\mathrm{Im}\log\det\hat{D}\mod 1 (C15)
=12πImlogdet(UtD^Ut)mod1.\displaystyle=\frac{1}{2\pi}\mathrm{Im}\log\det(U_{t}^{\dagger}\hat{D}U_{t})\mod 1. (C16)

Then we have the following relations

2π(q~xyq~xyu+2qxy(0))\displaystyle 2\pi(-\tilde{q}_{xy}-\tilde{q}_{xy}^{u}+2q_{xy}^{(0)}) (C17)
=\displaystyle= Imlogdet(UoD^Uo)Imlogdet(UuD^Uu)+\displaystyle-\mathrm{Im}\log\det(U_{o}^{\dagger}\hat{D}U_{o})-\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}U_{u})+
Imlogdet(UtD^Ut)\displaystyle\mathrm{Im}\log\det(U_{t}^{\dagger}\hat{D}U_{t})
=\displaystyle= Imlogdet(UoD^Uo)+Imlogdet(UuD^Uu)+\displaystyle\mathrm{Im}\log\det(U_{o}^{\dagger}\hat{D}^{\dagger}U_{o})+\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}^{\dagger}U_{u})+
Imlogdet(UoDUoUoD^UuUuDUoUuD^Uu)\displaystyle\mathrm{Im}\log\det\begin{pmatrix}U_{o}^{\dagger}DU_{o}&U_{o}^{\dagger}\hat{D}U_{u}\\ U_{u}^{\dagger}DU_{o}&U_{u}^{\dagger}\hat{D}U_{u}\\ \end{pmatrix}
=\displaystyle= Imlogdet(UoD^UoUoD^Uu0UuD^Uu)+\displaystyle\mathrm{Im}\log\det\begin{pmatrix}U_{o}^{\dagger}\hat{D}^{\dagger}U_{o}&U_{o}^{\dagger}\hat{D}^{\dagger}U_{u}\\ 0&U_{u}^{\dagger}\hat{D}^{\dagger}U_{u}\\ \end{pmatrix}+
Imlogdet(UoD^UoUoD^UuUuD^UoUuD^Uu)\displaystyle\mathrm{Im}\log\det\begin{pmatrix}U_{o}^{\dagger}\hat{D}U_{o}&U_{o}^{\dagger}\hat{D}U_{u}\\ U_{u}^{\dagger}\hat{D}U_{o}&U_{u}^{\dagger}\hat{D}U_{u}\\ \end{pmatrix}
=\displaystyle= Imlogdet[(UoD^UoUoD^Uu0UuD^Uu)(UoD^UoUoD^UuUuD^UoUuD^Uu)]\displaystyle\mathrm{Im}\log\det\left[\begin{pmatrix}U_{o}^{\dagger}\hat{D}^{\dagger}U_{o}&U_{o}^{\dagger}\hat{D}^{\dagger}U_{u}\\ 0&U_{u}^{\dagger}\hat{D}^{\dagger}U_{u}\\ \end{pmatrix}\begin{pmatrix}U_{o}^{\dagger}\hat{D}U_{o}&U_{o}^{\dagger}\hat{D}U_{u}\\ U_{u}^{\dagger}\hat{D}U_{o}&U_{u}^{\dagger}\hat{D}U_{u}\\ \end{pmatrix}\right]
=\displaystyle= Imlogdet(𝟙0UuD^UuUuD^UoUuD^UuUuD^Uu)\displaystyle\mathrm{Im}\log\det\begin{pmatrix}\mathds{1}&0\\ U_{u}^{\dagger}\hat{D}^{\dagger}U_{u}U_{u}^{\dagger}\hat{D}U_{o}&U_{u}^{\dagger}\hat{D}^{\dagger}U_{u}U_{u}^{\dagger}\hat{D}U_{u}\\ \end{pmatrix}
=\displaystyle= Imlogdet(UuD^UuUuD^Uu)\displaystyle\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}^{\dagger}U_{u}U_{u}^{\dagger}\hat{D}U_{u})
=\displaystyle= Imlogdet(UuD^Uu)+Imlogdet(UuD^Uu)\displaystyle\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}U_{u})^{\dagger}+\mathrm{Im}\log\det(U_{u}^{\dagger}\hat{D}U_{u})
=\displaystyle= 0mod2π.\displaystyle 0\mod 2\pi.

In the derivation, we have utilized the orthonormal properties UoUo=UuUu=𝟙U_{o}^{\dagger}U_{o}=U_{u}^{\dagger}U_{u}=\mathds{1}, UoUu=0U_{o}^{\dagger}U_{u}=0 and UoUo+UuUu=𝟙U_{o}U_{o}^{\dagger}+U_{u}U_{u}^{\dagger}=\mathds{1}. ∎

Therefore, we have the following relation

qxy+qxyu=0mod1.q_{xy}+q_{xy}^{u}=0\mod 1. (C18)

Combined with the relation that qxyu=qxyq_{xy}^{u}=q_{xy}, we get the conclusion that 2qxy=0mod12q_{xy}=0\mod 1, namely, qxyq_{xy} is quantized to 0 or 0.50.5 up to an integer. The result is consistent with our numerical results where all disorder configurations exhibit quantized quadrupole moments. We note that this proof remains valid when we replace Q^xy\hat{Q}_{xy} and qxy(0)q_{xy}^{(0)} in the definition of qxyq_{xy} with Q^f=j=1ncf(x^j,y^j)/(LxLy)\hat{Q}_{f}=\sum_{j=1}^{n_{c}}f(\hat{x}_{j},\hat{y}_{j})/(L_{x}L_{y}) and qf(0)=12j=1naf(xj,yj)/(LxLy)q_{f}^{(0)}=\frac{1}{2}\sum_{j=1}^{n_{a}}f(x_{j},y_{j})/(L_{x}L_{y}), respectively where f(x,y)f(x,y) is a general function so that the newly defined quantity is also quantized by chiral symmetry like the quadrupole moment. In addition, one can use the same procedure to prove the quantization of the octupole moment in 3D protected by chiral symmetry.

We note that while we have proved that the quadrupole moment is quantized to either 0 or 0.50.5 for each disorder configuration protected by chiral symmetry, for a disorder system, we need to consider many distinct samples and perform the average of the quadrupole moment over these samples. In this case, the averaged quadrupole moment may not be quantized since for some samples the quadruple moments are equal to 0.50.5 and for others they are equal to 0 when a system size is not large as shown in Fig. C1.

Refer to caption
Figure C1: (Color online) The quadrupole moment qxyq_{xy} for different disorder configurations when W=2.6W=2.6, showing that qxy=0.5q_{xy}=0.5 for most disorder samples and qxy=0q_{xy}=0 for others. Here Lx=Ly=80L_{x}=L_{y}=80 and mx=my=1.1m_{x}=m_{y}=1.1.

It is worth mentioning that Ref. Agarwala2020PRR has found that quadrupole topological insulators with quantized quadrupole moments can still exist even in amorphous systems without crystalline symmetries. We now can understand that the quantized quadrupole moment found in Ref. Agarwala2020PRR is protected by chiral symmetry.

Appendix D: Effective boundary Hamiltonian

In this section, we follow the transfer matrix method introduced in Ref. Oppen2017PRB to derive the effective boundary Hamiltonian of our system in the clean case. We will show that the effective boundary Hamiltonian at the yy-normal (xx-normal) edges are proportional to Hx(kx,mx)H_{x}(k_{x},m_{x}) [Hy(ky,my)H_{y}(k_{y},m_{y})] up to a nonzero factor, implying that the higher-order topology can be characterized by the topological invariant PP introduced in the main text.

Specifically, let us write the Hamiltonian as H^=c^Hc^\hat{H}=\hat{c}^{\dagger}H\hat{c} where c^=(c^1c^2c^Ly)\hat{c}^{\dagger}=(\begin{array}[]{cccc}\hat{c}_{1}^{\dagger}&\hat{c}_{2}^{\dagger}&\cdots&\hat{c}_{L_{y}}^{\dagger}\end{array}) with the index jj denoting the jjth layer consisting of sites along xx and HH reads

H=(h1V10000V1h2V20000V2h3V30000V3h4V40000V2Ly2h2Ly1V2Ly10000V2Ly1h2Ly)H=\left(\begin{array}[]{ccccccc}h_{1}&V_{1}^{\dagger}&0&0&0&\cdots&0\\ V_{1}&h_{2}&V_{2}^{\dagger}&0&0&\cdots&0\\ 0&V_{2}&h_{3}&V_{3}^{\dagger}&0&\cdots&0\\ 0&0&V_{3}&h_{4}&V_{4}^{\dagger}&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\ddots&\ddots&\vdots\\ 0&0&0&\cdots&V_{2L_{y}-2}&h_{2L_{y}-1}&V_{2L_{y}-1}^{\dagger}\\ 0&0&0&\cdots&0&V_{2L_{y}-1}&h_{2L_{y}}\\ \end{array}\right) (D1)

with VnV_{n} denoting the coupling between the nnth and (n+1)(n+1)th layer. In disordered systems, the parameters in hnh_{n} and V2n1V_{2n-1} describing the intra-cell hopping are randomly generated.

In the clean case, the system has the translational invariance of period 2 and thus there are two different layers described by the Hamiltonian h1h_{1} and h2h_{2}, respectively. If we view these two layers as a unit cell, we use V1V_{1} and V2V_{2} to describe the intra-cell and inter-cell layer coupling, respectively. Now HH can be simplified as

H=(h1V10000V1h2V20000V2h1V10000V1h2V20000V2h1V10000V1h2).H=\left(\begin{array}[]{ccccccc}h_{1}&V_{1}^{\dagger}&0&0&0&\cdots&0\\ V_{1}&h_{2}&V_{2}^{\dagger}&0&0&\cdots&0\\ 0&V_{2}&h_{1}&V_{1}^{\dagger}&0&\cdots&0\\ 0&0&V_{1}&h_{2}&V_{2}^{\dagger}&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\ddots&\ddots&\vdots\\ 0&0&0&\cdots&V_{2}&h_{1}&V_{1}^{\dagger}\\ 0&0&0&\cdots&0&V_{1}&h_{2}\\ \end{array}\right). (D2)

Considering the periodic boundaries along xx, we write h1h_{1}, h2h_{2}, V1V_{1} and V2V_{2} in momentum space as h1=h2=Hx(kx,mx)h_{1}=-h_{2}=H_{x}(k_{x},m_{x}), V1=imyσ0V_{1}=im_{y}\sigma_{0} and V2=σ0V_{2}=\sigma_{0}. When my=0m_{y}=0, it is clear to see that the effective boundary Hamiltonian is Hx(kx,mx)H_{x}(k_{x},m_{x}). When my0m_{y}\neq 0, we obtain the following two transfer matrices at energy EE

M1(E)=(i(Eσ0h1)/myiσ0/myσ002×2),\displaystyle M_{1}(E)=\left(\begin{array}[]{cc}i(E\sigma_{0}-h_{1})/m_{y}&-i\sigma_{0}/m_{y}\\ \sigma_{0}&0_{2\times 2}\\ \end{array}\right), (D3)
M2(E)=((Eσ0h2)imyσ0σ002×2),\displaystyle M_{2}(E)=\left(\begin{array}[]{cc}(E\sigma_{0}-h_{2})&-im_{y}\sigma_{0}\\ \sigma_{0}&0_{2\times 2}\\ \end{array}\right),

where the transfer matrices connect the eigenstate in neighboring layers through

(Ψ2nΨ2n1)=M1(Ψ2n1Ψ2n2),\displaystyle\left(\begin{array}[]{c}\Psi_{2n}\\ \Psi_{2n-1}\\ \end{array}\right)=M_{1}\left(\begin{array}[]{c}\Psi_{2n-1}\\ \Psi_{2n-2}\\ \end{array}\right), (D4)
(Ψ2n+1Ψ2n)=M2(Ψ2nΨ2n1),\displaystyle\left(\begin{array}[]{c}\Psi_{2n+1}\\ \Psi_{2n}\\ \end{array}\right)=M_{2}\left(\begin{array}[]{c}\Psi_{2n}\\ \Psi_{2n-1}\\ \end{array}\right),

with n1n\geq 1 and Ψn\Psi_{n} is the component in the nnth layer of an eigenstate with the energy EE.

Refer to caption
Figure D1: (Color online) The polarization (a) and (c) pxp_{x}, and (b) and (d) pyp_{y} for the 801801th to 10001000th iterations corresponding to different disorder configurations. The region between two vertical lines correspond to the Griffiths regime. In (a-b), mx=my=1.1m_{x}=m_{y}=1.1, and in (c-d), mx=my=0.5m_{x}=m_{y}=0.5. Here Lx=Ly=500L_{x}=L_{y}=500.

We now define the total transfer matrix at zero energy as

T=M2(E=0)M1(E=0)=imy(A(kx)σ0h1h1σ0),T=M_{2}(E=0)M_{1}(E=0)=-\frac{i}{m_{y}}\left(\begin{array}[]{cc}A(k_{x})\sigma_{0}&h_{1}\\ h_{1}&\sigma_{0}\\ \end{array}\right), (D5)

where A(kx)=1+mx2+my2+2mxsinkxA(k_{x})=1+m_{x}^{2}+m_{y}^{2}+2m_{x}\sin k_{x}. This matrix can be reduced to a diagonal block form through an elementary interchange transformation,

S24TS24=imy(H102×202×2H2),S_{24}TS_{24}=-\frac{i}{m_{y}}\left(\begin{array}[]{cc}H_{1}&0_{2\times 2}\\ 0_{2\times 2}&H_{2}\\ \end{array}\right), (D6)

where

S24=(1000000100100100),\displaystyle S_{24}=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&0&1\\ 0&0&1&0\\ 0&1&0&0\\ \end{array}\right), (D7)
H1=(A(kx)imx+eikximx+eikx1),\displaystyle H_{1}=\left(\begin{array}[]{cc}A(k_{x})&-im_{x}+e^{-ik_{x}}\\ im_{x}+e^{ik_{x}}&1\\ \end{array}\right),
H2=(1imx+eikximx+eikxA(kx)).\displaystyle H_{2}=\left(\begin{array}[]{cc}1&-im_{x}+e^{-ik_{x}}\\ im_{x}+e^{ik_{x}}&A(k_{x})\\ \end{array}\right).

Evidently, H1H_{1} and H2H_{2} have the same eigenvalues. Since TT is a symplectic matrix, its eigenvalues show up in pairs as (λ,1/λ)(\lambda,1/\lambda^{*}). Suppose that λ1my\lambda_{1}m_{y} (|λ1|>1|\lambda_{1}|>1) is an eigenvalue of H1H_{1}, then

TU\displaystyle TU =T(U11U12U21U22)\displaystyle=T\left(\begin{array}[]{cc}U_{11}&U_{12}\\ U_{21}&U_{22}\\ \end{array}\right) (D10)
=i(U11U12U21U22)(λ1σ002×202×2σ0/λ1),\displaystyle=-i\left(\begin{array}[]{cc}U_{11}&U_{12}\\ U_{21}&U_{22}\\ \end{array}\right)\left(\begin{array}[]{cc}\lambda_{1}\sigma_{0}&0_{2\times 2}\\ 0_{2\times 2}&\sigma_{0}/\lambda_{1}^{*}\\ \end{array}\right), (D15)

where UU is made up of eigenvectors of H1H_{1} and H2H_{2}. Then, the fixed-point boundary Green’s function is given by

G(E=0)=limNGN(E=0)=U21(U11)1(V0)1,G(E=0)=\lim_{N\rightarrow\infty}G_{N}(E=0)=U_{21}(U_{11})^{-1}(V_{0}^{\dagger})^{-1}, (D16)

where V0V_{0} can be chosen as any invertible matrix. By calculating eigenvectors of H1H_{1} and H2H_{2}, we obtain the effective boundary Hamiltonian along xx

Heff(kx)=G(E=0)1=g(kx)f(kx)Hx(kx,mx),H_{eff}(k_{x})=-G(E=0)^{-1}=\frac{g(k_{x})}{f(k_{x})}H_{x}(k_{x},m_{x}), (D17)

where g(kx)=(myλ11)g(k_{x})=-(m_{y}\lambda_{1}-1) and f(kx)=1+mx2+2mxsinkx(1|mx|)2>0f(k_{x})=1+m_{x}^{2}+2m_{x}\sin k_{x}\geq(1-|m_{x}|)^{2}>0 (mx=1m_{x}=1 is not considered as it corresponds to a phase boundary). Let us further prove that g(kx)<0g(k_{x})<0 for all kxk_{x}. Suppose my>0m_{y}>0, then λ1=12my[A(kx)+1+(A(kx)+1)24my2]\lambda_{1}=\frac{1}{2m_{y}}[A(k_{x})+1+\sqrt{(A(k_{x})+1)^{2}-4m_{y}^{2}}], we have

2g(kx)\displaystyle 2g(k_{x}) =\displaystyle= A(kx)+1(A(kx)+1)24my2\displaystyle-A(k_{x})+1-\sqrt{(A(k_{x})+1)^{2}-4m_{y}^{2}} (D18)
\displaystyle\leq (my2+mx22|mx|)\displaystyle-(m_{y}^{2}+m_{x}^{2}-2|m_{x}|)
[1+my2+(1mx)2]4my2\displaystyle-\sqrt{[1+m_{y}^{2}+(1-m_{x})^{2}]-4m_{y}^{2}}
<\displaystyle< (my2+mx22|mx|)\displaystyle-(m_{y}^{2}+m_{x}^{2}-2|m_{x}|)
|1my2+(1|mx|)2|,\displaystyle-|1-m_{y}^{2}+(1-|m_{x}|)^{2}|,

where we have used A(kx)(1|mx|)2+my2A(k_{x})\geq(1-|m_{x}|)^{2}+m_{y}^{2}. If 1my2+(1|mx|)2>01-m_{y}^{2}+(1-|m_{x}|)^{2}>0, then g(kx)<(1|mx|)2<0g(k_{x})<-(1-|m_{x}|)^{2}<0; otherwise, we have my2+mx22mx>2(1|mx|)2>0m_{y}^{2}+m_{x}^{2}-2m_{x}>2(1-|m_{x}|)^{2}>0, giving g(kx)<0g(k_{x})<0. Similarly,

Heff(ky)=g¯(ky)f¯(ky)Hy(ky,my),H_{eff}(k_{y})=\frac{\bar{g}(k_{y})}{\bar{f}(k_{y})}H_{y}(k_{y},m_{y}), (D19)

where g¯(ky)=(mxλ2+1)\bar{g}(k_{y})=-(m_{x}\lambda_{2}+1) and f¯(ky)=1+my2+2mysinky\bar{f}(k_{y})=1+m_{y}^{2}+2m_{y}\sin k_{y} with λ2=12mx[B(ky)+1+(B(ky)+1)24mx2]\lambda_{2}=-\frac{1}{2m_{x}}[B(k_{y})+1+\sqrt{(B(k_{y})+1)^{2}-4m_{x}^{2}}] and B(ky)=1+mx2+my2+2mysinkyB(k_{y})=1+m_{x}^{2}+m_{y}^{2}+2m_{y}\sin k_{y}.

Evidently, the higher-order topological phase arises when these effective boundary Hamiltonians become topological and thus can be characterized by the topological invariant PP.

Appendix E: Griffiths regime

In the main text, we have shown the existence of a Griffiths phase where topologically nontrivial and trivial samples coexist, leading to the topological invariant PP that is not quantized. In Fig. D1, we plot the polarizations in 200200 different iteration steps corresponding to different sample configurations. We see that in the Griffiths regime, some results show the polarization of 0.50.5 and others zero.

Appendix F: The finite-size analysis of the LSR and IPR

Refer to caption
Figure F1: (Color online) (a) The LSR with respect to the system size LL for W=1W=1 and (b) the IPR with respect to the system size LL for W=3.2W=3.2 for different energy levels around zero energy. The inset shows the fractal dimension D2D_{2} averaged over these energy levels as a function of LL. Here mx=my=1.1m_{x}=m_{y}=1.1.

In the main text, we have shown that the LSR at the band edge around zero energy in the region around W=1W=1 is close to 0.3860.386, indicating that the states are localized. Here we further plot the LSR for W=1W=1 with respect to the system size in Fig. F1(a), illustrating that the LSR approaches 0.3860.386 as the system size increases. We have also shown in the main text that for the localized states, the fractal dimension D2D_{2} can take negative values due to finite-size effects. Here we plot the IPR with respect to the system size in Fig. F1(b), showing the increase of the IPR with respect to the system size for a system with moderate sizes. Such an increase gives a negative fractal dimension. Yet, the increase slope declines as the system size is raised, indicating that D2D_{2} approaches zero in the thermodynamic limit.

References

  • (1) W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Quantized electric multipole insulators, Science 357, 61 (2017).
  • (2) M. Sitte, A. Rosch, E. Altman, and L. Fritz, Topological Insulators in Magnetic Fields: Quantum Hall Effect and Edge Channels with a Nonquantized θ\theta Term, Phys. Rev. Lett. 108, 126807 (2012).
  • (3) F. Zhang, C. L. Kane, and E. J. Mele, Surface State Magnetization and Chiral Edge States on Topological Insulators, Phys. Rev. Lett. 110, 046404 (2013).
  • (4) R.-J. Slager, L. Rademaker, J. Zaanen, and L. Balents, Impurity-bound states and Green’s function zeros as local signatures of topology, Phys. Rev. B 92, 085126 (2015).
  • (5) 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).
  • (6) 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).
  • (7) 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 (2018).
  • (8) L. Trifunovic and P. W. Brouwer, Higher-order bulk-boundary correspondence for topological crystalline phases, Phys. Rev. X 9, 011012 (2019).
  • (9) D. Călugăru, V. Juričić, and B. Roy, Higher-order topological phases: A general principle of construction, Phys. Rev. B 99, 041301(R) (2019).
  • (10) 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).
  • (11) B. Huang and W. V. Liu, Floquet Higher-Order Topological Insulators with Anomalous Dynamical Polarization, Phys. Rev. Lett. 124, 216601 (2020).
  • (12) H. Hu, B. Huang, E. Zhao, and W. V. Liu, Dynamical Singularities of Floquet Higher-Order Topological Insulators, Phys. Rev. Lett. 124, 057001 (2020).
  • (13) Y.-B. Yang, K. Li, L.-M. Duan, and Y. Xu, Type-II quadrupole topological insulators, Phys. Rev. Res. 2, 033029 (2020).
  • (14) A. Tiwari, M.-H. Li, B. A. Bernevig, T. Neupert, and S. A. Parameswaran, Unhinging the Surfaces of Higher-Order Topological Insulators and Superconductors, Phys. Rev. Lett. 124, 046801 (2020).
  • (15) H. Araki, T. Mizoguchi, and Y. Hatsugai, Phase diagram of a disordered higher-order topological insulator: A machine learning study, Phys. Rev. B 99, 085406 (2019).
  • (16) Z. Su, Y. Kang, B. Zhang, Z. Zhang, and H. Jiang, Disorder induced phase transition in magnetic higher-order topological insulator: A machine learning study, Chin. Phys. B 28, 117301 (2019).
  • (17) S. Franca, D. V. Efremov, and I. C. Fulga, Phase-tunable second-order topological superconductor, Phys. Rev. B 100, 075415 (2019).
  • (18) C.-A. Li and S.-S. Wu, Topological states in generalized electric quadrupole insulators, Phys. Rev. B 101, 195309 (2020).
  • (19) A. Agarwala, V. Juričić, and B. Roy, Higher-order topological insulators in amorphous solids, Phys. Rev. Res. 2, 012067(R) (2020).
  • (20) C. Wang and X. R. Wang, Disorder-induced quantum phase transitions in three-dimensional second-order topological insulators, Phys. Rev. Res. 2, 033521 (2020).
  • (21) F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008).
  • (22) J. Li, R.-L. Chu, J. K. Jain, and S.-Q. Shen, Topological Anderson Insulator, Phys. Rev. Lett. 102, 136806 (2009).
  • (23) C. W. Groth, M. Wimmer, A. R. Akhmerov, J. Tworzydło, and C. W. J. Beenakker, Theory of the Topological Anderson Insulator, Phys. Rev. Lett. 103, 196805 (2009).
  • (24) H. Jiang, L. Wang, Q.-F. Sun, and X. C. Xie, Numerical study of the topological Anderson insulator in HgTe/CdTe quantum wells, Phys. Rev. B 80, 165316 (2009).
  • (25) H.-M. Guo, G. Rosenberg, G. Refael, and M. Franz, Topological Anderson Insulator in Three Dimensions, Phys. Rev. Lett. 105, 216601 (2010).
  • (26) 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).
  • (27) 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).
  • (28) P. Titum, N. H. Lindner, M. C. Rechtsman, and G. Refael, Disorder-Induced Floquet Topological Insulators, Phys. Rev. Lett. 114, 056801 (2015).
  • (29) C. Liu, W. Gao, B. Yang, and S. Zhang, Disorder-Induced Topological State Transition in Photonic Metamaterials, Phys. Rev. Lett. 119, 183901 (2017).
  • (30) C.-Z. Chen, J. Song, H. Jiang, Q.-F. Sun, Z. Wang, and X. C. Xie, Disorder and Metal-Insulator Transitions in Weyl Semimetals, Phys. Rev. Lett. 115, 246603 (2015).
  • (31) S. Stützer, Y. Plotnik, Y. Lumer, P. Titum, N. H. Lindner, M. Segev, M. C. Rechtsman, and A. Szameit, Photonic topological Anderson insulators, Nature (London) 560, 461 (2018).
  • (32) 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).
  • (33) A. Altland and M. R. Zirnbauer, Nonstandard symmetry classes in mesoscopic normal-superconducting hybrid structures, Phys. Rev. B 55, 1142 (1997).
  • (34) C. W. J. Beenakker, Random-matrix theory of quantum transport, Rev. Mod. Phys. 69, 731 (1997).
  • (35) F. Haake, Quantum Signatures of Chaos (Springer-Verlag, Berlin, Heidelberg, 2006).
  • (36) 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).
  • (37) F. J. Dyson, The Dynamics of a Disordered Linear Chain, Phys. Rev. 92, 1331 (1953).
  • (38) G. Theodorou and M. H. Cohen, Extended states in a one-demensional system with off-diagonal disorder, Phys. Rev. B 13, 4597 (1976).
  • (39) T. P. Eggarter and R. Riedinger, Singular behavior of tight-binding chains with off-diagonal disorder, Phys. Rev. B 18, 569 (1978).
  • (40) A. Eilmes, R. A. Römer, and M. Schreiber, The two-dimensional Anderson model of localization with random hopping, Eur. Phys. J. B 1, 29 (1998).
  • (41) P. W. Brouwer, C. Mudry, B. D. Simons, and A. Altland, Delocalization in Coupled One-Dimensional Chains, Phys. Rev. Lett. 81, 862 (1998).
  • (42) R. Okugawa, S. Hayashi, and T. Nakanishi, Second-order topological phases protected by chiral symmetry, Phys. Rev. B 100, 235302 (2019).
  • (43) Q.-B. Zeng, Y.-B. Yang, and Y. Xu, Higher-order topological insulators and semimetals in generalized Aubry-André-Harper models, Phys. Rev. B 101, 241104(R) (2020).
  • (44) B. Kang, K. Shiozaki, and G. Y. Cho, Many-body order parameters for multipoles in solids, Phys. Rev. B 100, 245134 (2019).
  • (45) W. A. Wheeler, L. K. Wagner, and T. L. Hughes, Many-body electric multipole operators in extended systems, Phys. Rev. B 100, 245135 (2019).
  • (46) R. Resta, Quantum-Mechanical Position Operator in Extended Systems, Phys. Rev. Lett. 80, 1800 (1998).
  • (47) X. Dai, T. L. Hughes, X.-L. Qi, Z. Fang, and S.-C. Zhang, Helical edge and surface states in HgTe quantum wells and bulk insulators, Phys. Rev. B 77, 125319 (2008).
  • (48) Y. Peng, Y. Bao, and F. von Oppen, Boundary Green functions of topological insulators and superconductors, Phys. Rev. B 95, 235143 (2017).
  • (49) V. Oganesyan and D. A. Huse, Localization of interacting fermions at high temperature, Phys. Rev. B 75, 155111 (2007).
  • (50) C. Castellani and L. Peliti, Multifractal wavefunction at the localisation threshold, J. Phys. A 19, L429 (1986).
  • (51) A. MacKinnon and B. Kramer, The scaling theory of electrons in disordered solids: Additional numerical results, Z. Phys. B 53, 1 (1983).
  • (52) M. Serra-Garcia, V. Peri, R. Süsstrunk, O. R. Bilal, T. Larsen, L. G. Villanueva, and S. D. Huber, Observation of a phononic quadrupole topological insulator, Nature (London) 555, 342 (2018).
  • (53) C. W. Peterson, W. A. Benalcazar, T. L. Hughes, and G. Bahl, A quantized microwave quadrupole insulator with topologically protected corner states, Nature (London) 555, 346 (2018).
  • (54) 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).
  • (55) S. Mittal, V. V. Orre, G. Zhu, M. A. Gorlach, A. Poddubny, and M. Hafezi, Photonic quadrupole topological phases, Nat. Photon. 13, 692 (2019).
  • (56) C.-A. Li, B. Fu, Z.-A. Hu, J. Li, and S.-Q. Shen, Topological Phase Transitions in Disordered Electric Quadrupole Insulators, Phys. Rev. Lett. 125, 166801 (2020).
  • (57) W. Zhang, D. Zou, Q. Pei, W. He, J. Bao, H. Sun, and X. Zhang, Experimental Observation of Higher-Order Topological Anderson Insulators, arXiv:2008.00423.