Exact multiple anomalous mobility edges in a flat band geometry

Zhanpeng Lu , Hui Liu , Yunbo Zhang , Zhihao Xu

Front. Phys. ›› 2025, Vol. 20 ›› Issue (6) : 065201

PDF (3617KB)
Front. Phys. ›› 2025, Vol. 20 ›› Issue (6) : 065201 DOI: 10.15302/frontphys.2025.065201
RESEARCH ARTICLE

Exact multiple anomalous mobility edges in a flat band geometry

Author information +
History +
PDF (3617KB)

Abstract

Anomalous mobility edges (AMEs), separating localized from multifractal critical states, represent a novel form of localization transition in quasiperiodic systems. However, quasi-periodic models exhibiting exact AMEs remain relatively rare, limiting the understanding of these transitions. In this work, we leverage the geometric structure of flat band models to construct exact AMEs. Specifically, we introduce an anti-symmetric diagonal quasi-periodic mosaic modulation, which consists of both quasi-periodic and constant potentials, into a cross-stitch flat band lattice. When the constant potential is zero, the system resides entirely in a localized phase, with its dispersion relation precisely determined. For non-zero constant potentials, we use a simple method to derive analytical solutions for a class of AMEs, providing exact results for both the AMEs and the system’s localization and critical properties. Additionally, we propose a classical electrical circuit design to experimentally realize the system. This study offers valuable insights into the existence and characteristics of AMEs in quasi-periodic systems.

Graphical abstract

Keywords

localization / disorder / mobility tdge / flat band

Cite this article

Download citation ▾
Zhanpeng Lu, Hui Liu, Yunbo Zhang, Zhihao Xu. Exact multiple anomalous mobility edges in a flat band geometry. Front. Phys., 2025, 20(6): 065201 DOI:10.15302/frontphys.2025.065201

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

Multifractal critical states are special quantum states in disordered systems, characterized by strong amplitude fluctuations at the Anderson transition and displaying multifractal structures [17]. Moreover, recently, multifractal critical states have been observed in cold atom systems using the momentum lattice technique [8]. Since the behavior of multifractal critical states is completely different from either extended or localized states [911], they have attracted increasing attention from researchers over the past few decades. In terms of spatial distribution, the wave function of extended states is delocalized in real space but localized in the dual momentum space, whereas the opposite holds true for localized states [12]. This contrast suggests that extended and localized states can transform into each other under a position-momentum dual transformation. Multifractal critical states, in contrast, are delocalized in both real and momentum spaces [1324]. In position space, for an extended state, the amplitude between neighboring lattice sites remains constant in the thermodynamic limit, leading to a Lyapunov exponent λ1=0. In contrast, for a localized state, the amplitude decays exponentially between neighboring sites, resulting in λ1>0. In momentum space, extended states are characterized by λ1>0, while localized states correspond to λ1=0. However, multifractal critical states differ from both extended and localized states in that they exhibit zero Lyapunov exponents λ1=0 in both real space and momentum space, as emphasized in Refs. [9]. With interactions, single-particle multifractal critical states can evolve into some novel many-body quantum phases, such as a many-body critical phase [26] or a multifractal charge-density-wave order [27]. Recently, an important progress of multifractal critical states is anomalous mobility edges (AMEs) [25, 2831]. Unlike traditional MEs, which separate extended and localized states [3258], AMEs delineate energy boundaries that distinguish localized states from multifractal critical states, thereby broadening the scope of localization physics.

On the other hand, the study of flat band lattices has attracted widespread attention in multidisciplinary fields. Flat band lattices are translationally invariant systems that feature at least one dispersionless band and can be detected in phononic metamaterials [59]. Due to the freezing of electron motion in flat band systems, they offer an ideal platform for achieving exotic correlated phases or peculiar transport phenomena [6069], such as boundary flat bands with topological spin textures [70]. Recent theoretical studies have shown that introducing quasi-periodic modulations into flat band lattices can lead to exotic localization transitions, such as exact AMEs [71]. However, although some research has investigated AMEs in flat band systems, there has been limited focus on the precise engineering of exact AMEs within these lattices. This gap presents a promising direction for future research in this largely unexplored area.

In this work, we investigate a one-dimensional flat-band cross-stitch lattice subjected to diagonal mosaic quasi-periodic modulation, which incorporates both a constant potential and a quasi-periodic potential. By leveraging the unique geometric structure of flat-band models, we construct exact AMEs in such systems. Specifically, we introduce an anti-symmetric diagonal quasi-periodic mosaic modulation into the cross-stitch lattice. Through a local rotational transformation, the system acquires a structure that supports AMEs, characterized by mosaic-type quasi-periodic hopping coefficients and on-site potentials. This framework enables the analytical derivation of exact AMEs in flat-band quasicrystalline systems. Distinct from conventional quasi-periodic models where the system typically transitions from an extended phase to an intermediate phase with MEs, our model exhibits fundamentally different behavior. In the absence of constant potential modulation, all eigenstates are localized and no AMEs are observed. However, when the constant potential modulation is present, the system exhibits a class of exact AMEs. For weak quasi-periodic modulation, the system remains in a critical phase. As the strength of the quasi-periodic modulation increases, the system transitions to an intermediate phase featuring multiple exact AMEs, with multifractal critical states persisting even in the strong disorder limit. Analytically, we demonstrate that the localization properties of the system can be mapped onto a generalized AA model with both diagonal and off-diagonal modulations, which exhibits the energy-dependent transitions from multifractal to localized states. This energy dependence underlies the emergence and persistence of AMEs in our study. These findings reveal rich localization phenomena and provide new insights into the interplay between quasi-periodic modulation and AMEs in flat-band systems.

The rest of the paper is organized as follows: In Section 2, we first introduce the general model Hamiltonian of the 1D cross-stitch lattice with diagonal mosaic quasi-periodic modulation. In Section 3, we study the localization properties of the system for the mosaic quasi-periodic modulated cross-stitch lattice. An experimental scheme using electrical circuits is proposed to realize our model in Section 4. Finally, a summary is presented in Section 5.

2 Model

We consider a cross-stitch lattice [7274] with an on-site mosaic quasi-periodic modulation. The cross-stitch model, depicted in Fig.1, is defined for a unit cell comprising two sublattices, labeled A and B, within the blue dotted box region. Under the tight-binding approximation, the system is described by the Hamiltonian H^= H^0+ H^, where

H^0= Jn(c^n,A c^n+1,A+ c^n,Ac^n+1,B+H .c)+Jn( c^n,Bc^n+1,B+ c^n,Bc^n+1,A+H .c) tn( c^n,Ac^n,B+H.c),

and

H^=n(VA,n c^n,Ac^n,A+VB,n c^n,Bc^n,B),

where c^n,A( c^n, B) is the annihilation operator localized at sublattice A( B) of the n-th cell, J and t represent the intercell and intracell hopping amplitudes, respectively. Here, we set J=1 as the unit of energy. The on-site potentials on atoms A and B are considered to have an anti-symmetric form, i.e., V A,n= VB,n=Vn. The on-site mosaic modulation Vn is defined as [29, 45, 50]

Vn={Δ 1cos(2πβ n),n=κs,Δ 2, otherwise,

where the spatial modulation frequency set as the inverse of the golden mean, β =(51)/2, Δ1 and Δ2 denote the quasi-periodic and constant potential amplitudes, respectively, and κ is an integer. Since the quasi-periodic potential appears periodically with an interval of κ for the unit cell in the cross-stitch chain, we group the nearest κ unit cells into a cluster, indexed by s. If the total number of the cluster is N~, the system size will be L=2κ N~ with the number of unit cells N=κN ~. When κ= 1, H^ reduces to the standard AA potential. This AA-type modulated cross-stitch chain exhibits a critical-to-insulator transition as the strength of the quasi-periodic potentials increases at Δ1 =4 without showing any AMEs [72]. For κ 1 cases, we find that the localization properties strongly depend on whether Δ2 is present or absent.

The eigenvalue problem for the given tight-binding model is formulated as follows:

Eψn= t^ ψnT^ψn1T^ψn+1+ V^ψn,

where

t^=(0tt 0),T^=( 1111),V^=( Vn0 0 Vn).

Here, t^, T^, and V^ represent the intracell coupling, intercell coupling, and on-site potential matrices, respectively. The vector ψn= (anbn) T corresponds to the two-site wavefunction in the n-th unit cell of the periodic lattice. In the crystalline limit, where the on-site potential is set to zero, the system exhibits two flat bands with eigenvalues E±=±t2+4 [73, 74]. For such flat-band system, the lattice can be transformed into a Fano defect configuration using a local coordinate transformation [71]. Specifically, for the cross-stitch lattice, the transformation involves the matrix U^:

(pn fn)=U^ (an bn),U^=1 2 (111 1).

After this transformation, the eigenvalue equation becomes

(E+ t)p n=2 fn 1+ Vnf n,(E t) fn= 2p n+1+Vnpn.

Eliminating fn in favor of pn, we arrive at a reduced tight-binding form for pn:

B0p n=2 Vn 1pn 12 Vnp n+1+Vn2p n

with B 0=E2t24. This formulation described a 1D mosaic-modulated chain, referred to as the P-chain. Given the quasiperiodic modulation with an interval κ, we can define a quasicell that includes the nearest κ sites of the P-chain. The total number of the quasicell is denoted by N~, and the system size of the P-chain is given by N=κN~.

To investigate the localization properties of the eigenstates, one can calculate the inverse participation ratio (IPR) [7], IPR(i) = n=1N α={A, B}|ψ n,α(i )|4, where ψn,α(i ) represents the probability amplitude of the i-th eigenstate, with eigenvalue Ei, on the α sublattice in the n-th unit cell. For an extended state, the IPR scales as the inverse of the system size and approaches zero in the thermodynamic limit. Conversely, for a localized state, the IPR remains finite and independent of the system size L. In regions where the eigenstate ψ(i) exhibits multifractal behavior, IP R(i ) L Γi with Γi (0, 1). To quantify this behavior, the fractal dimension of the i-th eigenstate ψn,α( i) is defined as [24]

Γi=limL [ lnIPR( i) ln L].

Based on this definition, we can know that Γi 1 for an extended state; Γi 0 for a localized state; and when Γi(0,1 ), the corresponding state is a multifractal critical one. Additional, the mean inverse participation ratio can be used to characterize the localization properties within a specific region, M IP R(σ ~β~)=(1 /N σ~ β~)i σ~ β~IPR( i), where β~{ E~,C~, L~} specifies the region type – extended ( σ ~E~), critical ( σ ~C~), or localized ( σ~ L~) – and N σ~ β~ denotes the total number of eigenstates in the corresponding region. The scaling behavior of MIPR can be fitted as M IP R(σ ~β~) L Γ~. For a perfectly localized region, MIPR(σ~L~) maintains a finite value that hardly changes with the system size. For a perfectly extended region, M IP R(σ ~E~) changes linearly with 1 /L. For a critical region, Γ~(0, 1). This framework provides a method to analyze the localization and multifractal properties of eigenstates across various spectral regions.

To further verify the presence of multifractal critical states in the system, we analyze the standard deviations of the coordinates (σi) and the inverse of localization length (λi1) of the eigenstates as functions of their eigenvalues. The standard deviations σi, as defined in [25], captures the spatial distribution of the eigenstate and is given by

σi= j=1L (j j¯)2 |ψi (j)|2,

where j represents the lattice coordinate, and j¯=j=1Lj |ψ(i )(j)|2 is the center of mass position of the eigenstate. The value of σi provides the information about the spatial distribution of the eigenstates. For localized (extended) states, σi is relatively small (large). However, for multifractal critical states, σi lies between these two extremes and typically exhibits fluctuations. For a given energy Ei, the inverse of localization length λi1, also referred to as the Lyapunov exponent [7578], is another critical metric for characterizing eigenstates. It is defined as [7981]

λi1=lim N1Nln | | T(E i)||,

where T(Ei) denotes the norm of the total transfer matrix, T(Ei)= n=1NTn (Ei). According to Eq. (8), the transfer matrix for the n-th unit cell is expressed as

Tn(Ei)=( (Ei2 t2 4Vn 2) 2VnVn1 Vn10).

Here, λi1 measures the average growth rate of the wavefunction for the i-th eigenstate. A localized state corresponds to λi1>0, indicating exponential decay of the wavefunction. A delocalized state, whether extended or multifractal critical, is characterized by λi1=0. By examining the behaviors of σi and λi1, one can distinguish between localized, extended and multifractal critical states. In this paper, we apply the exact diagonalization method to numerically calculate the modulated cross-stitch chain under periodic boundary conditions (PBCs), with the intracell amplitude set to t=1.

3 Localization features

Let us first examine the case where Δ2 =0. For an arbitrary κ 2, the absence of Δ2 leads Eq. (8) to describe a series of disconnected lattice sites, which implies that all eigenstates are localized. Moreover, there exists a (κ2)N~-fold degeneracy at E=±t2+4, respectively. The degeneracy of the remaining L2(κ2)N~ energies is broken, and these energies are given by

E=±t2+4+Vκs2/2±4Vκs 2+V κs4 /4.

Fig.2(a) and (b) show the energy distributions with t=1, Δ1 =1, and L=1600 for κ=2 and κ= 4, respectively. In the case of κ= 2, there are no degenerate states, while for κ= 4, there is a 400-fold degeneracy at E=±5, respectively. The non-degenerate energies in Fig.2(a) and (b) correspond to the values predicted by Eq. (13), which agree well with our numerical results. Due to the emergence of the macroscopic degeneracy, arbitrary linear combinations of the degenerate eigenstates become solutions to the system. To mitigate the effects of this degeneracy, we introduce a small random disorder at each site in our numerical calculations. Specifically, we use a random disorder amplitude set to 108 in Fig.2(c) and (d), which plot the fractal dimension Γi as a function of eigenvalues for different system sizes for κ= 2 and κ=4, respectively. Notably, Γi approaches 0 for all states as the system size increases, indicating that the system remains in a localized phase in the Δ2=0 case.

For the case of Δ2 0, the system exhibits distinct localization properties, and the AMEs emerge. We choose κ= 2 and 3 as examples for our discussion. Results for arbitrary κ2 are provided in the Appendix.

For κ =2, the quasiperiodic potential periodically appears at even unit cells, while the modulation amplitude at the odd unit cells is set to a constant value Δ20. From Eq. (8), the effective eigenvalue equations for the P-chain can be expressed as

A1p 1,s=2V2,s1p 2,s12 Δ2 p2,s, B0p2,s= 2Δ2p1,s2V2, sp1,s+1+V2, s2p 2,s,

where A1=B 0Δ22. Here, the subscripts of pα~,s and Vα~,s denote the α~-th site in the s-th quasicell, with α~=1,2,,κ and s=1,2,,N~. By reducing Eq. (14), we obtain

( B0 4Δ22 A1)p2,s= 4Δ2A1(V2,s1p 2,s1+V2,sp2, s+1 ) +(4A1+1)V2, s2p 2,s.

This eigenvalue equation [Eq. (15)] represents a 1D generalized AA model with quasiperiodically modulated hopping and on-site terms. The system undergoes a critical-to-insulator transition when the on-site amplitude reaches twice the hopping amplitude, as discussed in Refs. [26, 31, 72]. Based on the localization transition of the generalized AA model, the analytical expressions for the AMEs are given by

Ec=± ±8Δ2 Δ1+t2+Δ 22,

which are shown as the black solid lines in Fig.3. Fig.3 illustrates the fractal dimension Γi as the function of eigenenergy Ei and Δ1 for L=1220. The yellow and blue regions correspond to critical and localized states, respectively. When the modulation amplitude Δ1 is weak, Γi for all the eigenstates deviates from 0 and 1, indicating that the system resides in a critical phase. As Δ1 increases to 1.65, a pair of AMEs emerges, marking the transition from the critical phase to a mixed phase of critical and localized states. When Δ1 4.1, another pair of AMEs emerges. In the regime of large Δ1, there are four AMEs in total, which separate critical and localized states. Notably, multifractal critical states persist even in regions with large disorder, highlighting the robustness of these states.

Fig.4(a) shows the fractal dimensions Γi as a function of eigenvalues for different system sizes in the weak disorder regime ( Δ1 =1). In this regime, the fractal dimensions for all states deviate from both 0 and 1 and exhibit minimal dependence on system size, indicating that all states are critical. Fig.4(b) and (c) show σi and the Lyapunov exponent λi1 as functions of the energy index for Δ1=1, respectively. In Fig.4(b), the large fluctuations in σi suggest the presence of multifractal critical states. Furthermore, as seen in Fig.4(c), λi1 approaches zero, confirming the delocalized nature of the eigenstates in the weak Δ1 regime. Additionally, the critical phase is explicitly confirmed by the scaling behavior of the MIPR. For Δ1 =1, all states are critical, and the corresponding fitting function is MIPR(σ~ C~) 0.1789N0.5291, as shown in Fig.4(d).

Fig.3 suggests multiple AMEs emerge when Δ1 >1.65. Fig.5(a) displays the fractal dimension Γi of various eigenstates as a function of their corresponding eigenvalues for different system sizes with large modulation amplitude Δ1 =8. The black dashed lines mark the positions of the AMEs, determined by Eq. (16). In Fig.5(a), the four AMEs partition the energy band into for five regions. The shade areas represent regions with system-size-independent Γi, indicating the emergence of multifractal critical states. In constrast, the remaining regions are localized, with Γi 0 as the system size increases. The standard deviation σi exhibits significant fluctuations in the critical regions, further confirming their critical nature, as shown in Fig.5(b). Additionally, the corresponding values of λ i1 approach 0 in these critical regions, indicating the delocalized behavior, as seen in Fig.5(c). Conversely, in the localized regions, λi1>0, and σi is very small, reflecting localized properties. The coexistence of both multifractal critical and localized states is a hallmark feature of the intermediate phase hosting AMEs. Based on the exact expressions for AMEs in Eq. (16) with κ= 2, we can distinguish the critical and localized regions in the spectrum. We further calculate the MIPR for a specific region as a function of 1/ N, as shown in Fig.5(d). One can see that the scaling of the MIPR for the two critical regions M IP R(σ ~C~)0.2779N 0.515 also exhibits critical behavior. The scalings of the MIPRs in the localized regions are shown in the insets of Fig.5(d). In these regions, the system-size-independent MI PR(σ~L~) tend to finite values in the thermodynamic limit.

For an arbitrary κ, the exact expression for AMEs (with a detailed derivation provided in the Appendix) is given by

2|(2) κ (Δ2)κ1Δ1 mκ 1Am|=| ( 4Aκ1+1)Δ12|,

where Am=A 1(2 Δ2 )2 Am1. For κ=3, Eq. (17) simplifies to

Ec=± 2+t2+Δ 22± 4+4Δ22±16Δ1Δ22.

Multiple AMEs emerge in Fig.6(a) marked by the black solid lines, where Γi is shown as a function of the eigenenergies and Δ1 for L=1260. As predicted by the analytical results, Γi approximately changes from 0 to 0.5 when the energies across these red lines. To illustrate the intermediate phase with AMEs, we take Δ1 =6 as a specific example. In Fig.6(b) and (c), we show the values of σi and λi for different eigenstates, respectively. The black dash lines represent the AMEs given by Eq. (18). In the critical regions, σi exhibits significant fluctuation, and the corresponding values of λ i1 approach 0, indicating delocalized behavior. In contrast, in the localized regions, λ i1>0 and σi is very small, reflecting the localized properties. Moreover, for any κ, one can obtain 4(κ1) AMEs described by Eq. (17), and the critical regions will persist even in the large disorder regime for κ2.

To clearly demonstrate the robustness of AMEs even if flat lattice lattice structure is disrupted, we introduce small random perturbations in intracell hopping and plot the fractal dimension Γi of different eigenstates as a function of the corresponding Ei and Δ1 or (a) κ= 2 and (b) κ=3 in Fig.7(a) and (b). Specifically, we use a random disorder amplitude set to 104 in Fig.7. Due to the emergence of the random perturbations, the ideal flat lattice structure will be destroyed. From Fig.7, one can see that the AMEs still exist in the system. However, the analytical solution of AMEs can no longer accurately determine their positions, indicating that the random perturbation causes a shift in the positions of AMEs within the system. These results indicate that AMEs are robust against small perturbations.

4 Experimental realization

This mosaic modulated cross-stitch lattice with multiple AMEs can be experimentally realized in electrical circuits. We designed a electrical circuit in Fig.8(a), corresponding to the model described by H^= H^0+ H^. In Fig.8(a), the intracell and the intercell hopping amplitudes can be controlled by the capacitor Ct, CJ, respectively. We can use a two-terminal configuration to achieve the negative capacitor value Ct, which consists of two capacitors and two operational amplifiers [40, 55, 56], shown in Fig.8(b). The inductor L isused to tune the resonant frequency of the circuit, and the on-site potentials at each lattice site is simulated by grounding capacitors Cna,Cnb. The whole system can be represented by the circuit Laplacian J(ω) of the circuit. The Laplacian is defined as the response of the grounded-voltage vector V to the vector I of input current by [40, 82]

I(ω)= J(ω)V(ω).

Using Eq. (19), the current of each node within the unit cell can be expressed as

Ina= iω CtI2Vnb+ iωCJ( 01 10)[Vn+ 1a+ Vn+1b+Vn 1a]+iω CJI2Vn 1b+[1iω L( 11 1 1) iω ( Cn,a+Ct+4 CJ ) I2]Vna

and

I nb= iω CtI2Vna+ iωCJ( 01 10)Vn 1a+iω CJI2[Vn+ 1b +Vn+ 1a+ Vn1b ]+ [1iω L( 11 1 1)iω (C n,b+ Ct+4 CJ)I2 ]Vnb,

where the vector Inα ¯=(I n,1α ¯,In,2α¯)T and Vnα¯=( Vn,1α ¯,Vn,2α ¯ ) T with α¯={a,b } represent the node currents and voltages of a and b sublattices within the nth unit cell, respectively, and I2 represents the 2× 2 identity matrix. We transform Eqs. (20) and (21) using a unitary matrix

U= 12( 111 1) ,

and achieve the transformed current−voltage relationship

I~na= iωCtI2 V~nb+iωCJ( 10 01)( V~n+1a + V~n+1b+ V~n1a)+iω CJI2 V~n1b+[1iωL (000 2) iω(Cn, a+C t+4C J)I2] V~na

and

I~nb= iω CtI2 V~na+iω CJ( 100 1) V~n1a+ iωCJI2 [ V~n+1b+ V~n+1a + V~n1b ]+[1iω L( 0002) iω (C n,b+Ct+ 4CJ)I2 ] V~nb.

As shown in Eqs. (23) and (24), the intercell hopping J appears in the second node of the circuit. We rewrite the current−voltage equations for the second nodes as

I~n,2a =iω Ct V~n,2b iωCJ( V~n+1,2a+ V~n+1,2b+ V~n1,2a )+ iω CJ V~n1,2b+iω [2ω 2L (C n,a+Ct+ 4CJ)] V~n,2a

and

I~n,2b =iω Ct V~n,2a iωCJV~n1 ,2a+ iω CJ[ V~n+1,2b+ V~n+1,2a+ V~n1 ,2b]+iω[2ω2 L(C n,b+Ct +4 CJ)] V~n,2b ,

where I~n,2α¯ and V~n,2α¯ denote the current and voltage of the second node within each sublattice in circuit, respectively. Therefore, we obtain the targeted circuit Laplacian J(ω) to simulate the our model H^= H^0 +H^ as

J(ω)=iω( C1,aCt CJC J00CtC 1,bCJ CJ 00 CJ CJ C2,aCt CJ 0CJ CJ Ct C2, b CJ 0 0 000Cn,aCt 00 00CtC n,b) +( 2ω2LCt4CJ) I2n,

where I2n represents the 2n×2n identity matrix. Hence the Hamiltonian for the cross-stitch flat band lattice with an on-site mosaic quasi-periodic modulation is achieved. The energy spectrum of the system can be obtained from the admittance spectrum of the circuit, and the distribution of states can be detected by measuring the voltage at each node [55, 56, 82, 83].

5 Summary

In this study, we investigated a class of exact AMEs in a cross-stitch lattice with quasi-periodic mosaic modulations. When Δ2=0, the system remains in a localized phase regardless of the disorder strength. However, when Δ2 0, we analytically demonstrate the existence of AMEs and derive their expressions, which show excellent agreement with numerical results. These AMEs are symmetrically distributed in energy spectrum and persist even in the presence of strong disorder. Moreover, we propose an experimental realization using electrical circuits, providing a practical platform for realizing our system. Our findings offer new insights into AMEs in flat band lattice.

References

[1]

B. Huckestein, Scaling theory of the integer quantum Hall effect, Rev. Mod. Phys. 67(2), 357 (1995)

[2]

A. Jagannathan, The Fibonacci quasicrystal: Case study of hidden dimensions and multifractality, Rev. Mod. Phys. 93(4), 045001 (2021)

[3]

S. Ostlund, R. Pandit, D. Rand, H. J. Schellnhuber, and E. D. Siggia, One-dimensional Schrödinger equation with an almost periodic potential, Phys. Rev. Lett. 50(23), 1873 (1983)

[4]

R. Merlin, K. Bajema, R. Clarke, F. Y. Juang, and P. K. Bhattacharya, Quasiperiodic GaAs–AlAs heterostructures, Phys. Rev. Lett. 55(17), 1768 (1985)

[5]

E. Maciá, Physical nature of critical modes in Fibonacci quasicrystals, Phys. Rev. B 60(14), 10032 (1999)

[6]

J. A. Ashraff and R. B. Stinchcombe, Nonuniversal critical dynamics on the Fibonacci-chain quasicrystal, Phys. Rev. B 40(4), 2278 (1989)

[7]

F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80(4), 1355 (2008)

[8]

T. Xiao, D. Xie, Z. Dong, T. Chen, W. Yi, and B. Yan, Observation of topological phase with critical localization in a quasi-periodic lattice, Sci. Bull. (Beijing) 66(21), 2175 (2021)

[9]

T. Liu and X. Xia, Predicted critical state based on invariance of the Lyapunov exponent in dual spaces, Chin. Phys. Lett. 41, 017102 (2024)

[10]

V. Goblot, A. Štrkalj, N. Pernet, J. L. Lado, C. Dorow, A. Lemaıtre, L. Le Gratiet, A. Harouri, I. Sagnes, S. Ravets, A. Amo, J. Bloch, and O. Zilberberg, Emergence of criticality through a cascade of delocalization transitions in quasiperiodic chains, Nat. Phys. 16(8), 832 (2020)

[11]

L. J. Zhai, G. Y. Huang, and S. Yin, Cascade of the delocalization transition in a non-Hermitian interpolating Aubry–André–Fibonacci chain, Phys. Rev. B 104(1), 014202 (2021)

[12]

Q. Dai, Z. Lu, and Z. Xu, Emergence of multifractality through cascade-like transitions in a mosaic interpolating Aubry–André–Fibonacci chain, Phys. Rev. B 108(14), 144207 (2023)

[13]

T. Liu, Dual-space invariance as a definitive signature of critical states in Anderson localization, arXiv: 2024)

[14]

A. Lagendijk, B. Tiggelen, and D. S. Wiersma, Fifty years of Anderson localization, Phys. Today 62(8), 24 (2009)

[15]

X. Lin, X. Chen, G. C. Guo, and M. Gong, General approach to the critical phase with coupled quasiperiodic chains, Phys. Rev. B 108(17), 174206 (2023)

[16]

S. Roy, T. Mishra, B. Tanatar, and S. Basu, Reentrant localization transition in a quasiperiodic chain, Phys. Rev. Lett. 126(10), 106803 (2021)

[17]

A. Padhan, M. K. Giri, S. Mondal, and T. Mishra, Emergence of multiple localization transitions in a one-dimensional quasiperiodic lattice, Phys. Rev. B 105(22), L220201 (2022)

[18]

Z. W. Zuo and D. Kang, Reentrant localization transition in the Su–Schrieffer–Heeger model with random-dimer disorder, Phys. Rev. A 106(1), 013305 (2022)

[19]

S. Aditya, K. Sengupta, and D. Sen, Periodically driven model with quasiperiodic potential and staggered hopping amplitudes: Engineering of mobility gaps and multifractal states, Phys. Rev. B 107(3), 035402 (2023)

[20]

S. Roy, S. Chattopadhyay, T. Mishra, and S. Basu, Critical analysis of the reentrant localization transition in a one-dimensional dimerized quasiperiodic lattice, Phys. Rev. B 105(21), 214203 (2022)

[21]

L. Z. Tang, G. Q. Zhang, L. F. Zhang, and D. W. Zhang, Localization and topological transitions in non-Hermitian quasiperiodic lattices, Phys. Rev. A 103(3), 033325 (2021)

[22]

C. W. Duncan, Critical states and anomalous mobility edges in two-dimensional diagonal quasicrystals, Phys. Rev. B 109(1), 014210 (2024)

[23]

C. Yang, W. Yang, Y. Wang, and Y. Wang, Exploring multifractal critical phases in two-dimensional quasiperiodic systems, Phys. Rev. A 110(4), 042205 (2024)

[24]

Y. Wang, L. Zhang, S. Niu, D. Yu, and X. J. Liu, Realization and detection of nonergodic critical phases in an optical Raman lattice, Phys. Rev. Lett. 125(7), 073204 (2020)

[25]

Y. C. Zhang, Y. Y. Zhang, Lyapunov exponent, and mobility edges, and critical region in the generalized Aubry–André model with an unbounded quasiperiodic potential, Phys. Rev. B 105(17), 174206 (2022)

[26]

Y. Wang, C. Cheng, X. J. Liu, and D. Yu, Many-body critical phase: Extended and nonthermal, Phys. Rev. Lett. 126(8), 080602 (2021)

[27]

M. Gonçalves,B. Amorim,F. Riche, E. V. Castro,P. Ribeiro, Incommensurability enabled quasi-fractal order in 1D narrow-band moiré systems, arXiv: 2023)

[28]

T. Liu, X. Xia, S. Longhi, and L. Sanchez-Palencia, Anomalous mobility edges in one-dimensional quasiperiodic models, SciPost Phys. 12(1), 027 (2022)

[29]

X. C. Zhou, Y. Wang, T. F. J. Poon, Q. Zhou, and X. J. Liu, Exact new mobility edges between critical and localized states, Phys. Rev. Lett. 131(17), 176401 (2023)

[30]

S. Z. Li,Y. C. Zhang,Y. C. Wang,S. C. Zhang,S. L. Zhu, Z. Li, Multifractal-enriched mobility edges and emergent quantum phases in Rydberg atomic arrays, arXiv: 2025)

[31]

F. Liu, S. Ghosh, and Y. D. Chong, Localization and adiabatic pumping in a generalized Aubry–André–Harper model, Phys. Rev. B 91(1), 014108 (2015)

[32]

Y. E. Kraus, Y. Lahini, Z. Ringel, M. Verbin, and O. Zilberberg, Topological states and adiabatic pumping in quasicrystals, Phys. Rev. Lett. 109(10), 106402 (2012)

[33]

M. Segev, Y. Silberberg, and D. N. Christodoulides, Anderson localization of light, Nat. Photonics 7(3), 197 (2013)

[34]

J. Biddle and S. Das Sarma, Predicted mobility edges in one-dimensional incommensurate optical lattices: An exactly solvable model of Anderson localization, Phys. Rev. Lett. 104(7), 070601 (2010)

[35]

S. Ganeshan, J. H. Pixley, and S. Das Sarma, Nearest neighbor tight binding models with an exact mobility edge in one dimension, Phys. Rev. Lett. 114(14), 146601 (2015)

[36]

Y. Wang, L. Zhang, W. Sun, T. F. J. Poon, and X. J. Liu, Quantum phase with coexisting localized. extended, and critical zones, Phys. Rev. B 106(14), L140203 (2022)

[37]

Y. Wang, L. Zhang, Y. Wan, Y. He, and Y. Wang, Two-dimensional vertex-decorated Lieb lattice with exact mobility edges and robust flat bands, Phys. Rev. B 107(14), L140201 (2023)

[38]

Z. Lu, Z. Xu, and Y. Zhang, Exact mobility edges and topological Anderson insulating phase in a slowly varying quasiperiodic model, Ann. Phys. 534(8), 2200203 (2022)

[39]

Z. Xu, H. Huangfu, Y. Zhang, and S. Chen, Dynamical observation of mobility edges in one-dimensional incommensurate optical lattices, New J. Phys. 22(1), 013036 (2020)

[40]

H. Liu, Z. Lu, X. Xia, and Z. Xu, Fate of localization features in a one-dimensional non-Hermitian flat-band lattice with quasiperiodic modulations, New J. Phys. 26(9), 093007 (2024)

[41]

Z. Xu, X. Xia, and S. Chen, Non-Hermitian Aubry-André model with power-law hopping, Phys. Rev. B 104(22), 224204 (2021)

[42]

S. Longhi, Dephasing-induced mobility edges in quasicrystals, Phys. Rev. Lett. 132(23), 236301 (2024)

[43]

T. Liu, H. Guo, Y. Pu, and S. Longhi, Generalized Aubry–André self-duality and mobility edges in non-Hermitian quasiperiodic lattices, Phys. Rev. B 102(2), 024205 (2020)

[44]

T. Liu and H. Guo, Mobility edges in off-diagonal disordered tight-binding models, Phys. Rev. B 98(10), 104201 (2018)

[45]

S. Longhi, Resonances , mobility edges, and gap-protected Anderson localization in generalized disordered mosaic lattices, Phys. Rev. B 110(18), 184201 (2024)

[46]

H. P. Lüschen, S. Scherg, T. Kohlert, M. Schreiber, P. Bordia, X. Li, S. Das Sarma, and I. Bloch, Single-particle mobility edge in a one-dimensional quasiperiodic optical lattice, Phys. Rev. Lett. 120(16), 160404 (2018)

[47]

X. Li and S. Das Sarma, Mobility edge and intermediate phase in one-dimensional incommensurate lattice potentials, Phys. Rev. B 101(6), 064203 (2020)

[48]

J. D. Bodyfelt, D. Leykam, C. Danieli, X. Yu, and S. Flach, Flatbands under correlated perturbations, Phys. Rev. Lett. 113(23), 236403 (2014)

[49]

H. Yao, T. Giamarchi, and L. Sanchez-Palencia, Lieb-Liniger bosons in a shallow quasiperiodic potential: Bose glass phase and fractal Mott lobes, Phys. Rev. Lett. 125(6), 060401 (2020)

[50]

Y. Wang, X. Xia, L. Zhang, H. Yao, S. Chen, J. You, Q. Zhou, and X. J. Liu, One-dimensional quasiperiodic mosaic lattice with exact mobility edges, Phys. Rev. Lett. 125(19), 196604 (2020)

[51]

J. Gao, I. M. Khaymovich, A. Iovan, X. W. Wang, G. Krishna, Z. S. Xu, E. Tortumlu, A. V. Balatsky, V. Zwiller, and A. W. Elshaari, Coexistence of extended and localized states in finite-sized mosaic Wannier-Stark lattices, Phys. Rev. B 108(14), L140202 (2023)

[52]

Z. Lu, Y. Zhang, and Z. Xu, Reentrant localization transitions in a topological Anderson insulator: A study of a generalized Su−Schrieffer−Heeger quasicrystal, Front. Phys. (Beijing) 20(2), 024204 (2025)

[53]

T. Liu, S. Cheng, R. Zhang, R. Ruan, and H. Jiang, Invariable mobility edge in a quasiperiodic lattice, Chin. Phys. B 31(2), 027101 (2022)

[54]

T. Liu and S. Cheng, Mobility edges generated by the non-Hermitian flatband lattice, Chin. Phys. B 32(2), 027102 (2023)

[55]

C. H. Xu, Z. C. Xu, Z. Y. Zhou, E. H. Cheng, and L. J. Lang, Electrical circuit simulation of non-Hermitian lattice models, Acta Phys. Sin. 72(20), 200301 (2023)

[56]

E. H. Cheng and L. J. Lang, Electrical circuit simulation of nonreciprocal Aubry–André models, Acta Phys. Sin. 71(16), 160301 (2022)

[57]

D. S. Borgnia, A. Vishwanath, and R. J. Slager, Rational approximations of quasiperiodicity via projected Green’s functions, Phys. Rev. B 106, 054204 (2022)

[58]

D. S. Borgnia and R. J. Slager, Localization as a consequence of quasiperiodic bulk-bulk correspondence, Phys. Rev. B 107, 085111 (2023)

[59]

M. M. Samak and O. R. Bilal, Direct observation of all-flat bands phononic metamaterials, Phys. Rev. Lett. 133(26), 266101 (2024)

[60]

M. Maksymenko, A. Honecker, R. Moessner, J. Richter, and O. Derzhko, Flat-band ferromagnetism as a Pauli-correlated percolation problem, Phys. Rev. Lett. 109(9), 096404 (2012)

[61]

J. W. Rhim and B. J. Yang, Singular flat bands, Adv. Phys. X 6(1), 1901606 (2021)

[62]

M. Hyrkäs, V. Apaja, and M. Manninen, Many-particle dynamics of bosons and fermions in quasi-one-dimensional flat-band lattices, Phys. Rev. A 87(2), 023614 (2013)

[63]

O. Derzhko, J. Richter, A. Honecker, M. Maksymenko, and R. Moessner, Low-temperature properties of the Hubbard model on highly frustrated one-dimensional lattices, Phys. Rev. B 81(1), 014421 (2010)

[64]

M. Goda, S. Nishino, and H. Matsuda, Inverse Anderson transition caused by flatbands, Phys. Rev. Lett. 96(12), 126401 (2006)

[65]

S. D. Huber and E. Altman, Bose condensation in flat bands, Phys. Rev. B 82(18), 184502 (2010)

[66]

D. Green, L. Santos, and C. Chamon, Isolated flat bands and spin-1 conical bands in two-dimensional lattices, Phys. Rev. B 82(7), 075104 (2010)

[67]

X. Li, J. Liu, and T. Liu, Localization–delocalization transitions in non-Hermitian Aharonov–Bohm cages, Front. Phys. (Beijing) 19(3), 33211 (2024)

[68]

Z. Ma, W. J. Chen, Y. Chen, J. H. Gao, and X. C. Xie, Flat band localization due to self-localized orbital, Front. Phys. (Beijing) 18(6), 63302 (2023)

[69]

Z. Ma, S. Li, M. M. Xiao, Y. W. Zheng, M. Lu, H. Liu, J. H. Gao, and X. C. Xie, Moiré flat bands of twisted few-layer graphite, Front. Phys. (Beijing) 18(1), 13307 (2023)

[70]

Y. Biao, Z. Yan, and R. Yu, Experimental observation of boundary flat bands with topological spin textures, Phys. Rev. B 110(24), L241110 (2024)

[71]

C. Danieli, J. D. Bodyfelt, and S. Flach, Flat-band engineering of mobility edges, Phys. Rev. B 91(23), 235134 (2015)

[72]

S. Lee, A. Andreanov, and S. Flach, Critical-to-insulator transitions and fractality edges in perturbed flat bands, Phys. Rev. B 107(1), 014204 (2023)

[73]

W. Maimaiti, A. Andreanov, H. C. Park, O. Gendelman, and S. Flach, Compact localized states and flat-band generators in one dimension, Phys. Rev. B 95(11), 115135 (2017)

[74]

W. Maimaiti and A. Andreanov, Non-Hermitian flat-band generator in one dimension, Phys. Rev. B 104(3), 035115 (2021)

[75]

A. Avila, Global theory of one-frequency Schrödinger operators, Acta Math. 215(1), 1 (2015)

[76]

A. Avila, J. You, and Q. Zhou, Sharp phase transitions for the almost Mathieu operator, Duke Math. J. 166(14), 2697 (2017)

[77]

Y. Liu, Q. Zhou, and S. Chen, Localization transition. spectrum structure.and winding numbers for one-dimensional non-Hermitian quasicrystals, Phys. Rev. B 104(2), 024201 (2021)

[78]

Y. Liu, Y. Wang, X. J. Liu, Q. Zhou, and S. Chen, Exact mobility edges PT-symmetry breaking and skin effect in one-dimensional non-Hermitian quasicrystals, Phys. Rev. B 103(1), 014203 (2021)

[79]

Z. H. Wang, F. Xu, L. Li, D. H. Xu, and B. Wang, Topological superconductors and exact mobility edges in non-Hermitian quasicrystals, Phys. Rev. B 105(2), 024514 (2022)

[80]

A. MacKinnon and B. Kramer, The scaling theory of electrons in disordered solids: Additional numerical results, Z. Phys. B 53(1), 1 (1983)

[81]

Y. Liu, Y. Wang, Z. Zheng, and S. Chen, Exact non-Hermitian mobility edges in one-dimensional quasicrystal lattice with exponentially decaying hopping and its dual lattice, Phys. Rev. B 103(13), 134208 (2021)

[82]

T. Hofmann, T. Helbig, C. H. Lee, M. Greiter, and R. Thomale, Chiral voltage propagation and calibration in a topolectrical Chern circuit, Phys. Rev. Lett. 122(24), 247702 (2019)

[83]

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(9), 925 (2018)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3617KB)

1154

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/