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

Disordered Parity Anomalous Semimetal

Shi-Hao Bi Department of Physics, The University of Hong Kong, Pokfulam Road, Hong Kong, China    Bo Fu [email protected] School of Sciences, Great Bay University, Dongguan 523000, China Great Bay Institute for Advanced Study, Dongguan 523000, China    Shun-Qing Shen [email protected] Department of Physics, The University of Hong Kong, Pokfulam Road, Hong Kong, China Quantum Science Center of Guangdong-Hong Kong-Macau Greater Bay Area, Shenzhen, China
Abstract

The parity anomalous semimetal is a topological state of matter characterized by its semi-metallic nature and a quantum Hall conductance of one-half e2/he^{2}/h (ee is the elementary charge and hh is the Planck constant). Here we investigate the topological phase transition driven by disorder in a semi-magnetic structure of topological insulator, and narrow-gap weak topological insulator film. We demonstrate that strong disorder not only leads to a topological transition from the parity anomalous semimetal to a diffusive metal with non-quantized anomalous Hall conductance, but also induces the topological phase in a trivial insulating phase. Our calculations of the local density of states provide clear picture for the formation of a single gapless Dirac fermion, which emerges as the disorder strength increases. The half quantized Hall effect is attributed to the existence of the gapless Dirac cone. Our findings of disorder-induced parity anomalous semimetal and diffusive metal significantly advance our understanding of the disorder-driven topological phase transition in magnetic topological insulators, opening up new avenues for further exploration in the field of quantum materials.

Introduction

The anomalous Hall effect and its quantized version are well-known phenomena in ferromagnets, arising from the spin-orbit coupling [1, 2, 3]. In metallic ferromagnets, the Hall conductivity is typically non-quantized and can be expressed as an integral of the Berry curvature over the occupied electron states [4]. Recent studies have proposed the possible realziation of one half quantization of the Hall conductivity in systems with gapless Wilson fermions, which can be regarded as a topological invariant for a parity anomalous semimetal (PAS) [5]. It was reported that the measured Hall conductivity approaches one-half in a semi-magnetic structure of Cr-doped topological insulator (Bi,Sb)2Te3\mathrm{(Bi,Sb)_{2}\mathrm{Te}_{3}} [6]. The system exhibits a band structure with a single gapless surface Dirac cone of electrons in the first Brillouin zone, which gives rise to the half quantization of the Hall conductivity [7, 8, 9]. This phenomenon bears a similarity to the parity anomaly of massless Dirac fermions in quantum field theory [10, 11, 12, 13]. Due to existence of a finite Fermi surface and nonzero longitudinal conductivity, the PAS is apparently distinct from the quantum anomalous Hall effect and fractional quantum anomalous Hall effect observed in an insulating phase, which are characterized by the Chern numbers and emergence of the chiral edge states [14, 15, 16, 17, 18, 19, 20, 21, 22]. Consequently, there has been a significant research effort to comprehend the origins of this effect, with numerous studies focusing on topics such as the realization, robustness, and dissipative properties of the half quantized Hall effect [23, 24, 25, 9, 26, 27].

The stability and robustness of this new topological phase in the presence of disorder pose significant challenges. Strong electron scattering near a finite Fermi surface, combined with the broken time-reversal symmetry in metallic ferromagnets, can have a profound impact [28]. Disorder in two dimensions is known to give rise to remarkable phenomena, including the metal-insulator transition [29, 30, 31], quantum Hall effect [32], and topological Anderson insulator [33, 34]. In addition to causing the metal-insulator transition, disorder also plays a crucial role in the creation of chiral edge states in topological phases [33, 34, 35, 36, 37, 38, 39, 40, 41, 42]. Therefore, understanding the effects of disorder on the stability and formation of the PAS is of utmost importance. In this Letter, we investigate the disorder-driven topological phase transition in a semi-magnetic structure of topological and narrow-gap band insulator (as depicted in Fig. 1(a)) to explore the fate and the formation of of the PAS in the presence of disorder. The phase diagram is established by calculating the Hall conductivity on a real space lattice and the arithmetical and geometric mean density of states (DOS) numerically. It is found that strong disorder not only drives a topological transition from the PAS to a diffusive metallic (DM) phase, but also strikingly induces the topological phase from an ordinary band insulating (BI) phase, as exhibited in Fig. 1(b). The half quantized Hall conductance is attributed to the emergence of a single gapless Dirac cone even in the presence of disturbing disorder, provided that the energy broadening does not smear the gap between the gapless and massive Dirac cone. Furthermore, an effective medium theory is developed to understand the formation and breaking down of the PAS.

Model and phase diagram

Consider a semi-magnetic structure of topological insulator as shown in Fig. 1(a), in which the magnetic ions are doped on the top layer of the film to form a ferromagnetic layer. The tight-binding model was introduced for topological insulator and studied extensively [43, 44, 17],

H0=𝐫iΨ𝐫iM0Ψ𝐫i+𝐫i,α=x,y,z(Ψ𝐫i𝒯αΨ𝐫i+𝐞α+H.c.),H_{0}=\sum_{{\bf r}_{i}}\Psi_{{\bf r}_{i}}^{\dagger}M_{0}\Psi_{{\bf r}_{i}}+\sum_{{\bf r}_{i},\alpha=x,y,z}(\Psi_{{\bf r}_{i}}^{\dagger}\mathcal{T}_{\alpha}\Psi_{{\bf r}_{i}+{\bf e}_{\alpha}}+\mathrm{H.c.}), (1)

where 𝒯α=tατzσ0iλα2τxσα\mathcal{T_{\alpha}}=t_{\alpha}\tau_{z}\sigma_{0}-\frac{\mathrm{i}\lambda_{\alpha}}{2}\tau_{x}\sigma_{\alpha} and M0=(m04t2tz)τzσ0+Vz(iz)τ0σzM_{0}=\left(m_{0}-4t_{\parallel}-2t_{z}\right)\tau_{z}\sigma_{0}+V_{z}(i_{z})\tau_{0}\sigma_{z}. Ψ𝐫i\Psi_{{\bf r}_{i}}^{\dagger} and Ψ𝐫i\Psi_{{\bf r}_{i}} are creation and annihilation operators of an electron at site ii. τα\tau_{\alpha} and σα\sigma_{\alpha}’s are Pauli matrices acting on the orbital and spin spaces, respectively. The magnetic doping is modeled by introducing a Zeeman potential Vz(iz)V_{z}(i_{z}). Vz(iz)=V0V_{z}(i_{z})=V_{0} for the top layer izLzMagi_{z}\leqslant L_{z}^{\mathrm{Mag}}, and 0 otherwise. The model in Eq. (1) incorporating four orbitals, |P1+,±12|P1_{-}^{+},\pm\frac{1}{2}\rangle and |P2+,12|P2_{+}^{-},\frac{1}{2}\rangle, was proposed to describe the topological nature of strong topological insulator Bi2Se3\mathrm{Bi_{2}Se_{3}} and Bi2Te3\mathrm{Bi_{2}Te_{3}} by taking m0=0.28m_{0}=0.28 eV [43, 44]. For the topologically trivial insulator, we propose the transition-metal pentatelluride ZrTe5\mathrm{ZrTe}_{5} as a possible candidate, which is commonly regarded as a weak topological insulator nearing the critical points for a transition to a strong topological insulator [45]. Notably, the low-energy states consist of four pyp_{y} orbitals from two Te atoms per unit cell, specifically |Te1/2py|\mathrm{Te}_{1/2}p_{y}\uparrow\rangle and |Te1/2py|\mathrm{Te}_{1/2}p_{y}\downarrow\rangle [46]. This case can also be effectively modeled in Eq. 1 by taking m0=0.02m_{0}=-0.02eV,.

Refer to caption
Fig. 1: (a) Schematic diagram for a semi-magnetic structure of thin films of topological insulator and/or band insulator. (b) Evolution of the Hall conductivity and the quasi-particle spectrum via disorder. The color stripes represent the energy broadening. The phase diagram of the Hall conductivity in the WEFW-E_{F} plane for (c) topological insulator and (d) narrow-gap band insulators. The bright yellow areas highlight the PAS phase, and the solid white circle lines indicate the phase boundaries by means of the effective medium theory. Parameters used are Lx=Ly=20L_{x}=L_{y}=20, Lz=10L_{z}=10, LzMag=3L_{z}^{\mathrm{Mag}}=3, λ=0.41\lambda_{\parallel}=0.41 eV, λz=0.44\lambda_{z}=0.44 eV, t=0.566t_{\parallel}=0.566 eV, tz=0.40t_{z}=0.40 eV, V0=0.1V_{0}=0.1 eV, and lattice constants a=b=1a=b=1 nm and c=0.5c=0.5 nm. 50 samples are averaged for each point.

We then study the impact of disorder on the electrical Hall conductivity of the semi-magnetic structure of topological insulator thin film. For this purpose, we follow the common practice in the study of Anderson localization and introduce disorder through random on-site energy u𝐫iu_{{\bf r}_{i}} that is uniformly distributed in [W/2,+W/2]\left[-W/2,+W/2\right], and the impurity Hamiltonian is given by Himp=𝐫iΨ𝐫iu𝐫iτ0σ0Ψ𝐫iH_{\mathrm{imp}}=\sum_{\mathbf{r}_{i}}\Psi_{{\bf r}_{i}}^{\dagger}u_{{\bf r}_{i}}\tau_{0}\sigma_{0}\Psi_{{\bf r}_{i}}. The phase diagrams of the electrical Hall conductivity in the disorder strength WW and the chemical potential EFE_{F} plane are presented in Fig. 1(c) and (d) for the topological insulator and narrow-gap band insulator, respectively. The Hall conductivity is numerically calculated by utilizing the real space noncommuative Kubo formula [47],

σxy=e2h2πiTr{P[i[x,P],i[y,P]]}imp,\sigma_{xy}=\frac{e^{2}}{h}\left\langle 2\pi\mathrm{i}\mathrm{Tr}\left\{P\left[-\mathrm{i}\left[x,P\right],-\mathrm{i}\left[y,P\right]\right]\right\}\right\rangle_{\mathrm{imp}}, (2)

where PP denotes the projector onto the occupied states, and xx and yy are the coordinate operators. imp\left\langle\cdots\right\rangle_{\mathrm{imp}} denotes the disorder-average. We take periodic boundary condition in the xx and yy directions, and open boundary condition in the zz direction. The yellow area indicates the PAS with one half quantum Hall conductivity e22h\frac{e^{2}}{2h}. For the topologically nontrivial case, the PAS phase appears from the pristine state and exhibits a gradual departure for a finite disorder WW. The deviation of Hall conductivity from the half-quantized value defines the phase boundary between PAS and DM, while a finite chemical potential can also modulate the transition disorder strength. For the topologically trivial case, with weak disorder, the system remains in the BI phase with negligible Hall conductivity, indicated by the dark-purple region in the lower left corner. With increasing disorder strength, the PAS phase emerges strikingly and persist over a finite range, distinctively signifying the existence of the metallic ferromagnetic phase. When the disorder strength is even stronger, the half-quantized Hall plateau tends to diminish, and the system eventually evolves into the DM phase. The finite-size effect analysis categorically supports the robustness of the emergent half-quantized Hall phase. For details refer to Ref. [48].

Spectral function and gapless Dirac cone

The quasiparticle picture is valid when the disorder strength is far from reaching the Anderson transition point from DM to Anderson insulator. In this situation, the disorder only renormalizes the energy spectrum of the quasiparticles and introduces a finite lifetime. Hence, by examining the evolution of the spectral function with varying the disorder strength, we can clearly observe the changes occurring during the phase transitions. The spectral function A(ϵ,𝐤)A(\epsilon,\mathbf{k}) is defined by A(ϵ,𝐤)=nψn𝐤|δ(ϵH)|ψn𝐤A(\epsilon,{\bf k})=\sum_{n}\left\langle\psi_{n{\bf k}}\right|\delta\left(\epsilon-H\right)\left|\psi_{n{\bf k}}\right\rangle, where |ψn𝐤=1S|un𝐤ei𝐤𝐫\left|\psi_{n{\bf k}}\right\rangle=\frac{1}{\sqrt{S}}\left|u_{n{\bf k}}\right\rangle\mathrm{e}^{\mathrm{i}{\bf k}\cdot{\bf r}} are the Bloch states for the clean system with H0|un𝐤=ϵn|un𝐤H_{\mathrm{0}}\left|u_{n{\bf k}}\right\rangle=\epsilon_{n}\left|u_{n{\bf k}}\right\rangle. For a disordered system, A(ϵ,𝐤)A(\epsilon,\mathbf{k}) can be numerically evaluated using the standard kernel polynomial method [52, 53], where the Wannier wavefunctions in DOS calculations are replaced by the Bloch eigenstates.

Refer to caption
Fig. 2: Evolution of the spectral function A(ϵ,𝐤)A(\epsilon,\mathbf{k}) with the disorder strength WW. The panels (a-e) are for the topological insulator, and the panels (f-j) are for the band insulator. The lattice size: Lx=Ly=400L_{x}=L_{y}=400, Lz=10L_{z}=10, and LzMag=3L_{z}^{{\rm Mag}}=3.We use M=6000M=6000 Chebyshev moments to achieve high spectral resolution. The momentum path is along the high-symmetry points MΓXM-\Gamma-X.

The spectral functions for different disorder strengths are clearly displaced in Fig. 2. Fig 2(a) shows the band structure of topologically nontrivial case in the absence of disorder W=0W=0. A gapless Dirac cone appears at the Γ\Gamma point while a series of gapped Dirac cones also accompanies. We can identify the gap of one massive Dirac cone is manipulated by the Zeeman field VzV_{z} in the magnetic layer. Thus the gapless Dirac cone is the surface states located at the bottom layer, and the gapped Dirac cone is the surface state located at the top layer of the sample. This is a straightforward evidence to illustrate the band structure is topologically nontrivial. Recent studies [5] reveal that the gapless Dirac cone of fermions gives rise to PAS as a signature of parity anomaly of gapless Dirac fermions. This is consistent with the fact that the Hall conductivity is one half as shown in Fig. 1(c). With increasing the disorder strength in PAS phase (from Fig. 2(a) to 2(c)), we notice that the gapless Dirac cone exhibits remarkable resilience to disorder, persisting even as other gapped energy dispersions are smeared. However, the broadening of these dispersions is still smaller than their respective band gap, and the Hall condcuctivity remains half-quantized. At PAS-DM transition point (Fig. 2(d)), the broadening induced by impurities reaches one half of the band gap in the gapped bands. With further increases in disorder strength, the chemical potential inevitably lies within the broadened gapped bands. Subsequently, the Hall conductivity derivates from e22h\frac{e^{2}}{2h}.

On the contrary, Fig. 2(f) shows the trivial band structure has a finite gap at the Γ\Gamma point, and is a topologically trivial band insulator in the absence of disorder. With increasing the disorder strength, the band gap progressively shrinks and ultimately closes near W=2.0W=2.0 eV, giving rise to a single, gapless Dirac cone. Correspondingly, the Hall conductivity is close to e22h\frac{e^{2}}{2h} as demonstrated in Fig. 1(d). The formation of the gapless Dirac cone is the signature of the disorder-induced PAS, which is consistent with the calculated Hall conductivity. For even stronger disorder in Fig. 2(e) or (j), the gapless Dirac cone collapses as anticipated, and the Hall conductivity deviates from half-quantization, giving rise to the DM. From the quasiparticle perspective, our core finding in this calculation can be summarized in Fig. 1(b), where in a disordered semi-magnetic topological insulator/band insulator film, the PAS phase occurs when the chemical potential solely intersects the gapless Dirac cone (despite broadening) within the broadened gapped bands.

Arithmetic and geometric mean density of states

These phase transitions can be characterized by analyzing the two types of means of the local density of states (DOS): the arithmetic mean ρa\rho_{\mathrm{a}} and the geometric mean ρt\rho_{\mathrm{t}}. The local DOS, ρ𝐫,α(ϵ)=𝐫,α|δ(ϵH)|𝐫,α\rho_{{\bf r},\alpha}(\epsilon)=\left\langle{\bf r},\alpha\right|\delta\left(\epsilon-H\right)\left|{\bf r},\alpha\right\rangle, quantifies the amplitude of the wave function at site 𝐫{\bf r} for a given energy ϵ\epsilon, where |𝐫,α\left|{\bf r},\alpha\right\rangle denotes an α\alpha-orbital electron wave function at that site. The spatial distribution of ρ𝐫,α(ϵ)\rho_{{\bf r},\alpha}(\epsilon) contains direct information about the localization properties, which are closely intertwined with the topology of the quantum system [54, 55, 56]. The arithmetic and geometric mean DOS for a disordered system are defined as ρa(ϵ)=1Vi=1Vα=14ρ𝐫i,α(ϵ)imp\rho_{\mathrm{a}}(\epsilon)=\left\langle\frac{1}{V}\sum_{i=1}^{V}\sum_{\alpha=1}^{4}\rho_{{\bf r}_{i},\alpha}(\epsilon)\right\rangle_{\mathrm{imp}} and ρt(ϵ)=exp[1Nsi=1Nslnα=14ρ𝐫i,α(ϵ)imp]\rho_{\mathrm{t}}(\epsilon)=\exp\left[\frac{1}{N_{s}}\sum_{i=1}^{N_{s}}\left\langle\ln\sum_{\alpha=1}^{4}\rho_{{\bf r}_{i},\alpha}(\epsilon)\right\rangle_{\mathrm{imp}}\right], respectively [57, 58]. Here we randomly choose a finite number of lattice sites NsV=LxLyLzN_{s}\ll V=L_{x}L_{y}L_{z} to improve the statistics of ρt\rho_{\mathrm{t}} [59]. A zero value of arithmetic mean DOS ρa(ϵ)\rho_{{\rm a}}(\epsilon) at a finite interval defines an energy gap, such as a band insulator, while a nonzero value of ρa(ϵ)\rho_{{\rm a}}(\epsilon) means that there exists the states at the energy ϵ\epsilon, which can be either localized or delocalized. The geometric mean DOS ρt(ϵ)\rho_{{\rm t}}(\epsilon) may reveal more information for the localization of the states. For energy ϵ\epsilon where all states are extended, local DOS are uniformly distributed throughout the system, and we have ρtρa\rho_{{\rm t}}\simeq\rho_{{\rm a}}. However, for the states whose wave-functions are localized in real space, there is no contribution to ρt\rho_{{\rm t}} [52]. Generally, comparison of ρa(ϵ)\rho_{\mathrm{a}}(\epsilon) to ρt(ϵ)\rho_{\mathrm{t}}(\epsilon) can investigate the localization properties of particle states. In the case of quasi-2D topological insulator films featuring surface states, the analysis of these properties becomes more complex. Further study shows that since the surface states are localized near the top or bottom surface in zz direction and extended in xx and yy direction, these states do not contribute to the value of ρt\rho_{{\rm t}}, but do to ρa\rho_{{\rm a}}. For a non-vanishing ρa\rho_{{\rm a}}, if the ratio ρt/ρa\rho_{\mathrm{t}}/\rho_{\mathrm{a}} is 0, it can define two different phases: one is that all states are localized in real space in all directions, i.e., Anderson insulator, and another one is that the states are only localized at the top or bottom surface, i.e., the surface states. For a finite ratio of ρt/ρa\rho_{\mathrm{t}}/\rho_{\mathrm{a}} it indicates the existence of the delocalized states, defining a DM.

Refer to caption
Fig. 3: The arithmetic and geometric mean DOS ρa(E)\rho_{\mathrm{a}}(E) and ρt(E)\rho_{\mathrm{t}}(E) versus disorder strength at E=0.01E=0.01 eV for (a) the topological insulator and (b) for the band insulator.The system size are the same as those in Fig. 2. M=6000M=6000 Chebyshev moments and Ns=30N_{s}=30 are used in the simulation. 20 samples are averaged for each point.

ρa\rho_{\mathrm{a}} and ρt\rho_{\mathrm{t}} can be calculated by using the standard kernel polynomial method. As illustrated in Fig. 3(a) and (b), by analyzing ρa\rho_{\mathrm{a}} and ρt\rho_{\mathrm{t}} at EF=0.01E_{F}=0.01 eV , we establish that the topologically nontrivial case undergoes two distinct quantum phase transitions: PAS\toDM\toAI (Anderson insulator) with varying disorder strength and the topologically trivial case exhibits three quantum phase transitions: BI\toPAS\toDM\toAI. For nontrivial case, when disorder strength is small (W<Wc2.60W<W_{c}\approx 2.60 eV), the gapless Dirac cone on the bottom surface experiences broadening as a result of self-scattering by impurities present on the same surface. Consequently, we observe that ρa0\rho_{{\rm a}}\neq 0 and ρt0\rho_{{\rm t}}\simeq 0, leading to a vanishingly small ratio ρt/ρa\rho_{{\rm t}}/\rho_{{\rm a}} (approximately 10110210^{-1}\sim 10^{-2} as calculated), which defines the PAS phase. Unlike in conventional metals, where ρt\rho_{{\rm t}} and ρa\rho_{{\rm a}} typically track each other, the absence of ρt\rho_{{\rm t}} provides the numerical evidence for the stability of the PAS phase under weak disorder. As disorder increases (Wc<W<Wl13.55W_{c}<W<W_{l}\approx 13.55 eV), both ρt\rho_{{\rm t}} and ρa\rho_{{\rm a}} are finite, defining the DM phase. With further increasing the disorder (W>WlW>W_{l}), all the states are expected to be Anderson localized, characterized by a vanishing ρt\rho_{t} and a finite ρa\rho_{{\rm a}} over the whole range of the spectrum. Starting from topologically trivial case and considering weak disorder (W<1.70W<1.70 eV), the presence of a finite band gap results in no states within this gap, leading to ρt=ρa=0\rho_{{\rm t}}=\rho_{{\rm a}}=0, which we refer as BI phase. Nevertheless, the numerical calculation inevitably introduce a vanishingly small broadening parameter, which effectively introduces a uniformly distributed nonzero local DOS in BI phase. This results in a ratio of ρt/ρa\rho_{{\rm t}}/\rho_{{\rm a}}\simeq1 in this regime. As disorder increases, a phase transition from BI to PAS occurs at Wc,1=2.0W_{c,1}=2.0 eV, characterized by a strongly suppressed ρt/ρa\rho_{{\rm t}}/\rho_{{\rm a}}. Continuing to increase the disorder strength leads to encountering two more quantum phase transitions, analogous to those observed in the topologically nontrivial case.

The effective medium theory

To understand the origin of the phase transitions between PAS to DM, and BI to DM, we can employ the effective medium theory in conjunction with the Kubo formula for electrical conductivities [60, 61]. The self-consistent Born approximation is a very powerful tool and applied extensively to investigate the physics in topological Anderson insulator [33, 34, 62, 63, 64, 65, 66, 37, 67, 68, 69]. In the semi-magnetic structure with multiple layers, the electron wave function has an additional layer degree of freedom, which lead to a matrix structure of the Green’s function and self-energy. We extend the self-consistent Born approximation to the layered structure, in which the retarded self energy is diagonal in the layer degree of freedom subspace and can be expressed as

ΣizizR(EF)=W212S𝐤GizizR(𝐤,EF),\Sigma_{i_{z}i_{z}}^{R}(E_{F})=\frac{W^{2}}{12S}\sum_{{\bf k}^{\prime}}G_{i_{z}i_{z}}^{R}({\bf k}^{\prime},E_{F}), (3)

where izi_{z} is the layer number, GR(𝐤,EF)=[EFH0(𝐤)ΣR]1G^{R}(\mathbf{k}^{\prime},E_{F})=\left[E_{F}-H_{\mathrm{0}}(\mathbf{k}^{\prime})-\Sigma^{R}\right]^{-1} is the dressed Green’s function. Despite the translational symmetry breaking in the zz direction, our scheme is applicable to any layer number and doping. The self-consistent equation for self-energy can be solved numerically [68]. With the self-energy, the Hall conductivity can be calculated by means of the Kubo-Bastin formula . For technique details refer to Ref.[48].

Numerical solution shows the renormalized mass m~0(iz)=m0+14Tr[ΣizizR(EF)σ0τz]\widetilde{m}_{0}\left(i_{z}\right)=m_{0}+\frac{1}{4}\mathrm{Tr}\left[\Sigma_{i_{z}i_{z}}^{R}(E_{F})\sigma_{0}\tau_{z}\right] exhibits weak layer dependence and is remarkably enhanced by disorder irrespective of its initial values. The evolution of the quasiparticle band structure shown in Fig. 2 is primarily due to this effect. For the trivial case, m~0(iz)\widetilde{m}_{0}\left(i_{z}\right) increases with increasing the disorder strength, and change its sign at a critical value, which causes a topological transition from the bulk insulating phase into a topological insulating phase, leading to the emergence of the Dirac cone of the surface states. This represents the disorder-induced topological Anderson insulator phase transition [33, 34, 62, 63, 64, 65, 66, 37, 67, 68, 69]. As shown in Fig. 4, the green circles indicate the band gap between the conduction and valence bands, extracted from the spectral function in Fig. 2. The green solid line, representing results from the effective medium theory, demonstrates good consistency with these findings. When the gap closes at Wc,1W_{c,1}, a half-quantized Hall plateau emerges (indicated by the vertical green dashed line).

Refer to caption
Fig. 4: The Hall conductivity versus disorder strength WW for a topologically trival structure . Here we take Lx=Ly=28L_{x}=L_{y}=28. The blue line is for the Hall conductivity calculated from the effective medium theory. The green line represents the effective band gap at Γ\Gamma point, as evidenced by the green empty circles directly extracted from the spectral function in Fig. 2. The orange line is the energy broadening at the top layer (iz=1i_{z}=1). Two purple pentagons at Wc,1W_{c,1} and Wc,2W_{c,2} indicate the left and right boundaries of the PAS phase.

Another important result of the effective medium theory is the imaginary part of the self-energy, η(iz)=14ImTr[ΣizizR(EF)]\eta(i_{z})=-\frac{1}{4}\mathrm{Im}\mathrm{Tr}\left[\Sigma_{i_{z}i_{z}}^{R}(E_{F})\right], which is responsible for energy level broadening and serves as an order parameter for describing the phase transition from PAS to DM. Numerical results show that imaginary part η\eta slightly depends on the layer index izi_{z}. The ηbottom\eta_{\mathrm{bottom}} on the bottom surface is always present, regardless of disorder strength, while the ηtop\eta_{\mathrm{top}} on the top (the orange line in Fig. 4) surges at a critical value Wc,23.22W_{c,2}\approx 3.22 eV of the disorder strength. It is observed that breaking down of the quantized Hall conductivity accompanies with the emergence of the band broadening near the top layer ηtop\eta_{\mathrm{top}} as indicated by the vertical orange dashed line. As the wave functions of the surface states near the crossing point mainly reside at the bottom surface, the tiny energy broadening ηbottom\eta_{\mathrm{bottom}} has no impact to the Hall conductivity. The band broadening ηtop\eta_{\mathrm{top}} on the top mains affects the higher energy states of the Dirac cone, which are mainly distributed in the whole bulk, and the are the key source to the nonzero Hall conductivity [8]. The phase boundary for the PAS-DM transition, indicated by the white dot-line in Fig. 1, is given by the emergence of ηtop\eta_{\mathrm{top}}, which shows good agreements with the independent numerical results of the real-space Hall conductivity. Therefore, the nonzero ηtop\eta_{\mathrm{top}} causes the breaking down of the PAS to DM. In the DM region, the longitudinal conductivity is finite while the Hall conductivity decays to vanish.

Conclusion

To conclude, the PAS appears in the disordered semi-magnetic structure based on numerical calculation of the Hall conductivity and local DOS, and can be understood very well in the framework of the effective medium theory. We demonstrate that strong disorder not only leads to a topological transition from the PAS to DM phase, but also drives a BI into the PAS phase. Our findings unequivocally illuminate the genesis of half-quantized Hall phase in both topologically trivial and non-trivial cases, substantiates the paradigm of disorder-driven topological phase transition, and provide a sturdy foundation for the exploration and development of half-quantized Hall effects in quantum materials and devices.

Acknowledgements.
We thank Dr. Huan-Wen Wang and Rui Chen for helpful discussions. This work was supported by the National Key R&D Program of China under Grant No. 2019YFA0308603 and the Research Grants Council, University Grants Committee, Hong Kong under Grant Nos. C7012-21G and 17301823.

References

  • [1] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Anomalous Hall effect, Rev. Mod. Phys. 82, 1539 (2010).
  • [2] D. Xiao, M.-C. Chang, and Q. Niu, Berry phase effects on electronic properties, Rev. Mod. Phys. 82, 1959 (2010).
  • [3] C.-Z. Chang, C.-X. Liu, and A. H. MacDonald, Colloquium: Quantum anomalous Hall effect, Rev. Mod. Phys. 95, 011002 (2023).
  • [4] F. D. M. Haldane, Berry Curvature on the Fermi Surface: Anomalous Hall Effect as a Topological Fermi-Liquid Property, Phys. Rev. Lett. 93, 206602 (2004).
  • [5] B. Fu, J.-Y. Zou, Z.-A. Hu, H.-W. Wang, and S.-Q. Shen, Quantum anomalous semimetals, npj Quantum Mater. 7, 94 (2022).
  • [6] M. Mogi, Y. Okamura, M. Kawamura, R. Yoshimi, K. Yasuda, A. Tsukazaki, K. Takahashi, T. Morimoto, N. Nagaosa, M. Kawasaki, et al., Experimental Signature of the Parity Anomaly in a Semi-magnetic Topological Insulator, Nat. Phys. 18, 390 (2022).
  • [7] J.-Y. Zou, B. Fu, H.-W. Wang, Z.-A. Hu, and S.-Q. Shen, Half-quantized Hall effect and power law decay of edge-current distribution Phys. Rev. B 105, L201106 (2022)
  • [8] J.-Y. Zou, R. Chen, B. Fu, H.-W. Wang, Z.-A. Hu, and S.-Q. Shen, Half-quantized Hall effect at the parity-invariant Fermi surface Phys. Rev. B 107, 125153 (2023).
  • [9] H.-W. Wang, B. Fu, and S.-Q. Shen, Signature of parity anomaly: Crossover from one half to integer quantized Hall conductance in a finite magnetic field, Phys. Rev. B 109, 075113 (2024).
  • [10] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • [11] A. N. Redlich, Parity violation and gauge noninvariance of the effective gauge field action in three dimensions, Phys. Rev. D 29, 2366 (1984).
  • [12] G. W. Semenoff, Condensed-Matter Simulation of a Three-Dimensional Anomaly, Phys. Rev. Lett. 53, 2449 (1984).
  • [13] S.-Q. Shen, Half quantized Hall effect, Coshare Science 2, 1 (2024).
  • [14] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Quantized Hall Conductance in a Two-Dimensional Periodic Potential, Phys. Rev. Lett. 49, 405 (1982).
  • [15] F. D. M. Haldane, Model for a Quantum Hall Effect without Landau Levels: Condensed-Matter Realization of the "Parity Anomaly", Phys. Rev. Lett. 61, 2015 (1988).
  • [16] R. Yu, W. Zhang, H.-J. Zhang, S.-C. Zhang, X. Dai, and Z. Fang, Quantized Anomalous Hall Effect in Magnetic Topological Insulators, Science 329, 61 (2010).
  • [17] R.-L. Chu, J. Shi, and S.-Q. Shen, Surface edge state and half-quantized Hall conductance in topological insulators, Phys. Rev. B 84, 085312 (2011).
  • [18] C.-Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y. Ou, P. Wei, L.-L. Wang, Z.-Q. Ji, Y. Feng, S. Ji, X. Chen, J. Jia, X. Dai, Z. Fang, S.-C. Zhang, K. He, Y. Wang, L. Lu, X.-C. Ma, and Q.-K. Xue, Experimental Observation of the Quantum Anomalous Hall Effect in a Magnetic Topological Insulator, Science 340, 167 (2013).
  • [19] J. G. Checkelsky, R. Yoshimi, A. Tsukazaki, K. S. Takahashi, Y. Kozuka, J. Falson, M. Kawasaki, and Y. Tokura, Trajectory of the anomalous Hall effect towards the quantized state in a ferromagnetic topological insulator, Nat. Phys. 10, 731 (2014).
  • [20] J. Cai, E. Anderson, C. Wang, X. Zhang, X. Liu, W. Holtzmann, Y. Zhang, F. Fan, T. Taniguchi, K. Watanabe, et al., Signatures of fractional quantum anomalous Hall states in twisted MoTe2, Nature 622, 63 (2023).
  • [21] H. Park, J. Cai, E. Anderson, Y. Zhang, J. Zhu, X. Liu, C. Wang, W. Holtzmann, C. Hu, Z. Liu, et al., Observation of fractionally quantized anomalous Hall effect, Nature 622, 74 (2023).
  • [22] Z. Lu, T. Han, Y. Yao, A. P. Reddy, J. Yang, J. Seo, K. Watanabe, T. Taniguchi, L. Fu, and L. Ju, Fractional quantum anomalous Hall effect in multilayer Graphene, Nature 626, 759 (2024).
  • [23] M. Gong, H. Liu, H. Jiang, C.-Z. Chen, and X.-C. Xie, Half-quantized helical hinge currents in axion insulators, National Science Review 10, nwad025 (2023).
  • [24] H. Yang, L. Song, Y. Cao, and P. Yan, Realization of Wilson fermions in topolectrical circuits, Communications Physics 6, 211 (2023).
  • [25] Z. Ning, X. Ding, D.-H. Xu, and R. Wang, Robustness of half-integer quantized Hall conductivity against disorder in an anisotropic Dirac semimetal with parity anomaly, Phys. Rev. B 108, L041104 (2023).
  • [26] Y.-H. Wan and Q.-F. Sun, Quarter-quantized thermal Hall effect with parity anomaly, Phys. Rev. B 109, 195408 (2024).
  • [27] H. Zhou, C.-Z. Chen, Q.-F. Sun, and X. C. Xie, Dissipative chiral channels, Ohmic scaling, and half-integer Hall conductivity from relativistic quantum Hall effect, Phys. Rev. B 109, 115305 (2024).
  • [28] S. Hikami, A. I. Larkin, and Y. Nagaoka, Spin-orbit interaction and magnetoresistance in the two dimensional random system, Progress of Theoretical Physics 63, 707 (1980).
  • [29] E. Abrahams, P. W. Anderson, D. C. Licciardello, and T. V. Ramakrishnan, Scaling Theory of Localization: Absence of Quantum Diffusion in Two Dimensions, Phys. Rev. Lett. 42, 673 (1979).
  • [30] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008).
  • [31] E. Abrahams, 50 Years of Anderson Localization (World Scientific Publishing Co. Pte. Ltd, 2010).
  • [32] S. M. G. Richard E. Prange, The Quantum Hall Effect, Graduate Texts in Contemporary Physics (Springer New York, NY, 2012).
  • [33] J. Li, R.-L. Chu, J. K. Jain, and S.-Q. Shen, Topological Anderson Insulator, Phys. Rev. Lett. 102, 136806 (2009).
  • [34] 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).
  • [35] S. Stützer, Y. Plotnik, Y. Lumer, P. Titum, N. H. Lindner, M. Segev, M. C. Rechtsman, and A. Szameit, Photonic Topological Anderson Insulators, Nature 560, 461 (2018).
  • [36] 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).
  • [37] 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).
  • [38] G.-G. Liu, Y. Yang, X. Ren, H. Xue, X. Lin, Y.-H. Hu, H.-x. Sun, B. Peng, P. Zhou, Y. Chong, and B. Zhang, Topological Anderson Insulator in Disordered Photonic Crystals, Phys. Rev. Lett. 125, 133603 (2020).
  • [39] W. Zhang, D. Zou, Q. Pei,W. He, J. Bao, H. Sun, and X. Zhang, Experimental Observation of Higher-Order Topological Anderson Insulators, Phys. Rev. Lett. 126, 146802 (2021).
  • [40] Q. Lin, T. Li, L. Xiao, K.Wang,W. Yi, and P. Xue, Observation of non-Hermitian topological Anderson insulator in quantum dynamics, Nat. Commun. 13, 3229 (2022).
  • [41] T. Dai, A. Ma, J. Mao, Y. Ao, X. Jia, Y. Zheng, C. Zhai, Y. Yang, Z. Li, B. Tang, et al., A programmable topological photonic chip, Nat. Mater. 23, 928 (2024).
  • [42] M. Ren, Y. Yu, B. Wu, X. Qi, Y. Wang, X. Yao, J. Ren, Z. Guo, H. Jiang, H. Chen, X.-J. Liu, Z. Chen, and Y. Sun, Realization of Gapped and Ungapped Photonic Topological Anderson Insulators, Phys. Rev. Lett. 132, 066602 (2024).
  • [43] H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, and S.-C. Zhang, Topological insulators in Bi2Se3\mathrm{Bi_{2}Se_{3}}, Bi2Te3\mathrm{Bi_{2}Te_{3}} and Sb2Te3\mathrm{Sb_{2}Te_{3}} with a single Dirac cone on the surface, Nat. Phys. 5, 438 (2009).
  • [44] C.-X. Liu, X.-L. Qi, H. Zhang, X. Dai, Z. Fang, and S.- C. Zhang, Model Hamiltonian for topological insulators, Phys. Rev. B 82, 045122 (2010).
  • [45] H. Weng, X. Dai, and Z. Fang, Transition-Metal Pentatelluride ZrTe5\mathrm{ZrTe}_{5} and HfTe5\mathrm{HfTe}_{5}: A Paradigm for Large-Gap Quantum Spin Hall Insulators, Phys. Rev. X 4, 011002 (2014).
  • [46] R. Y. Chen, Z. G. Chen, X.-Y. Song, J. A. Schneeloch, G. D. Gu, F. Wang, and N. L. Wang, Magnetoinfrared Spectroscopy of Landau Levels and Zeeman Splitting of Three-Dimensional Massless Dirac Fermions in ZrTe5\mathrm{ZrTe}_{5}, Phys. Rev. Lett. 115, 176404 (2015).
  • [47] E. Prodan, Disordered Topological Insulators: a Non-Commutative Geometry Perspective, J. Phys. A: Math. Theor. 44, 113001 (2011).
  • [48] See Supplemental Material at [URL to be added by publisher] for details of (Sec. SI) finite-size effect of the Hall conductivity in real space, (Sec. SII) features of arithmetic mean (ρa)(\rho_{{\rm a}}) and geometric mean (ρt)(\rho_{{\rm t}}) density of states in clean topological insulator film, (Sec. SIII) diffusive metallic phase to Anderson insulator transition, (Sec. SIV) the effective medium theory, and (Sec. SV) the Kubo-Bastin formula for the Hall conductivity, which includes Refs. [49, 50, 51, 60, 61].
  • [49] A. MacKinnon and B. Kramer, The scaling theory of electrons in disordered solids: Additional numerical results, Zeitschrift für Physik B Condensed Matter 53, 1 (1983).
  • [50] A. Yamakage, K. Nomura, K.-I. Imura, and Y. Kuramoto, Criticality of the metal-topological insulator transition driven by disorder, Phys. Rev. B 87, 205141 (2013).
  • [51] K. Slevin and T. Ohtsuki, Corrections to Scaling at the Anderson Transition, Phys. Rev. Lett. 82, 382 (1999).
  • [52] A. Weiße, G. Wellein, A. Alvermann, and H. Fehske, The Kernel Polynomial Method, Rev. Mod. Phys. 78, 275 (2006).
  • [53] Z. Fan, J. H. Garcia, A. W. Cummings, J. E. Barrios-Vargas, M. Panhans, A. Harju, F. Ortmann, and S. Roche, Linear Scaling Quantum Transport Methodologies, Phys. Rep. 903, 1 (2021).
  • [54] A. M. M. Pruisken, Topological Principles in the Theory of Anderson Localization, International Journal of Modern Physics B 24, 1895 (2010).
  • [55] C. Tian, Y. Chen, and J. Wang, Emergence of integer quantum Hall effect from chaos, Phys. Rev. B 93, 075403 (2016).
  • [56] S.-C. Zhang and D. P. Arovas, Effective field theory of electron motion in the presence of random magnetic flux, Phys. Rev. Lett. 72, 1886 (1994).
  • [57] Y.-Y. Zhang, R.-L. Chu, F.-C. Zhang, and S.-Q. Shen, Localization and mobility gap in the topological Anderson insulator, Phys. Rev. B 85, 035107 (2012).
  • [58] Y.-Y. Zhang and S.-Q. Shen, Algebraic and geometric mean density of states in topological Anderson insulators, Phys. Rev. B 88, 195145 (2013).
  • [59] J. H. Pixley, P. Goswami, and S. Das Sarma, Anderson Localization and the Quantum Phase Diagram of Three Dimensional Disordered Dirac Semimetals, Phys. Rev. Lett. 115, 076601 (2015).
  • [60] A. Bastin, C. Lewiner, O. Betbeder-matibet, and P. Nozieres, Quantum Oscillations of the Hall Effect of a Fermion Gas with Random Impurity Scattering, Journal of Physics and Chemistry of Solids 32, 1811 (1971).
  • [61] Y. Chen and H. Zhai, Hall conductance of a non-Hermitian Chern insulator, Phys. Rev. B 98, 245130 (2018).
  • [62] H.-M. Guo, G. Rosenberg, G. Refael, and M. Franz, Topological Anderson Insulator in Three Dimensions, Phys. Rev. Lett. 105, 216601 (2010).
  • [63] 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).
  • [64] R. Chen, D.-H. Xu, and B. Zhou, Topological Anderson Insulator Phase in a Dirac-Semimetal Thin Film, Phys. Rev. B 95, 245305 (2017).
  • [65] R. Chen, C.-Z. Chen, J.-H. Sun, B. Zhou, and D.-H. Xu, Phase Diagrams ofWeyl Semimetals with Competing Intraorbital and Interorbital Disorders, Phys. Rev. B 97, 235109 (2018).
  • [66] C.-B. Hua, R. Chen, D.-H. Xu, and B. Zhou, Disorder-Induced Majorana Zero Modes in a Dimerized Kitaev Superconductor Chain, Phys. Rev. B 100, 205302 (2019).
  • [67] C. Wang and X. R. Wang, Disorder-induced quantum phase transitions in three-dimensional second-order topological insulators, Phys. Rev. Res. 2, 033521 (2020).
  • [68] D. Vu and S. Das Sarma, Weak Quantization of Noninteracting Topological Anderson Insulator, Phys. Rev. B 106, 134201 (2022).
  • [69] R. Chen, X.-X. Yi, and B. Zhou, Four-dimensional topological Anderson insulator with an emergent second Chern number, Phys. Rev. B 108, 085306 (2023).