Quantum anomalous Hall effect in monolayers Ti2X2 (X = P, As, Sb, Bi) with tunable Chern numbers by adjusting magnetization orientation

Keer Huang , Lei Li , Wu Zhao , Xuewen Wang

Front. Phys. ›› 2025, Vol. 20 ›› Issue (2) : 024203

PDF (3693KB)
Front. Phys. ›› 2025, Vol. 20 ›› Issue (2) : 024203 DOI: 10.15302/frontphys.2025.024203
RESEARCH ARTICLE

Quantum anomalous Hall effect in monolayers Ti2X2 (X = P, As, Sb, Bi) with tunable Chern numbers by adjusting magnetization orientation

Author information +
History +
PDF (3693KB)

Abstract

Despite extensive research, the achievement of tunable Chern numbers in quantum anomalous Hall (QAH) systems remains a challenge in the field of condensed matter physics. Here, we theoretically proposed that Ti2X2 (X = P, As, Sb, Bi) can realize tunable Chern numbers QAH effect by adjusting their magnetization orientations. In the case of Ti2P2 and Ti2As2, if the magnetization lies in the xy plane, and all C2 symmetries are broken, a low-Chern-number phase with C = 1 will manifest. Conversely, if the magnetization is aligned to the z-axis, the systems enter a high-Chern number phase with C = 3. As for Ti2Sb2 and Ti2Bi2, by manipulating the in-plane magnetization orientation, these systems can periodically enter topological phases (C = ±1) over a 60° interval. Adjusting the magnetization orientation from +z to −z will result in the systems’ Chern number alternating between ±1. The non-trivial gap in monolayer Ti2X2 (X = P, As, Sb, Bi) can reach values of 23.4, 54.4, 60.8, and 88.2 meV, respectively. All of these values are close to the room-temperature energy scale. Furthermore, our research has revealed that the application of biaxial strain can effectively modify the magnetocrystalline anisotropic energy, which is advantageous in the manipulation of magnetization orientation. This work provides a family of large-gap QAH insulators with tunable Chern numbers, demonstrating promising prospects for future electronic applications.

Graphical abstract

Keywords

tunable Chern number / ferromagnetic / quantum anomalous Hall effect / Ti2X2 (X = P, As, Sb, Bi) / first-principles

Cite this article

Download citation ▾
Keer Huang, Lei Li, Wu Zhao, Xuewen Wang. Quantum anomalous Hall effect in monolayers Ti2X2 (X = P, As, Sb, Bi) with tunable Chern numbers by adjusting magnetization orientation. Front. Phys., 2025, 20(2): 024203 DOI:10.15302/frontphys.2025.024203

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

In recent years, thanks to the development of topological physics theory and two-dimensional (2D) material preparation, two-dimensional topological materials have received keen attention, enriching condensed matter theory and facilitating the discovery of next-generation quantum functional materials [13]. Among them, the quantum anomalous Hall (QAH) effect is a special 2D topological phase of great research significance, which is characterized by the presence of a non-zero Chern number and spin-polarized helical edge electron states within the energy gap of the bulk band structure [4]. The quantum anomalous Hall insulator, also known as a Chern insulator, is distinguished from a conventional insulator due to its unique properties. In the absence of an external magnetic field, internal magnetic exchange interactions can disrupt time-reversal symmetry and open a gap caused by spin−orbit coupling, resulting in quantized anomalous Hall conductivity [5]. Due to the absence of dissipation in its chiral edge states, the QAH effect holds promise for the development of advanced electronic devices with low power consumption. Additionally, it offers an advantageous platform for the exploration of emerging quantum phenomena, such as Majorana fermions and topological magnetoelectric effects [610].

Hence, achieving the QAH insulators has become an exceedingly significant focal point within the realm of condensed matter physics. To date, many methodologies and approaches have been proposed to achieve QAH insulators [1124]. The first observation of the QAH effect occurred in thin films of (Bi/Sb)2Te3 doped with Cr at extremely low temperatures (below 30 mK) due to the weak magnetic coupling in the doped Cr atoms and the narrow band gap of the material [25]. However, the impracticality of these extreme experimental conditions limits the potential application of this effect. Since the QAH insulator is a particular type of two-dimensional ferromagnetic semiconductor characterized by a non-trivial bandgap opened by spin−orbit coupling (SOC), both strong ferromagnetic (FM) coupling and SOC strength are crucial to reaching the high-temperature QAH state. Hence, the objective of increasing the observation temperature is to discover an inherent magnetic topological insulator. Fortunately, the QAH effect has been observed in a thin film of MnBi2Te4 at a temperature of 1.4 K. By applying an external magnetic field to induce a transition from antiferromagnetic to ferromagnetic interlayer magnetic coupling, this temperature can be raised to 6.5 K [26]. This finding strongly suggests that the intrinsic magnetic insulator has the potential to support the QAH effect at high temperatures, making it suitable for practical applications.

Furthermore, apart from the temperature challenge, the effective control and manipulation of the topological state of Chern insulators is another crucial concern, particularly regarding increasing the Chern number. Enhancing the Chern number in high-Chern-number QAH systems can significantly enhance functionality by inducing the additional chiral edge channels, which exhibit dissipationless properties [2730]. So far, several scientific studies have shown that the QAH effect can be achieved theoretically with adjustable Chern numbers in different magnetic topological materials, such as MnBi2Te4 and multilayer structures of magnetically doped topological insulators [31, 32]. In addition, by altering the direction of magnetization, the Chern number of the materials can also be tuned [3336]. Topological states are typically protected by different symmetries in the system, which means that different topological states require different symmetries. However, the low symmetry of 2D systems poses a significant challenge to the emergence of multiple topological states. Therefore, it is imperative to undertake in-depth investigations into novel 2D materials that exhibit the QAH effect and possess tunable Chern numbers.

In recent years, a series of honeycomb-type lattice structures, such as MX [37], MX2 [17, 3847], and M2X3 [4851] families, have been theoretically predicted to host the QAH effect. Inspired by the above research, we propose a family of 2D monolayer materials called Ti2X2 (X = P, As, Sb, Bi) in buckled honeycomb lattice structures. Based on first-principles calculations, we predict they are all QAH insulators. It is crucial to note that all four proposed materials with thermal and dynamical stability have robust FM ground states, beneficial to practical applications and experimental observation. We show that Ti2P2 and Ti2As2 (Ti2Sb2 and Ti2Bi2) are xy easy-plane (z easy-axis) ferromagnetic half-metals. For Ti2P2 and Ti2As2, due to C3 and inversion symmetries, six Dirac points appear in the first Brillouin zone. After considering SOC, the nontrivial band gaps open at the Dirac points, allowing for the hosting of the QAH effect. Remarkably, we find that the Chern number depends on the orientation of magnetization. When the magnetization lies in the xy plane and all C2 symmetries are broken, C=±1. When the magnetization aligns with the z-axis, a high-Chern-number C=±3 phase arises. For Ti2Sb2 and Ti2Bi2, a spin-polarized Dirac point exists at each K and K point in the first Brillouin zone. When SOC is introduced, different orientations of magnetization lead to distinct topological states. By tuning the orientation of in-plane magnetization, the system can periodically enter topological phases (C=±1) over a 60 interval. By tuning magnetization orientations from +z to –z, the Chern number of the system alternatively changes between ±1. Finally, we have also discovered that biaxial strain can effectively adjust the magnetocrystalline anisotropic energy (MAE), enabling the manipulation of magnetization orientation and the achievement of tunable Chern numbers for the QAH insulators.

2 Results and discussion

The Ti2X2 (X = P, As, Sb, Bi) crystal structure exhibits a double-layered hexagonal lattice of buckled honeycomb with the space group of P-3m1 (No. 164), the same as that of MnSe monolayer [52]. As depicted in Fig.1(a), each unit cell contains two Ti atoms and two X atoms, where the lower layer X is situated beneath the upper layer Ti, while the lower layer Ti is located directly beneath the upper layer X. Tab.1 lists the values of the lattice constants (a), Ti1−Ti2 bond length (d), and Ti1−X−Ti3 angle ( α). There are C3 rotation axes through z( C3 z), space inversion (P), and C2 rotation axes through x( C2 x) in the corresponding first Brillouin zone. Especially, due to the existence of C 3z, there are three C2 rotation axes, i.e., C2x, C 2+, and C 2, as shown in Fig.1(b).

To confirm the stability of the Ti2X2, we first calculated the cohesive energy Ec as Ec=(2E Ti+2E XE Ti2X2)/4, in which E Ti2X2 is the energy of each unit cell of Ti2X2, and E Ti, EX are energies of free Ti and X atoms, respectively. The cohesive energies for Ti2P2, Ti2As2, Ti2Sb2, and Ti2Bi2 are 3.51, 3.16, 2.64, and 2.35 eV, respectively. Besides, the dynamic stability is confirmed by the phonon spectra. All phonon modes of Ti2X2 display positive frequencies across the momentum space [Figs. S1(a)−(d)], indicating the stability of each structure. Moreover, the thermodynamic stability could be checked by ab initio molecular dynamics simulations. After annealing for three ps at 300 K, the structures remain intact with the fluctuating temperature and total energy [Figs. S1(e)−(h)], further confirming the thermodynamic stability of Ti2X2.

Then, we investigate the magnetic ground state of the monolayer Ti2X2 (X = P, As, Sb, Bi). The calculations were performed in a 2×1 supercell, with four magnetic phases considered, including one FM and three AFM configurations, i.e., AFM-Néel, AFM-Zigzag, and AFM-Stripy [Fig. S2]. The total energies listed in Table S1 indicate that the ferromagnetic state is preferred in all Ti2X2. In the monolayer Ti2X2, each Ti atom provides three electrons to the corresponding X atom, resulting in partially-occupied valence-electron shells of 3d14s0 for Ti atoms and fully-occupied stable shells for X atoms. According to the crystal field theory, the 3d orbitals of Ti atoms are split into three groups: (d xz, d yz), d z2, and (dxy, d x2 y2). Additionally, after the exchange interaction is introduced, the spin-up and spin-down channels are split, and only one 3d electron occupies one of the (d xz, d yz) orbitals, resulting in a net magnetic moment of 1 μB per Ti atom [Fig.1(c)].

To uncover the underlying mechanism of magnetic coupling, we calculated the magnetic exchange parameters using the total-energy mapping method (see details in Section 2 in the supplementary material). The results are presented in Tab.1. J1 and J2 represent the parameters of the exchange interactions between the nearest-neighbor (NN) and the next-nearest-neighbor (NNN) Ti atoms, respectively. They are determined by comparing the total energies of the materials with the FM, AFM-s, and AFM-z magnetic structures. Clearly, J1 is positive and significantly larger than J 2, giving rise to ultrastable FM ground states in all the Ti2X2 (X = P, As, Sb, Bi) monolayers. Because of the short distance between NN Ti atoms (~3.0 Å), the Ti−Ti hybridization is strong even for 3d orbitals, which is also demonstrated through the dispersive 3d states near the Fermi level [Fig.2(e)−(h)]. Combined with the specific orbital occupation of Ti 3d orbitals, it leads to a ferromagnetic (FM) state in Ti2X2. For NNN Ti atoms, the magnetic exchange parameter J2 is determined by the superexchange (SE) mechanism. According to the Goodenough−Kanamori−Anderson (GKA) SE rule [5355], the SE interaction between the NNN Ti atoms via the X p orbitals is FM for a bond angle α of 90 (Ti1−X−Ti3) and AFM for a bond angle of 180. As presented in Tab.1, the bond angles of Ti−X−Ti in Ti2X2 (X = P, As, Sb, Bi) are 114.0 , 111.4, 105.4, and 102.7, respectively, gradually approaching 90 . So, the NNN exchange parameter (J2) changes from antiferromagnetic coupling for Ti2P2 to ferromagnetic coupling for Ti2X2 (X = As, Sb, Bi). As listed in Tab.1, D denotes the MAE per Ti ion, defined as ΔE= EE, where E /(E) is the total energy when the magnetic moment is oriented along the out-of-plane (in-plane) direction. The MAE illustrates that the magnetization lies in the xy plane for Ti2P2 and Ti2As2, and is aligned with the z-axis for Ti2Sb2 and Ti2Bi2.

We now investigate the electronic band structures of the Ti2X2 (X = P, As, Sb, Bi) monolayers. In Fig.2(a)−(d), one can observe that all monolayers are half-metals with wide spin windows (1.55, 1.41, 0.98, and 0.76 eV for Ti2P2, Ti2As2, Ti2Sb2, and Ti2Bi2, respectively). The spin-down bands exhibit a large energy gap around the Fermi level, while the spin-up bands display a metallic electronic structure, promising them candidates in spintronics, such as magnetic tunnel junctions and spin valves [56, 57]. For Ti2P2 and Ti2As2, a spin-polarized Dirac point appears along the ΓK high symmetry line at the Fermi level. The presence of C 3z rotation and inversion symmetries results in six spin-polarized Dirac points in the first Brillouin zone. But in terms of Ti2Sb2 and Ti2Bi2, at the Fermi level, there exists a spin-polarized Dirac point located at each K and K points, respectively. Consequently, Ti2P2 (Ti2As2) and Ti2Sb2 (Ti2Bi2) exhibit distinct electronic structural properties. Subsequently, we will categorize them into two groups and analyze each individually.

Firstly, we will investigate the topological properties of Ti2P2 and Ti2As2. Owing to their similar electronic structure, we take monolayer Ti2As2 as an example in the following. After the spin-orbit coupling (SOC) is turned on, various topological phases can be formed depending on the orientation of magnetization. As illustrated in Fig.1(b), there exist three C2 rotation axes along the high symmetry lines of ΓK(K). When the magnetization aligns with the C2 rotation axis (e.g., x-direction), the spin-polarized Dirac points are still preserved along the Γ K line [see Fig. S3(a)]. In order to open a gap in the in-plane magnetization, it is crucial to break all C2 rotation symmetries. When the magnetization direction is oriented at a 30 angle with the x-axis, φ =30 [Fig. S4], as shown in Fig.3(a), a global gap of about 54.4 meV opens in the monolayer with a Chern number of C = 1. As plotted in Fig.3(b), a single chiral edge state connecting the valence and conductance bands demonstrates the nontrivial band topology with an exotic QAH effect. Additionally, from Tab.1, the tiny MAE (~0.7 meV) suggests that by applying a magnetic field of a few Tesla, it is possible to experimentally orient the magnetization from the xy plane to the z direction [33, 45]. As illustrated in Fig.3(e), the magnetization aligned with the z direction leads to the monolayer Ti2As2 entering a high-Chern-number phase (C = 3) with a band gap of 8.4 meV. As shown in Fig.3(f), there are three chiral gapless edge states connecting the valence and conduction bands in direct agreement with the high Chern number. Further, we continuously adjust the orientations of magnetization to investigate the topological phase transition. Fig.3(c) illustrates that when the magnetization lies in the plane and aligns with the C2 rotation axis, i.e., φ =0 (180 ), 60 (210 ), 120 (300 ), the band gaps close at the ΓK(K, K) [Fig. S5] high symmetry line [Fig. S6]. Consequently, the Chern number of the system alternates between ±1 with a periodicity of 60 . In the case of the out-of-plane magnetization, starting from z direction (θ=90), along the path of φ= 30, the global band gap gradually decreases with the decrease of polar angle θ. Through the first-principles calculations, it has been determined that the gap will completely close at θ=85.6, 98.6 ( 265.6, 278.6) [Fig. S7], and subsequently reopen. The critical polar angle determines the phase boundaries between low and high Chern numbers. Fig.3(g) represents the phase diagram with distinct Chern numbers. Additionally, it is observed that the maximum gap value does not occur at θ= 0 but rather slightly deviates, manifesting at θ=10.

For other paths (φ=0 and φ=5 ), the phase diagrams are presented in Fig. S8. In the case of Ti2P2, phase diagrams of the Chern number as a function of azimuthal angle φ and polar angle θ are shown in Fig. S9.

In order to explain the results mentioned above, a two-band effective Hamiltonian can be employed using the kp model. In the absence of the SOC, the spin and orbital degrees of freedom become independent, resulting in the spin being a well-defined quantum number within its own subspace. As such, the Γ−K path has the little group C2, whose symmetry operators are E and C 2x. The effective Hamiltonian H e ff in an effective kp model can be constrained by C 2xH eff(kx, ky)C2x1=H eff(kx, k y). The H e ff can be written as H eff= h0(q)+h( q)σ, where q is referenced to the Dirac Point, and the Pauli matrices σ act on the orbital space (see detail in Section 3.1 in the supplementary material), there, h0(q)=Cqx, h(q)=(Aqx,Bq y,0).

When considering the SOC, out-of-plane ferromagnetism breaks the C2 symmetry as the magnetization direction is perpendicular to the C2 rotation axes. The SOC introduces a Dirac mass term H =Mσz into the effective Hamiltonian, leading to the formation of a Dirac gap. This phenomenon remains valid even when the magnetization direction lies in the plane but does not align with the C2 rotational axes. When the magnetization direction is oriented along the C2 rotational axes, the rotational symmetry of C2 remains preserved, preventing the introduction of a Dirac mass term by the SOC. To understand the Chern number variation, we computed the distribution of Berry curvature in the Brillouin zone. Fig.3(d) and (h) present the distribution of Berry curvature for in-plane and out-of-plane magnetizations, respectively. In the case of in-plane magnetization, two-thirds of the Brillouin zones exhibit Berry curvatures with consistent signs, while the remaining one-third of the zone displays Berry curvatures with opposite signs. Conversely, for out-of-plane magnetization, the Berry curvatures give the same contribution in all Brillouin zones. Since each gapped Dirac point in a system contributes to a Chern number of ±1/2, consequently, the in-plane magnetization results in a small net value of low Chern number (C = 1), while the out-of-plane magnetization leads to a high Chern number (C = 3). Therefore, multiple Dirac cones within the first Brillouin zone are necessary to achieve tunable Chern numbers in the Chern insulator by adjusting magnetization orientation.

Next, we will analyze the topological characteristics of Ti2Sb2 and Ti2Bi2. Given their similar properties, Ti2Sb2 will be utilized as an illustrative example. As previously mentioned, in the absence of the SOC, there exists a spin-polarized Dirac point at each K and K point. However, when SOC is introduced, different orientations of magnetization lead to distinct topological states. When the magnetization aligns with the C2 rotation axis, the spin-polarized Dirac points still exist but slightly deviate from the K(K) point on the K(K)−Γ path [Fig. S3(b)]. On the other hand, when the magnetization direction is at an angle of φ =30 , as shown in Fig.4(a), a global gap of approximately 10.2 meV emerges in the system, and as plotted in Fig.4(b), one gapless chiral edge state within the SOC gap is clearly visible, accompanied by a Chern number of C = 1. By tuning the orientation of in-plane magnetization, similar to Ti2As2, the system can periodically enter topological phases with alternative Chern numbers of C=±1 in the interval of 60 [Fig.4(c)]. When the magnetization direction is oriented towards the z-axis, as shown in Fig.4(d), a global gap of 60.8 meV emerges in the system, and as plotted in Fig.4(e), one chiral gapless edge states connect the valence and conduction bands, concomitant with a Chern number of C = 1. We continuously tune magnetization orientation from +z to −z (along φ=30). As shown in Fig.4(f), the Chern number of the system also alternatively changes between ±1. However, it is worth noting that the phase transition point does not precisely occur at θ= 0, but rather at θ= 11.6. In the case of Ti2Bi2, phase diagrams of the Chern number as a function of azimuthal angle φ and polar angle θ are shown in Fig. S10. The difference in Chen number in Ti2P2 (Ti2As2) and Ti2Sb2 (Ti2Bi2) can be analyzed by their electronic structure properties. As shown in Fig.2(e)−(h), as the atomic number of elements in the group VA increases, the relative positions of two different irreducible energy levels near the Fermi level at K point change, until Ti2Sb2, where the two energy levels undergo energy level inversion at K point, which results in Ti2P2 (Ti2As2) and Ti2Sb2 (Ti2Bi2) having different Chern number when magnetization direction along +/−z.

The above results can also be described using a two-band effective Hamiltonian through the kp method. The d xz and dyz orbitals of Ti ions form the basis functions of the 2D irreducible representation at the K( K) valley located around the Fermi level. Owing to the constraint of the corresponding little-group, the effective Hamiltonian Heff can be written as H ef f=A1kxσx+A1kyσy (see detail in Section 3.2 in the supplementary material). Consequently, a Dirac cone with linear dispersions emerges at the K( K) valley, resembling the characteristic Dirac cones observed in graphene. When including the SOC and magnetization direction aligns with the C2 rotational axes (e.g., x-direction), the C2 rotational symmetry is preserved and introduces a A0σx term into the effective Hamiltonian, results in the Dirac point deviate slightly from K point on the KΓ path. However, in the other cases, the SOC will introduce a Dirac mass term and open a Dirac gap.

Because of the inherent flexibility of two-dimensional materials, the manipulation of strain has proven to be an effective method for modifying various material properties [5862]. In this study, we investigate the impact of strain on monolayer Ti2X2 (X = P, As, Sb, Bi) by subjecting it to biaxial strain ranging from −6% to 6%. As the results depicted in Fig.5(a)−(d), the MAE decreases monotonically from tensile strain to compressive strain, suggesting that strain plays a beneficial role in facilitating the transition of the QAH effect between different topological phases.

3 Conclusion

In conclusion, we systematically investigate these four materials’ structural stability, magnetic properties, and topologically electronic properties based on density functional theory, kp model, and symmetry analysis. The phonon calculations and molecular dynamics simulations confirmed the dynamical and thermal stabilities of the system. Their FM ground states are determined by both the NN Ti−Ti direct exchange coupling and NNN Ti−Ti superexchange coupling. We find that Ti2X2 (X = P, As, Sb, Bi) can realize the QAH effect with tunable Chern numbers by adjusting the orientation of magnetization. For Ti2P2 and Ti2As2, the QAH effect with low Chern numbers is formed for the in-plane magnetization, while the QAH effect with high Chern numbers arises when the magnetization is aligned along the z direction. Similarly, for Ti2Sb2 and Ti2Bi2, by tuning the orientation of in-plane magnetization, the system can periodically enter topological phases with Chern numbers of C=± 1 within a 60 interval and tuning magnetization orientations from +z to −z, the Chern number of the system alternatively changes between ±1. By subjecting the material to a stretching force, the magnitude of the MAE can be reduced, which is beneficial for controlling the orientation of magnetization using an external magnetic field. Our work provides an ideal platform to realize tunable Chern numbers for the QAH effect via adjusting magnetization orientation, which is expected to be applied in fields such as spintronics, topological superconductivity, and quantum computing.

References

[1]

L. Fu, C. L. Kane, and E. J. Mele, Topological insulators in three dimensions, Phys. Rev. Lett. 98(10), 106803 (2007)

[2]

P. Li, X. Yang, Q. S. Jiang, Y. Z. Wu, and W. Xun, Built-in electric field and strain tunable valley-related multiple topological phase transitions in VSiXN4 (X = C, Si, Ge, Sn, Pb) monolayers, Phys. Rev. Mater. 7(6), 064002 (2023)

[3]

H. Huan, Y. Xue, B. Zhao, G. Y. Gao, H. R. Bao, and Z. Q. Yang, Strain-induced half-valley metals and topological phase transitions in MBr2 monolayers (M = Ru, Os), Phys. Rev. B 104(16), 165427 (2021)

[4]

P. Li, Prediction of intrinsic two dimensional ferromagnetism realized quantum anomalous Hall effect, Phys. Chem. Chem. Phys. 21(12), 6712 (2019)

[5]

B. I. Halperin, Quantized Hall conductance, current-carrying edge states, and the existence of extended states in a two-dimensional disordered potential, Phys. Rev. B 25(4), 2185 (1982)

[6]

F. D. M. Haldane, Nobel Lecture: Topological quantum matter, Rev. Mod. Phys. 89(4), 040502 (2017)

[7]

M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82(4), 3045 (2010)

[8]

X. L. Qi and S. C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83(4), 1057 (2011)

[9]

H. Chi and J. S. Moodera, Progress and prospects in the quantum anomalous Hall effect, APL Mater. 10(9), 090903 (2022)

[10]

C. Z. Chang, C. X. Liu, and A. H. MacDonald, Colloquium: Quantum anomalous Hall effect, Rev. Mod. Phys. 95(1), 011002 (2023)

[11]

J. Cai, E. Anderson, C. Wang, X. Zhang, X. Liu, W. Holtzmann, Y. Zhang, F. Fan, T. Taniguchi, K. Watanabe, Y. Ran, T. Cao, L. Fu, D. Xiao, W. Yao, and X. Xu, Signatures of fractional quantum anomalous Hall states in twisted MoTe2, Nature 622(7981), 63 (2023)

[12]

P. Deng, P. Zhang, C. Eckberg, S. K. Chong, G. Yin, E. Emmanouilidou, X. Che, N. Ni, and K. L. Wang, Quantized resistance revealed at the criticality of the quantum anomalous Hall phase transitions, Nat. Commun. 14(1), 5558 (2023)

[13]

G. M. Ferguson, R. Xiao, A. R. Richardella, D. Low, N. Samarth, and K. C. Nowack, Direct visualization of electronic transport in a quantum anomalous Hall insulator, Nat. Mater. 22, 1100 (2023)

[14]

P. J. Guo, Z. X. Liu, and Z. Y. Lu, Quantum anomalous Hall effect in collinear antiferromagnetism, npj Comput. Mater. 9(1), 70 (2023)

[15]

Y. L. Jiao, X. T. Zeng, C. Chen, Z. Gao, K. Y. Song, X. L. Sheng, and S. A. Yang, Monolayer and bilayer PtCl3: Energetics, magnetism, and band topology, Phys. Rev. B 107(7), 075436 (2023)

[16]

J. Li and R. Q. Wu, Electrically tunable topological phase transition in van der Waals heterostructures, Nano Lett. 23(6), 2173 (2023)

[17]

S. S. Li, X. Y. Li, W. X. Ji, P. Li, S. S. Yan, and C. W. Zhang, Quantum anomalous Hall effect with a high and tunable Chern number in monolayer NdN2, Phys. Chem. Chem. Phys. 25(27), 18275 (2023)

[18]

X. Y. Li, X. F. Xu, H. Zhou, H. X. Jia, E. Wang, H. X. Fu, J. T. Sun, and S. Meng, Tunable topological states in stacked Chern insulator bilayers, Nano Lett. 23(7), 2839 (2023)

[19]

H. Park, J. Cai, E. Anderson, Y. Zhang, J. Zhu, X. Liu, C. Wang, W. Holtz-mann, C. Hu, Z. Liu, T. Taniguchi, K. Watanabe, J. H. Chu, T. Cao, L. Fu, W. Yao, C. Z. Chang, D. Cobden, D. Xiao, and X. Xu, Observation of fractionally quantized anomalous Hall effect, Nature 622(7981), 74 (2023)

[20]

P. Zhang, P. P. Balakrishnan, C. Eckberg, P. Deng, T. Nozaki, S. K. Chong, P. Quarterman, M. E. Holtz, B. B. Maranville, G. Qiu, L. Pan, E. Emmanouilidou, N. Ni, M. Sahashi, A. Grutter, and K. L. Wang, Exchange biased quantum anomalous Hall effect, Adv. Mater. 35(31), 2300391 (2023)

[21]

S. Zhang, X. M. Zhang, Z. Q. He, L. Jin, C. Liu, Y. Liu, and G. D. Liu, Weyl nodal lines, Weyl points and the tunable quantum anomalous Hall effect in two-dimensional multiferroic metal oxynitride: Tl2NO2, Nanoscale 15(34), 14018 (2023)

[22]

Y. F. Zhao, R. Zhang, J. Cai, D. Zhuo, L. J. Zhou, Z. J. Yan, M. H. W. Chan, X. Xu, and C. Z. Chang, Creation of chiral interface channels for quantized transport in magnetic topological insulator multilayer heterostructures, Nat. Commun. 14(1), 770 (2023)

[23]

L. J. Zhou, R. B. Mei, Y. F. Zhao, R. X. Zhang, D. Y. Zhuo, Z. J. Yan, W. Yuan, M. Kayyalha, M. H. W. Chan, C. X. Liu, and C. Z. Chang, Confinement-induced chiral edge channel interaction in quantum anomalous Hall insulators, Phys. Rev. Lett. 130(8), 086201 (2023)

[24]

Y. Gao, Y. Y. Zhang, J. T. Sun, L. Zhang, S. Zhang, and S. Du, Quantum anomalous Hall effect in two-dimensional Cu-dicyanobenzene coloring-triangle lattice, Nano Res. 13(6), 1571 (2020)

[25]

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(6129), 167 (2013)

[26]

Y. Deng, Y. Yu, M. Z. Shi, Z. Guo, Z. Xu, J. Wang, X. H. Chen, and Y. Zhang, Quantum anomalous Hall effect in intrinsic magnetic topological insulator MnBi2Te4, Science 367(6480), 895 (2020)

[27]

J. Wang, B. Lian, H. Zhang, Y. Xu, and S. C. Zhang, Quantum anomalous Hall effect with higher plateaus, Phys. Rev. Lett. 111(13), 136801 (2013)

[28]

C. Z. Chang, W. Zhao, D. Y. Kim, P. Wei, J. K. Jain, C. Liu, M. H. W. Chan, and J. S. Moodera, Zero-field dissipationless chiral edge transport and the nature of dissipation in the quantum anomalous Hall state, Phys. Rev. Lett. 115(5), 057206 (2015)

[29]

Y. X. Wang and F. Li, High Chern number phase in topological insulator multilayer structures, Phys. Rev. B 104(3), 035202 (2021)

[30]

M. Bosnar, A. Y. Vyazovskaya, E. K. Petrov, E. V. Chulkov, and M. M. Otrokov, High Chern number van der Waals magnetic topological multilayers MnBi2Te4/hBN, npj 2D Mater. Appl. 7(1), 33 (2023)

[31]

J. Ge, Y. Liu, J. Li, H. Li, T. Luo, Y. Wu, Y. Xu, and J. Wang, High-chernnumber and high-temperature quantum Hall effect without landau levels, Natl. Sci. Rev. 7(8), 1280 (2020)

[32]

Y. F. Zhao, R. Zhang, R. Mei, L. J. Zhou, H. Yi, Y. Q. Zhang, J. Yu, R. Xiao, K. Wang, N. Samarth, M. H. W. Chan, C. X. Liu, and C. Z. Chang, Tuning the Chern number in quantum anomalous Hall insulators, Nature 588(7838), 419 (2020)

[33]

Y. Li, J. Li, Y. Li, M. Ye, F. Zheng, Z. Zhang, J. Fu, W. Duan, and Y. Xu, High-temperature quantum anomalous Hall insulators in lithium-decorated iron-based superconductor materials, Phys. Rev. Lett. 125(8), 086401 (2020)

[34]

H. Huan, Y. Xue, B. Zhao, H. Bao, L. Liu, and Z. Yang, Tunable Weyl half-semimetals in two-dimensional iron-based materials MFeSe (M = Tl, In, Ga), Phys. Rev. B 106(12), 125404 (2022)

[35]

Z. Li, Y. Han, and Z. Qiao, Chern number tunable quantum anomalous Hall effect in monolayer transitional metal oxides via manipulating magnetization orientation, Phys. Rev. Lett. 129(3), 036801 (2022)

[36]

L. Jin, L. R. Wang, X. M. Zhang, Y. Liu, X. F. Dai, H. L. Gao, and G. D. Liu, Fully spin-polarized Weyl fermions and in/out-of-plane quantum anomalous Hall effects in a two-dimensional d0 ferromagnet, Nanoscale 13(11), 5901 (2021)

[37]

K. Wang, Y. Li, H. Mei, P. Li, and Z. X. Guo, Quantum anomalous Hall and valley quantum anomalous Hall effects in two-dimensional d0 orbital XY monolayers, Phys. Rev. Mater. 6(4), 044202 (2022)

[38]

L. Li, X. Kong, X. Chen, J. Li, B. Sanyal, and F. M. Peeters, Monolayer 1T-LaN2: Dirac spin-gapless semiconductor of p-state and Chern insulator with a high Chern number, Appl. Phys. Lett. 117(14), 143101 (2020)

[39]

S. D. Guo, W. Q. Mu, and B. G. Liu, Valley-polarized quantum anomalous Hall insulator in monolayer RuBr2, 2D Mater. 9(3), 035011 (2022)

[40]

X. Feng, L. Cai, Z. Chen, Y. Dai, B. Huang, and C. Niu, Tunable second-order topological insulators in Chern insulators 2H-FeX2 (X = Cland Br), Appl. Phys. Lett. 122(19), 193104 (2023)

[41]

Y. Wu, J. Tong, L. Deng, F. Luo, F. Tian, G. Qin, and X. Zhang, Realizing spontaneous valley polarization and topological phase transitions in monolayer ScX2 (X = Cl, Br, I), Acta Mater. 246, 118731 (2023)

[42]

X. L. Sheng and B. K. Nikolić, Monolayer of the 5d transition metal trichloride OsCl3: A playground for two-dimensional magnetism, room-temperature quantum anomalous Hall effect, and topological phase transitions, Phys. Rev. B 95(20), 201402 (2017)

[43]

J. He, X. Li, P. Lyu, and P. Nachtigall, Near-room-temperature Chern insulator and Dirac spin-gapless semiconductor: Nickel chloride monolayer, Nanoscale 9(6), 2246 (2017)

[44]

S. D. Guo, W. Q. Mu, X. B. Xiao, and B. G. Liu, Intrinsic room-temperature piezoelectric quantum anomalous Hall insulator in Janus monolayer Fe2IX (X = Cl and Br), Nanoscale 13(30), 12956 (2021)

[45]

J. Y. You, C. Chen, Z. Zhang, X. L. Sheng, S. A. Yang, and G. Su, Two-dimensional Weyl half-semimetal and tunable quantum anomalous Hall effect, Phys. Rev. B 100(6), 064408 (2019)

[46]

H. Zhang, W. Yang, P. Cui, X. Xu, and Z. Zhang, Prediction of monolayered ferromagnetic CrMnI6 as an intrinsic high-temperature quantum anomalous Hall system, Phys. Rev. B 102(11), 115413 (2020)

[47]

J. Y. You, Z. Zhang, B. Gu, and G. Su, Two-dimensional room-temperature ferromagnetic semiconductors with quantum anomalous Hall effect, Phys. Rev. Appl. 12(2), 024063 (2019)

[48]

B. Zhang, F. Deng, X. Chen, X. Lv, and J. Wang, Quantum anomalous Hall effect in M2X3 honeycomb Kagome lattice, J. Phys. Condens. Matter 34(47), 475702 (2022)

[49]

H. P. Wang, W. Luo, and H. J. Xiang, Prediction of high-temperature quantum anomalous Hall effect in two-dimensional transition-metal oxides, Phys. Rev. B 95(12), 125430 (2017)

[50]

P. Li and Z. X. Guo, The Dirac half-semimetal and quantum anomalous Hall effect in two-dimensional Janus Mn2X3Y3 (X, Y = F, Cl, Br, I), Phys. Chem. Chem. Phys. 23(35), 19673 (2021)

[51]

P. Li, Y. Ma, Y. Zhang, and Z. X. Guo, Room temperature quantum anomalous Hall insulator in a honeycomb-Kagome lattice, Ta2O3, with huge magnetic anisotropy energy, ACS Appl. Electron. Mater. 3(4), 1826 (2021)

[52]

M. Aapro, M. N. Huda, J. Karthikeyan, S. Kezilebieke, S. C. Ganguli, H. G. Herrero, X. Huang, P. Liljeroth, and H. P. Komsa, Synthesis and properties of monolayer MnSe with unusual atomic structure and antiferromagnetic ordering, ACS Nano 15(8), 13794 (2021)

[53]

J. B. Goodenough, Theory of the role of covalence in the perovskite-type manganites [La,M(II)]MnO3, Phys. Rev. 100(2), 564 (1955)

[54]

J. Kanamori, Superexchange interaction and symmetry properties of electron orbitals, J. Phys. Chem. Solids 10(2−3), 87 (1959)

[55]

J. Kanamori, Crystal distortion in magnetic compounds, J. Appl. Phys. 31(5), S14 (1960)

[56]

W. E. Pickett and J. S. Moodera, Half metallic magnets, Phys. Today 54(5), 39 (2001)

[57]

I. Žutić, J. Fabian, and S. Das Sarma, Spintronics: Fundamentals and applications, Rev. Mod. Phys. 76(2), 323 (2004)

[58]

Z. Liu, M. Amani, S. Najmaei, Q. Xu, X. Zou, W. Zhou, T. Yu, C. Qiu, A. G. Birdwell, F. J. Crowne, R. Vajtai, B. I. Yakobson, Z. Xia, M. Dubey, P. M. Ajayan, and J. Lou, Strain and structure heterogeneity in MoS2 atomic layers grown by chemical vapour deposition, Nat. Commun. 5(1), 5246 (2014)

[59]

G. H. Ahn, M. Amani, H. Rasool, D. H. Lien, J. P. Mastandrea, J. W. III Ager, M. Dubey, D. C. Chrzan, A. M. Minor, and A. Javey, Strain-engineered growth of two-dimensional materials, Nat. Commun. 8(1), 608 (2017)

[60]

C. Zhang, M. Y. Li, J. Tersoff, Y. Han, Y. Su, L. J. Li, D. A. Muller, and C. K. Shih, Strain distributions and their influence on electronic structures of WSe2‒MoS2 laterally strained heterojunctions, Nat. Nanotechnol. 13(2), 152 (2018)

[61]

Z. Dai, L. Liu, and Z. Zhang, 2D materials: Strain engineering of 2D materials: Issues and opportunities at the interface, Adv. Mater 31(45), 1970322 (2019)

[62]

Z. Li, Y. Lv, L. Ren, J. Li, L. Kong, Y. Zeng, Q. Tao, R. Wu, H. Ma, B. Zhao, D. Wang, W. Dang, K. Chen, L. Liao, X. Duan, X. Duan, and Y. Liu, Efficient strain modulation of 2D materials via polymer encapsulation, Nat. Commun. 11(2), 1151 (2020)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3693KB)

Supplementary files

fop-25019-of-lilei_suppl_1

2530

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/