Topology and ferroelectricity in group-V monolayers
Rehman Mutee Ur1, Hua Chenqiang1, 2, Lu Yunhao1, 2, ‡
State Key Laboratory of Silicon Materials, School of Materials Science and Engineering, Zhejiang University, Hangzhou 310027, China
Zhejiang Province Key Laboratory of Quantum Technology and Device, Department of Physics, Zhejiang University, Hangzhou 310027, China

 

† Corresponding author. E-mail: luyh@zju.edu.cn

Project supported by the National Natural Science Foundation of China (Grant Nos. 11974307 and 61574123), Zhejiang Provincial Natural Science Foundation, China (Grant No. D19A040001), the Fundamental Research Funds for the Central Universities of China, and the 2DMOST, Shenzhen University (Grant No. 2018028).

Abstract

The group-V monolayers (MLs) have been studied intensively after the experimental fabrication of two-dimensional (2D) graphene and black phosphorus. The observation of novel quantum phenomena, such as quantum spin Hall effect and ferroelectricity in group-V elemental layers, has attracted tremendous attention because of the novel physics and promising applications for nanoelectronics in the 2D limit. In this review, we comprehensively review recent research progress in engineering of topology and ferroelectricity, and several effective methods to control the quantum phase transition are discussed. We then introduce the coupling between topological orders and ferroelectric orders. The research directions and outlooks are discussed at the end of the perspective. It is expected that the comprehensive overview of topology and ferroelectricity in 2D group-V materials can provide guidelines for researchers in the area and inspire further explorations of interplay between multiple quantum phenomena in low-dimensional systems.

1. Introduction

Two-dimensional (2D) materials have revolutionized the new generation of devices. Decreasing the thickness of materials down to single (or several) atomic layer has attracted tremendous interest since they could display exotic phenomena which are different from their bulk counterparts, e.g., quantum spin Hall effect,[14] 2D ferroelectricity,[5] superconductivity,[68] or magnetism.[911] Motivated by the remarkable progress of graphene,[12,13] 2D materials other than graphene have been a pronounced research field of interest. Recent experimental successes in the elemental 2D layers have invigorated the search for more fascinating structures and properties based on elemental 2D systems, for example, Si (silicene),[1416] Ge (germanene),[17,18] Sn (stanene),[19] B (borophene),[20] P (phosphorene),[21] As (arsenene),[22] Sb (antimonene),[23] Bi (bismuthene),[4] Te (tellurene),[24] and Se (selenene).[25]

Graphene, the first successfully isolated elemental material in 2D limit with the honeycomb lattice, holds the linear band dispersion and crossing at the K point in the 1st Brillouin zone (BZ), generating the Dirac fermions. Actually, a tiny energy gap opens up at the Dirac points due to spin–orbital coupling (SOC) and graphene is a quantum spin Hall (QSH) insulator.[26] QSH insulators, also named as 2D topological insulators (TI), hold counter-propagating and oppositely spin-polarized conducting states at their edges, which are protected by the time reversal symmetry against the weak localized perturbations. Therefore, these robust edge states endow 2D TIs with possibilities for a wide range of applications in dissipationless electronics, topological quantum computing, and so on.[2730] However, the SOC effect in graphene is too small to achieve the experimental measurement and topological electronic devices,[31] which stimulates scientists to search for large-gap 2D TIs not only in compounds but also in elemental materials. For example, graphene-like honeycomb structures have been widely explored including group-IV and group-V materials.[32,33] For group-V materials, except for the graphene-like structures, the puckered structures in black phosphorus lattice are also able to exhibit topological properties.[34] This puckered structure is unique for group-V elements due to the one extra valence electron compared to group-IV elements and provides a feasible way to tune the crystal symmetry and distortion of the 2D system, allowing intimate coupling and control between various order parameters.

Ferroelectricity is also an exciting ferroic order, possessing reversible spontaneous electronic polarization which is important for applications in non-volatile memory devices, field effect transistors, solar cells, and sensors.[3538] Normally, ferroelectricity would disappear when the film thickness is below a critical value, due to the depolarizing electric field and electron screening. A new class of materials, called as hyperferroelectrics, have been identified to remain in stable polar states in ultrathin atomic layers against unscreened depolarizing field.[39] However, the experimental evidence of these materials is still rare. Recently, the ferroelectric instability becomes of interest in 2D van der Waals materials[40,41] and 2D ferroelectricity is identified in 2D group-V binary compounds[42] and elemental monolayers (MLs),[43] which is impossible for their 3D bulk counterparts. Conventionally, the intrinsic ferroelectricity has been reported only for compounds in which the required positive and negative charge centers are carried by different elements. The elemental materials are always stable in centrosymmetric structures without intrinsic ferroelectricity. While, the lack of constrain in one dimension could break the inversion symmetry and drive the 2D elemental layers into a ferroelectric state.[4345]

The coupling between different ordering parameters is important not only for fundamental research, but also meaningful for cross-control of different orderings in applications. The most striking of such controls is switching magnetic ordering and/or ferroelectric polarization ordering (referred as multiferroicity or magnetoelectric coupling) by electric and magnetic fields, which is important for the design of new electronic devices.[46,47] There are a lot of work and reviews discussing coupling between magnetic and electronic orderings. As far as we know, there is still no review touching interplay between electronic topological ordering and ferroelectric polarization/charge ordering in 2D materials. In this paper, we will systematically report the recent advances about the topology and ferroelectricity in group-V 2D materials. Starting from structure classification of group-V MLs, we will summarize the recent work on 2D group-V MLs about the topology and ferroelectricity with their possible approaches of modulation (e.g., electric field, chemical functionalization, and fabricating the binary systems). Lastly, we would like to discuss the possible coupling between these two interesting quantum properties. Here, the topology refers to the electronic structure, not the topological domain pattern in ferroelectric materials that is the topology of the geometrical structure. We certainly do not make any attempt to cover everything in this mini-review, and the topics chosen are more or less biased by the author’s own research interests. The review mainly focuses on group-V MLs and only the advances in the last several years will be discussed.

2. Structural allotropes
2.1. Bulk allotropes

Before the discussion of 2D layers, we firstly discuss the structure character of bulk group-V elements. There are five valence electrons in group-V elements and diverse valence states (–3, +3, and +5) are possible for them, resulting in a variety of structural allotropes for bulk of group-V elements at ambient conditions. Phosphorus, the most studied element of group-V, has several allotropes, which are named as white, red, blue, violet, and black phosphorus.[4851] In addition, amorphous phosphorus is also very popular.[51,52] The discovery of black phosphorus, the most stable allotrope of phosphorus at ambient conditions, is dated back in 1914[50,53] and it crystallizes in parallel puckered layer orthorhombic structure with space group Cmca as shown in Fig. 1(a). The 3s and 3p orbitals of phosphorus undergo sp3 hybridization iw three σ-covalent bonds along with an out-of-plane oriented lone pair of electrons. Two of these three covalent bonds lie parallel to the plane while the third one is almost perpendicular to the plane, forming a puckered layer. And, the out-of-plane lone pair electrons combine the puckered layers together, crystalizing in the orthorhombic structure (α phase). Phosphorus can also crystallize into a rhombohedral geometry, i.e., blue phosphorus (β phase) with space group .[54] Blue phosphorus also adopts a layered structure with buckled layers consisting of knitted honeycomb rings as shown in Fig. 1(b), where the three σ-covalent bonds within the layer are equivalent, resulting in the hexagonal structure.

Fig. 1 Two typical stable phases of group-V elements: (a) orthorhombic α phase with space group Cmca, and (b) rhombohedral β phase with space group . Left and right panels show side and top views, respectively.

Although the most stable crystal structure of phosphorus is the α phase with orthorhombic space group, the other elements of group-V (As, Sb, and Bi) favor the β phase due to the weak s–p hybridization.[55,56] The delocalized p–p coupling causes more and more metallic bonding between atoms from As, Sb to Bi and the hexagonal honeycomb structure becomes more stable. For example, the most common and stable allotrope of As and Sb has the rhombohedral layered structure with space group as shown in Fig. 1(b), consisting of two interpenetrating, face-centered-cubic lattices which are displaced in the [111] direction and slightly distorted in the [111] direction. While the α phase is a metastable phase, the phase transformation between β and α is observed for both As and Sb.[57] For bulk Bi, the last element of group-V, only the rhombohedral structure (β phase) has been observed in experiment.

2.2. Two-dimensional allotropes

Generally, there are more crystal allotropes for 2D materials than the corresponding 3D ones due to the lack of periodic boundary constrain along one dimension. For 2D group-V layers, a lot of allotropes have been predicted by theoretical calculations,[5861] as shown in Fig. 2(a). Although several phases could be found under unique experimental conditions,[48] the α and β phases are more stable than the others (Fig. 2(b)), which can be viewed as single layers cleaved from the bulk (Fig. 3(a)). For example, α-P (black phosphorene) ML has already been exfoliated from bulk α phase along [001] direction[62] and β-P (blue phosphorene) ML can be grown along [0001] direction of bulk β phase by chemical vapor deposition (CVD).[63] Both α and β group-V films have been experimentally obtained,[6468] although the experimental evidence of β-P ML is still under debate.[69,70] Their dynamic stability has also been confirmed through phonon spectra as shown in Fig. 3(b), without any appreciable imaginary modes. In fact, most of group-V layers fabricated in experiment are α and β phases.[4,21,7173] These two phases are more popular and observed in experiment for all group-V elements.[74] Thus, we only focus on these two phases in the following.

Fig. 2 (a) Top views of 9 group-V ML allotropes with five typical honeycomb structures (α, β, γ, δ, ε) and four non-honeycomb structures (ζ, η, θ, ι). (b) Comparison of stabilities in terms of energies for different allotropes of group-V MLs.[59]
Fig. 3 (a) Top view and side view of α and β phases with two atomic layers (green and blue). The buckling height h is marked for α phase. When h is non-zero, the inversion symmetry of the structure breaks. (b) Phonon dispersions of group-VA MLs in α and β phases.[59]

For the β phase, all group-V MLs have buckled honeycomb lattices with two atomic layers, the same as silicene, which are not in the same plane due to the special valence electrons (s2p3) and mixed sp3 hybridization. There is buckling difference for different group-V MLs, but the space group (symmetry) is the same for all of them. For the α phase, all group-V MLs take the puckered lattice with orthogonal structure and there is also buckling difference (Fig. 3(a)) for different group-V MLs, e.g., the buckling height (h) of α-P is zero and that of α-Bi varies from zero to nonzero. Importantly, this nonzero buckling breaks the inversion symmetry of the α phase, quite different from the buckling in the β phase, and largely correlates to the topological and ferroelectric properties of the α phase, which will be discussed below.

3. Quantum spin Hall effect in 2D group-V monolayer

Topological insulators, as a new quantum state of matter, have attracted intensive research interest due to their applications in spintronics and topological quantum computation as well as academic importance as a novel quantum phenomenon.[75,76] Actually, topological properties in 3D group-V have attracted attention for more than ten years. The valence band of bulk Sb is equivalent to that of a strong topological insulator and bulk Bi is a conventional insulator due to the band inversion at L point. Both bulk Bi and Sb are semimetal with electron and hole pockets at different K points. Since they both have the rhombohedral structure, the Bi–Sb alloy is introduced to open a global bandgap in BZ and maintains the nontrival character.[77]

Two-dimensional TIs also named as QSH insulators are considered as more valuable than 3D TIs for spin transport applications, because the edge states in the former are more robust against back-scattering caused by non-magnetic defects than the surface states in the latter. QSH insulators possess edge states protected by the time reversal symmetry. As stated above, graphene was the first predicted QSH insulator. But, due to the lack of strong SOC, it is difficult to open up a bandgap for experimental observation in graphene. The 2D MLs of group-V elements (P, As, Sb, and Bi) are considered as new emerging 2D materials with much stronger SOC effect than graphene. Moreover, there are diversified ways to tune the topological character of 2D MLs of group-V elements and the cross-control of quantum states may be more easily accessed in the 2D systems. Therefore, as new members in the 2D material family, they can be potential candidates to study the QSH effect under appropriate conditions.

3.1. The α phase monolayer

Firstly, we discuss the α phase without buckling (h = 0 in Fig. 3(a)). In the inversion symmetric α phase, the structure has a non-symmorphic D2h (7) space group with two sublattices (A and B in Fig. 4(a)). Based on the tight-binding model, two types of Dirac fermions have been predicted in the 1st BZ as shown in Figs. 4(b) and 4(c).[78] A pair of type-I Dirac points locate along ΓX2 and are mainly determined by the pz orbital (Fig. 4(c)) between the two atomic layers, which are very sensitive to the interaction or distance between the layers. There are four type-II Dirac points protected by mirror symmetry, so that they carry opposite Chern numbers and appear in pairs (Fig. 4(b)). Type-II Dirac points are mainly determined by the px orbital (Fig. 4(d)) and sensitive to the in-plane strain. When a strain is applied, type I Dirac points only move along ΓX2 line while type II points are located at generic k-points. Thus, the α phase without buckling could hold six Dirac points in the 1st BZ, like graphene, although their space groups are different. The main difference is that the Dirac points in α phase can move in BZ without changing symmetry while they are pinned at K points in graphene. Therefore, a QSH insulator with large topological gap can also be realized in α phase, although a strain or external field may be needed for the phase transition. For example, intrinsic α-P is a normal insulator with a bandgap around Γ point and a large strain (∼ 16 %) along the zigzag direction is needed for the emergence of type-II Dirac points. While, only 6 % strain is needed for the phase transition of α-As.[79] Besides the strain modulation, application of electric field is also proposed to tune the bandgap and create the type-I Dirac points. In Fig. 5, an electric field (0.3 V/Å) perpendicular to the 2D plane is applied to close the bandgap of 4 ML α-P and causes the topological phase transition (0.55 V/Å for 3 ML).[80] Adsorption of alkali atom on the surface of α-P has similar effect as electric field and Dirac states have been observed in experiment by surface doping of potassium.[81] The structure parameters and electronic coupling between atoms of other group-V elements are different from those of α-P and the critical values for the topological phase transition are also different. Especially, the interlayer coupling of α-Bi is already weak enough and it is an intrinsic non-trivial 2D TI with large SOC gap[78] (gapped type-I Dirac points as aforementioned). More importantly, the large-scale growth of α-Bi has been realized[65] and the evidence of topological edge state along the edge of α-Bi has also been observed in experiment.[73] Thus, α-Bi is a good elemental 2D material with intrinsic topological properties to investigate different kinds of quantum states controlled by thickness, gating, and substrate interaction. Very recently, α-Sb has been successfully grown on Td-WTe2 substrate[74] and another experimental work claimed that 2 ML or 4 ML α-Sb becomes non-trivial due to the layer stacking.[82]

Fig. 4 (a) Basis sites in a unit cell of centrosymmetric puckered α structure. (b) 2D rectangular BZ with Dirac points marked by red dots. A pair of type-I Dirac points are marked as D and D′ on ΓX2, while two pairs of type-II Dirac points are marked (red dots) near F and F′ on ΓX1. (c) Schematic energy dispersions around D and F points respectively illustrating type-I and type-II Dirac points. (d) Band structures of 2D phosphorene (α phase with h = 0) type lattice structure of group-V elements in the absence of SOC, the red and green dots size shows the projection on the pz (px) atomic orbitals.[78]
Fig. 5 (a) Band structure of phosphorene layers (4 L) with applied external electric field of (a) 0.3 V/Å and (b) 0.45 V/Å without including SOC. The red and blue lines show the top valence and bottom conduction states, respectively. (c) Inversion energy Δinv and (d) Z2 invariant of 2-, 3-, and 4-layer phosphorene as a function of electric field. The negative values of Δinv show non-trivial topological phases. It can be seen that the phase transition takes place for 3- and 4-layer phosphorene at a critical field of 0.55 V/Å and 0.3 V/Å, respectively.[80]

As mentioned above, the atomic buckling in α phase is related to the inversion symmetry and quite important for topological properties. Except for α-P, the atomic buckling (h) of other group-V elements could be nonzero and different from each other due to different hybridization between two atomic layers or substrate effect. This buckling breaks the inversion symmetry and degeneracy of crossing points, and even causes the topological phase transition. This effect is most remarkable in α-Bi. Because of the negligible s–p hybridization and the formation of lone pair,[43] the spontaneous buckling height of α-Bi can vary in a wide range (0.0–0.5 Å). The different buckling height results in different pz interaction and thus changes the topological gap. The non-trivial QSH effect (phase transition) is determined by atomic buckling (h), as shown in Fig. 6.[73] The α-Bi is topologically non-trivial at low buckling heights (e.g., h < 0.1 Å ) and transforms to trivial insulator when h is large enough. Normally, atomic buckling is sensitive to charge ordering or charge transfer between substrate and epitaxial film. Thus, this buckling-dependent QSH effect provides a unique way to control the topological character at atomic level and very helpful for the design of topological device.

Fig. 6 (a) Side and top views of α-Bi ML with buckling. The upper and lower atomic layers are presented in blue and green colors, respectively. The buckling height h is defined by the distance between the atoms within the same ML. (b) The h-dependent energy gap (Δ) at Dirac point of α-Bi ML. (c) Band structures of non-buckled (h = 0) α-Bi without and with SOC as well as buckled α-Bi with SOC at different buckling heights (h = 0.1 Å and 0.5 Å).[73]
3.2. The β phase monolayer

Bismuth ML in β phase has attracted attention immediately after the discovery of graphene due to the similar honeycomb structure. Bismuth is the heaviest group-V element and actually, it is also the heaviest element in periodic table without radioactivity. So, it possesses large SOC effect and is the most promising candidate to achieve 2D QSH phase. In 2006, β-Bi ML had been predicted to be topologically non-trivial with protected edge states[83] and the non-trivial gap is large enough for experimental measurement. Several experimental works provide evidences of the topological edge states in β-Bi ML as shown in Fig. 7, where the dI/dV mapping shows the conducting states at the edges[84] and the STS profile crossing the edge shows a Dirac-point-like signal.[3,4]

Fig. 7 (a) Crystal structure of β-Bi. Red and blue colors label the upper and lower atomic layers of the topmost bilayer, respectively. Here, bilayer means two (upper and lower) atomic layers. (b) Topography around the hexagonal pit-like defect, differential conductance is shown with color plot, showing high conductance at alternative edges.[89] (c) STS of the step edge (blue) and the inner terrace (red) of Bi/Bi2Te3 island averaged over the area as marked in the inset of (c). The STS, the bands aligned Dirac point “DP” as indicated by the blue dashed line. The orange dashed lines mark the band gap of the topmost Bi layer.[3] (d) Differential conductivity dI/dV at different distances from the edge. A large gap of ∼ 0.8 eV is observed in bulk bismuthene (black curve). Inset shows the STS measurement locations at uphill substrate step causing the boundary.[4]

The β-Sb and β-As in ground state are trivial insulators. However, they can sustain large tensile strain (e.g., 18.4 % for As[85] and 18 % for Sb[86]) that assists the topological phase transition. It was proposed that the topological phase transition from normal insulator to 2D QSH insulator occurs under biaxial strain of 11.7 % and 14.5 % for Sb and As, respectively.[85,86] This large strain may be achieved by some substrate as buffer layer during epitaxial growth. As the second heaviest element of group-V, β-Sb could become 2D TI by other ways. It was predicted that β-Sb becomes 2D TI when the thickness increases between 4 ML and 7 ML, where the surface band splitting provides a robust QSH gap at room temperature.[87] With further increased film thickness, a crossover from the 2D TI to the 3D TI occurs.[87] On the other hand, coupling with substrate could also cause the topological phase transition in β-Sb and most recently, an experimental evidence of nontrivial topological edge states has been observed in epitaxially grown 2D β-Sb islands on copper Cu (111) substrate,[72] where the topological gap lies ∼ 1.5 eV above the Fermi level (Fig. 8). Similarly, β-P was also predicted to be a large indirect bandgap semiconductor ( > 2 eV),[58] and it has been proposed to be topologically non-trivial under 16 % tensile strain.[88]

Fig. 8 Band structure of freestanding (β phase) antimonene ML with and without SOC, represented by red and blue lines, respectively. (b) STM topography of a hexagonal honeycomb shaped antimonene island. The lower panel shows three dI/dV curves measured at the edges and center of the island marked by blue, green, and red arrows in the upper panel. (c) Plot of spatially resolved data of 35 dI/dV curves along the black dashed arrow in panel (b). The three dI/dV curves marked by arrows correspond to those shown in panel (d) with peaks at 1.68 eV (green), 1.85 eV (blue), and 1.95 eV (red). (d) The intensity plot of the dI/dV curves shown in panel (c).[72]

In general, crystal defects are really important for semiconductors, which could introduce additional defect states in bandgap affecting the transport properties. For topological materials, these defect states in bulk could introduce additional conducting states overcoming the advantage of the protected boundary states within bandgap. In comparison with compound systems, elemental systems are better with less defect problems, which is very important for the application of topological materials. This is one reason that researchers struggle to realize topological character in elemental systems. Recently, due to rising interest in QSH effect in group-V elemental MLs, 2D honeycomb structures of group-V binary compounds have also gained much attention,[90] e.g., SbAs, BiP, BiAs, BiSb, AsP, and SbP. Because all elements are group-V elements with isovalent electrons, there is no additional state appearing in the bandgap and the gap size can be tuned by applying a lattice strain. They all exhibit metal–indirect gap semiconductor–direct gap semiconductor–TI transitions under strain.[91] This broadens the potential candidates of 2D TIs in hexagonal β phase of group-V elements.

3.3. Chemical functionalization of group-V MLs

Chemical functionalization is an important tool to control electronic properties of materials and oxidation process is easy for group-V elements. This could change the topological character and help the emergence of new topological states. For example, β-P ML (blue phosphorus) is a trivial insulator with large bandgap. This gap size decreases largely after oxidation (P2O2) due to the saturation of the lone pair electrons. And, a moderate strain can easily tune the three bands around the Fermi energy, resulting in tunable quantum phase transition between pseudospin-1 fermions and 2D double Weyl fermions (Figs. 9(a)9(c)).[92,93] β-As ML also becomes a topological insulator exhibiting the quantum spin Hall effect when surface functionalized with H, F, OH, and CH3 chemical group.[94,95]

Fig. 9 (a) Electronic band structure of blue phosphorus oxide (BPO) without strain. A and B label the non-degenerate state at the conduction band bottom and the doubly-degenerate state at the valence band top, respectively. (b) Energies of A and B states versus strain. The crossing of the two corresponds to the quantum phase transition between a semiconductor phase and a symmetry protected semimetal (SSM) phase. (c) Energy dispersion around the band crossing point at Fermi level for critical strain +0.6 % (upper panel) and (lower panel) +4 % strain. The low-energy quasiparticles around the crossing point are described by 2D pseudospin-1 fermions and 2D double Weyl fermions.[92] Band structures of the (d) BiF and (e) SbF monolayers with (red) and without (gray) SOC. The bands around the Fermi level consist of the px and py orbitals. The size of the symbols is proportional to the population of the px and py orbitals.[96]

Hydrogenation or halogenation is another effective method to introduce topological phase transition in group-V MLs. Here, we firstly discuss the functionalized β phases. After functionalization, Bi2X2 or Sb2X2 (X = H, F, Cl, Br) were predicted to be topologically non-trivial insulators and stable above room temperature (Figs. 9(d) and 9(e)).[96] These features make Bi2X2 (or Sb2X2) an ideal platform for the realization of many striking quantum phenomena at room temperature. Similarly, halogen functionalized As2X2 also opens up a large non-trivial gap, signifying its potential for future applications as high-speed devices. Furthermore, if functionalized 2D TI is grown on the magnetic substrate, the quantum anomalous Hall insulator may be achieved, e.g., Sb2H2 on antiferromagnetic substrate LaFeO3 (111).[97] For the α phase of group-V, it is predicted that halogen-decorating on one side of MLs would generate a Dirac nodal line around the Fermi energy without SOC or hourglass fermions protected by nonsymmorphic symmetry when considering the SOC.[98]

As both α and β phases of group-V MLs have one lone pair of electrons, functionalization by monovalence chemical groups is normally preferred to maintain chemical stability with tunable electronic structures. Although chemical functionalization is a good way to change the electronic structure of 2D materials including the topological properties of 2D group-V MLs, this functionalization is permanent with fixed change of electronic structure. It is hard to reversibly tune the electronic structure by chemical reaction which mostly evolves irreversible bond formation. It is more desirable for device application to change or switch the electronic character between two or more topological states reversibly.

4. Ferroelectricity in 2D group-V materials

Ferroelectricity is a quantum order characterized by spontaneous alignment of electronic dipoles with switchable macroscopic electronic polarization. Ferroelectric materials with this switchable electrical polarization have attracted much research interest due to their applications in non-volatile memory devices, field effect transistors, sensors, solar cells, and so on.[37,38,99] Normally, ferroelectricity disappears when the film thickness is below a critical value because of the depolarizing electric field and electron screening. A new class of materials, called as hyperferroelectrics, has been identified, which can endure stable polar states in ultrathin atomic layers against unscreened depolarizing field.[39] This requires instability of a longitudinal optical mode in addition to the transverse-optical-mode instability of normal ferroelectrics. However, there is still lack of solid experimental evidence of these materials. Recently, the ferroelectric instability becomes of interest in 2D van der Waals materials. A lot of atomically thin ferroelectric materials have been predicted and experimentally realized in 2D van der Waals materials[100102] and readers may refer to two recent topic review papers.[41,46] Such depolarization field effect in a true 2D system may no longer be an issue. Moreover, ferroelectric transition driven by electric field is even observed in 2D semimetals,[103,104] expanding our understanding of ferroelectricity.[105] These results show again that dimensionality makes major impact on materials properties and underlying physics.

4.1. The α phase monolayer

By scanning tunneling microscope and transport measurement, the 2D ferroelectricity has been confirmed in group-IV monochalcogenides, SnTe ML, which has puckered atomic structure.[100] The key condition of this ferroelectricity is the resonance bonding between cations and anions composing the puckered structure.[106] After that, ferroelectricity has been identified in a lot of 2D systems by replacing elements with isoconvalent state as SnTe, e.g., MX (M = Ge, Sn; X = S, Se).[45] The α phase of 2D group-V elemental MLs has similar structure and the same valence electrons with 2D group-IV monochalcogenides. Thus, they are also possible to present ferroelectricity in puckered α phase.

Ferroelectricity has been mostly observed in compounds of different elements rather than elemental materials. The appearance of FE in elemental materials does not contradict the requirement of broken inversion symmetry. The major obstacle is that most nonmetallic elemental materials are stabilized in nonpolar structures of high symmetry, which prevents the electric polarization. This obstacle could be overcome in 2D system. As mentioned above, the atomic buckling in α phase breaks the inversion symmetry and very recently, Xiao et al. predicted the existence of FE in buckled α-As, α-Sb, and α-Bi elemental MLs for the first time, as shown in Figs. 10(a)10(c).[43] The spontaneous polarization in these buckled MLs is along the in-plane x-axis with a predicted high Curie temperature, which makes them promising candidates for ferroelectric devices. Further analysis of the free-energy contour and phonon spectrum confirmed that beyond the two FE phases (A and A′), there also exist anti-ferroelectric (AFE) metastable phases (C and C′) (Figs. 10(d) and 10(e)).

Fig. 10 (a) Top view of α phase group-V elemental ML. The unit cell is indicated with black dashed rectangle lines. (b) Free energy contour for α-As ML versus the buckling heights (hU, hL). The phases A, A′, and B are marked. (c) Side views of the two energy-degenerate distorted noncentrosymmetric structures (A and A′ phases) and undistorted centrosymmetric structure (phase B). The height differences between upper (red) and lower (blue) atomic layers are marked as hU and hL, respectively. (d) Side view of α-Bi ML in antiferroelectric (AFE) configuration. The atomic sites shifted up and down are colored in red and green, respectively. (e) Free energy contour of Bi MLs versus buckling heights (h(R), h(G)). For Bi, besides the FE phases A and A′, there also exist metastable AFE phases C and C′.[43]

Compared with the 2D FE group-IV monochalcogenide materials, besides the fundamental distinction (compound vs. elemental), the mechanisms for the ferroelectric polarization of 2D group-V MLs are also different. The group-IV monochalcogenide materials break the centrosymmetry via in-plane distortion from the cubic structure as well as the occupation of lattice sites with different kinds of elements. In comparison, the group-V MLs break centrosymmetry by the out-of-plane atomic-layer buckling, whereas the in-plane distortion does not break the symmetry. As a result, the transformation between two degenerate FE structures of group-IV monochalcogenide materials requires the breaking and reforming bonds between atoms. This is the reason that the resonance bonding between atoms is important and required for the generation of ferroelectronic polarization in group-IV monochalcogenide compounds.[107] Whereas for the group-V monolayers, there is no significant bonding and bond-breaking character in the transformation between two degenerate FE structures. Moreover, the puckered structure of the elemental MLs permits the presence of an AFE phase (for α-Bi), which is absent in the compound materials. The coexistence of FE and AFE phases in the same materials is important. As the symmetry and geometrical structure of AFE phase are different from those of FE phase, its electronic structures could be largely changed in topological nature. This interplay between geometrical phase transition and electronic polarization allows an extraordinary control of the electronic structure (e.g., electronic topological properties) via external field. In addition, the electronic polarization is sensitive to buckling height h and interaction with the substrate, giving us more freedom to engineer the electronic structures and phase transition of the van der Waals heterostructure devices based on group-V elemental MLs.

It has also been predicted that 2D group-V binary compounds composed of both light and heavy group-V elements present ferroelectricity in α or γ phase with notable large and strain tunable spontaneous polarization, e.g., BiN,[107] SbN, BiP.[108] SbP, and SbAs.[109] Such large polarization is caused by the immense electronegativity difference of the light and heavy group-V elements and their displacement in the polarization plane. The mechanism for the ferroelectric transition is almost the same as that of 2D group-IV monochalcogenide compounds requiring the breaking and reforming bonds between atoms. In addition, ferroelectricity is also predicted in halogen-intercalated α-P bilayer without buckling, where the ferroelectric transition is determined by the shift of the halogen atoms covalently bonded to up-layer or down-layer.[110] This concept may be extended to other van der Waals bilayers, but it is really challenging in experiment to precisely functionalize the target material to control these quantum orders.

4.2. The β phase monolayer

In hexagonal honeycomb β phase, the out-of-plane atomic-layer buckling does not break the inversion symmetry and the group-V elemental MLs remain centrosymmetric without ferroelectricity no matter whether they are planar or buckled 2D layers. One possible way to break centrosymmetry is to fabricate AB binary monolayer with a bipartite corrugated honeycomb lattice, where the A and B atoms are different elements. Binary group-V compounds show a spontaneous FE polarization in freestanding buckled β phase and their polarization is switched by changing the relative position of the two atoms along out-of-plane direction. The presence of a diatomic basis that breaks the inversion symmetry even in the planar geometry leads to the emergence of Zeeman-like spin-split bands with coupled spin-valley physics analogous to the MoS2 case.[111] Meanwhile, it is mainly responsible for the onset of the FE phase when the honeycomb lattice buckles, allowing for an electrically controllable Rashba-like spin-texture around the valleys, whose chirality is locked to the polarization direction and therefore fully reversible upon FE switching.[42] These spin split bands can be effectively detected by spin-resolved spectroscopic techniques and binary β phase compounds allow for the engineering of 2D spin field-effect-transistors.[112]

Chemical functionalization of polar chemical group is another direct and possible way to transform centrosymmetric β phase into polar structure. It is proposed that ferroelectricity can be introduced into β-As ML by passivation of CH2OCH3 ligand.[90] The fully relaxed structure of CH2OCH3 functionalized β-As is shown in the inset of Fig. 11(a), where arsenene is flat without any out-of-plane buckling and remains centrosymmetry. When CH2OCH3 is adsorbed on β-As ML, it becomes polarized with local dipole. The in-plane electronic polarization is achieved by ligand molecules alignment and a ferroelectric transition can be driven by a ligand rotation mechanism (Fig. 11(b)). Actually, this concept can date back to an early work[114] and is expected to be a more general phenomenon in 2D-layered materials. Similar to β-As passivated by simple ligands as mentioned before, CH2OCH3-functionalized β-As becomes a topologically nontrivial insulator with QSH effect. The electronic structure and topological character are mainly determined by the bonding between β-As and ligand, not the relative position or rotation of CH2OCH3 sitting on top of the As atoms. Thus, the topological phase and nontrivial band gap are robust and well preserved when ligand CH2OCH3 molecules are rotated (Fig. 11(c)) without interplay between the topological character and ferroelectricity. On the other hand, the total energies of multiple rotated configurations with different polarizations are quite close to each other, which means that the ligand molecules may not be orderly arranged as shown in Fig. 11(a). However, this controllable in-plane ferroelectricity maybe has wide applications in memory or logical device designs and the storage density can be quite high as the information bit can be directly related to one ligand arrangement.

Fig. 11 (a) Energy variation and kinetic barrier for CH2OCH3 rotations in functionalized arsenene; the insets show the structural models. (b) Variation of the polarization along the x and y directions. (c) Electronic band structures of the symmetric ligand configuration with the rotation angle of 0° (left) and the the asymmetric ligand configuration with the rotation angle of 120° (right). The topological order is preserved under the ligand rotation, although there is a band splitting due to an out-of-plane symmetry breaking.[113]
5. Relationship of FE and TI

As both ferroelectricity and electronic topological character are closely linked to the geometrical symmetry and structural distortion, the interplay between them can be mediated by geometrical distortion and their coexistence could enable precise control of electronic topology using electric fields (determined by geometrical distortion in phase transition), enhancing the applicability of topological materials in electronics and spintronics. Unfortunately, the crystal structure of 3D group-V is the centrosymmetric one without electronic polarization. Very recently, it is found that both electronic polarization and nontrivial topology can coexist in some compounds containing group-V elements.[115117] An extremely rich physics strongly links to the ferroelectric properties, ranging from electric control of bulk Rashba effect to the transitions among a 3D topological insulator, normal insulator, and Weyl semimetals. It is also proposed to realize topologically protected Weyl semimetal and nodal line semimetal phases in IV–VI semiconducting compounds (isovalent as group-V in average) with the low-temperature ferroelectric phase.[118,119] These pave a way to explore the coupling between topological properties and electronic polarization in low dimensional systems.

There is more freedom allowing crystal-symmetry variation and distortion in 2D group-V MLs, which is important for the emergence of multiple quantum ordering. In puckered α phase, out-of-plane atomic-layer buckling is flexible and can vary in a wide range. This buckling breaks centrosymmetry and also changes the electronic topological character. For α-Bi, it is an intrinsic 2D TI with QSH effect when the buckling is zero. This corresponds to a paraelectric state. When the buckling is strong and large ( > 0.1 Å), α-Bi becomes a normal insulator without centrosymmetry and corresponds to a ferroelectric state. Thus, the topological phase transition with and without topological protected boundary states may be controlled by the ferro–paraelectric phase transition through geometrical distortion. It is interesting that when the buckling is weak ( > 0.1 Å) under strain, the QSH effect is persisted and coexists with FE in this single phase material. As there are two ferroelectric states with opposite polarizations and they are related to each other via an inversion operation, the chirality and direction of the spin-locked current at boundary rely on the direction of the ferroelectric polarization, as shown in Figs. 12(a) and 12(c). Thus, it may promote novel applications in conceptually new devices, where the direction of spin current is fully controlled by ferroelectricity. Since the direction of spin-locked current can be viewed as a ferroic order, this is some kind of multiferroics.

Fig. 12 Schematic representation of ferroelectric topological insulators (a) with two polar states and (b) with two polar TI states and an antipolar normal insulator (NI) state. (c) Proposed switching device corresponding to the two polar TI states with ferroelectric polarization. (d) Switching boundary conductivity on/off of an antipolar NI and two polar TI states. The chirality of electronic current is labeled by red and blue lines respectively. The up and down spins are indicated by short black arrows on the lines.

One the other hand, AFE could coexist with FE in 2D group-V elemental ML. The crystal symmetry and distortion of AFE are distinctly different from those of FE, resulting in different electronic structure and topological character. For example, AFE and FE states may correspond to an antipolar normal insulator (NI) state and two polar topological insulator (TI) states, respectively, as shown in Fig. 12(b). There is no protected boundary state within the bandgap of NI, while the dissipationless boundary states are expected at the edges of TI which are protected by the topological nature. It could enable precise control of topology and switching on/off the topological boundary states using electric fields (Fig. 12(d)), enhancing the applicability of topological materials in electronics and spintronics.

6. Perspective

The recent progress in QSH insulators, 2D ferroelectricity, and experimental fabrication of group-V elemental MLs has motivated the researchers to revisit group-V materials. They are potential candidates for realization of 2D TIs and their topological properties can be controlled by electric field, strain, substrate effect, or chemical functionalization. Recent experimental evidences of non-trivial topology in 2D Bi and Sb have robustly boosted the research interest. Moreover, group-V MLs are the first predicted ferroelectric elemental materials in 2D limit, which violate the traditional concept and broaden the 2D ferroelectric family.

Ferroelectricity and QSH effect are two intriguing quantum phenomena in modern condensed matter physics and both of them are related to the integration of Berry phase in BZ. The co-existing QSH effect and ferroelectricity in the same material may lead to novel physical properties, e.g., topological magneto-electric effect.[120] The effective integration of these two properties is not only important for the novel applications in devices, but also leads to the discoveries of unusual quantum phenomena, e.g., the formation of quantum Hall ferroelectric states[121,122] and nonlinear Hall effect.[123,124] In addition, as 2D materials, they can easily grow or deposit on substrates with other quantum properties. Actually, it has been confirmed in experiment that superconductivity can be introduced into 2D Bi ML by deposition on a superconducting substrate due to the superconductor proximity effect.[125] Therefore, 2D group-V MLs provide a good platform to study the interplay between multiple quantum properties.

Reference
[1] Liu C C Feng W Yao Y 2011 Phys. Rev. Lett. 107 076802
[2] Sheng D N Weng Z Y Sheng L Haldane F D M 2006 Phys. Rev. Lett. 97 036808
[3] Murakami S 2006 Phys. Rev. Lett. 97 236805
[4] Reis F Li G Dudy L Bauernfeind M Glass S Hanke W Thomale R 2017 Science 357 287
[5] Wu M Dong S Yao K Liu J Zeng X C 2016 Nano Lett. 16 7309
[6] Baskaran G 2018 Many-body Approaches at Different Scales: A Tribute to Norman H. March on the Occasion of his 90th Birthday Angilella G G N Amovilli C Cham Springer International Publishing 43
[7] Liao M Zang Y Guan Z Li H Gong Y Zhu K Hu X P Zhang D Xu Y Wang Y Y He K Ma X C Zhang S C Xue Q K 2018 Nat. Phys. 14 344
[8] Xing Y Fu H L Liu H Sun Y Wang F Lin X Wang J Zhang H M Peng J P Ma X C Xue Q K 2015 Science 350 542
[9] Gibertini M Koperski M Morpurgo A F Novoselov K S 2019 Nat. Nanotechnol. 14 408
[10] Gong C Li L Li Z Ji H Stern A Xia Y Cao T Bao W Wang C Wang Y Qiu Z Q Cava R J Louie S G Xia J Zhang X 2017 Nature 546 265
[11] Huang B Clark G Navarro-Moratalla E Klein D R Cheng R Seyler K L Zhong D Schmidgall E McGuire M A Cobden D H Yao W Xiao D Jarillo-Herrero P Xu X 2017 Nature 546 270
[12] Novoselov K S Geim A K Morozov S V Jiang D Katsnelson M I Grigorieva I V Dubonos S V Firsov A A 2005 Nature 438 197
[13] Geim A K Novoselov K S 2007 Nat. Mater. 6 183
[14] Feng B Ding Z Meng S Yao Y He X Cheng P Chen L Wu K 2012 Nano Lett. 12 3507
[15] Fleurence A Friedlein R Ozaki T Kawai H Wang Y Yamada-Takamura Y 2012 Phys. Rev. Lett. 108 245501
[16] Vogt P De Padova P Quaresima C Avila J Frantzeskakis E Asensio M C Resta A Ealet B Le Lay G 2012 Phys. Rev. Lett. 108 155501
[17] Li L Lu S Z Pan J Qin Z Wang Y Q Wang Y Cao G Y Du S Gao H J 2014 Adv. Mater. 26 4820
[18] Zhang L Bampoulis P Rudenko A N Yao Q Van Houselt A Poelsema B Katsnelson M I Zandvliet H J W 2016 Phys. Rev. Lett. 116 256804
[19] Zhu F Chen W Xu Y Gao C Guan D Liu C Qian D Zhang S C Jia J 2015 Nat. Mater. 14 1020
[20] Li W Kong L Chen C Gou J Sheng S Zhang W Li H Chen L Cheng P Wu K 2018 Sci. Bull. 63 282
[21] Li L Yu Y Ye G J Ge Q Ou X Wu H Feng D Chen X H Zhang Y 2014 Nat. Nanotechnol. 9 372
[22] Vishnoi P Mazumder M Pati S K and R Rao C N 2018 New J. Chem. 42 14091
[23] Shao Y Liu Z L Cheng C Wu X Liu H Liu C Wang J O Zhu S Y Wang Y Q Shi D X Ibrahim K Sun J T Wang Y L Gao H J 2018 Nano Lett. 18 2133
[24] Huang X Guan J Lin Z Liu B Xing S Wang W Guo J 2017 Nano Lett. 17 4619
[25] Qin J Qiu G Jian J Zhou H Yang L Charnas A Zemlyanov D Y Xu C Y Xu X Wu W Wang H Ye P D 2017 ACS Nano 11 10222
[26] Kane C L Mele E J 2005 Phys. Rev. Lett. 95 226801
[27] Chen Y L Analytis J G Chu J H Liu Z K Mo S K Qi X L Zhang H J Lu D H Dai X Fang Z Zhang S C Fisher I R Hussain Z Shen Z X 2009 Science 325 178
[28] Hsieh D Xia Y Qian D Wray L Meier F Dil J H Osterwalder J Patthey L Fedorov A V Lin H Bansil A Grauer D Hor Y S Cava R J Hasan M Z 2009 Phys. Rev. Lett. 103 146401
[29] Qi X L Zhang S C 2011 Rev. Mod. Phys. 83 1057
[30] Zhang S Zhu W Sun Q 2013 J. Phys. Condens. Matter 25 295301
[31] Yao Y Ye F Qi X L Zhang S C Fang Z 2007 Phys. Rev. 75 041401
[32] Xu Y Yan B Zhang H J Wang J Xu G Tang P Duan W Zhang S C 2013 Phys. Rev. Lett. 111 136804
[33] Cahangirov S Topsakal M Aktürk E Şahin H Ciraci S 2009 Phys. Rev. Lett. 102 236804
[34] Vishnoi P Pramoda K Rao C N R 2019 ChemNanoMat 5 1062
[35] Auciello O Scott J F Ramesh R 1998 Phys. Today 51 22
[36] Nuraje N Su K 2013 Nanoscale 5 8752
[37] Dawber M Rabe K M Scott J F 2005 Rev. Mod. Phys. 77 1083
[38] Scott J F 2007 Science 315 954
[39] Garrity K F Rabe K M Vanderbilt D 2014 Phys. Rev. Lett. 112 127601
[40] Wu M Jena P 2018 Wiley Interdiscip. Rev. Comput. Mol. Sci. 8 e1365
[41] Guan Z Hu H Shen X Xiang P Zhong N Chu J Duan C 2020 Adv. Electron. Mater. 6 1900818
[42] Di Sante D Stroppa A Barone P Whangbo M H Picozzi S 2015 Phys. Rev. 91 161401
[43] Xiao C Wang F Yang S A Lu Y Feng Y Zhang S 2018 Adv. Funct. Mater. 28 1707383
[44] Wu M Zeng X C 2016 Nano Lett. 16 3236
[45] Fei R Kang W Yang L 2016 Phys. Rev. Lett. 117 097601
[46] Lu C Wu M Lin L Liu J M 2019 Natl. Sci. Rev. 6 653
[47] Liu J Fang C Fu L 2019 Chin. Phys. 28 047301
[48] Zhang L Huang H Zhang B Gu M Zhao D Zhao X Li L Zhou J Wu K Cheng Y Zhang J 2019 Angew. Chem. Int. Ed. 58 1074
[49] Brown A Rundqvist S 1965 Acta Crystallogr. 19 684
[50] Bridgman P W 1914 J. Am. Chem. Soc. 36 1344
[51] Hultgren R Gingrich N S Warren B E 1935 J. Chem. Phys. 3 351
[52] Ruck M Hoppe D Wahl B Simon P Wang Y Seifert G 2005 Angew. Chem. Int. Ed. 44 7616
[53] Jacobs R B 1937 J. Chem. Phys. 5 945
[54] Jamieson J C 1963 Science 139 1291
[55] Saito M Takemori Y Hashi T Nagao T Yaginuma S 2007 Jpn. J. Appl. Phys. 46 7824
[56] Namari N A P Saito M 2019 Jpn. J. Appl. Phys. 58 061003
[57] Burford N Carpenter Y Y Conrad E Saunders C D L 2010 Biol. Chem. Arsenic Antimony Bismuth 1
[58] Zhu Z Tománek D 2014 Phys. Rev. Lett. 112 176802
[59] Zhang S Xie M Li F Yan Z Li Y Kan E Liu W Chen Z Zeng H 2016 Angew. Chem. Int. Ed. 55 1666
[60] Wu M Fu H Zhou L Yao K Zeng X C 2015 Nano Lett. 15 3557
[61] Zhuo Z Wu X Yang J 2016 J. Am. Chem. Soc. 138 7091
[62] Liu H Neal A T Zhu Z Luo Z Xu X Tománek D Ye P D 2014 ACS Nano 8 4033
[63] Zhang J L Zhao S Han C Wang Z Zhong S Sun S Guo R Zhou X Gu C D Di K Li Z Chen W 2016 Nano Lett. 16 4903
[64] Gibaja C Rodriguez-San-Miguel D Ares P Gómez-Herrero J Varela M Gillen R Maultzsch J Hauke F Hirsch A Abellán G Zamora F 2016 Angew. Chem. Int. Ed. 55 14345
[65] Yang Z Wu Z Lyu Y Hao J 2019 InfoMat 1 98
[66] Shi Z Q Li H Yuan Q Q Song Y H Lv Y Y Shi W Jia Z Y Gao L Chen Y B Zhu W Li S C 2019 Adv. Mater. 31 1806130
[67] Yuhara J Fujii Y Nishino K Isobe N Nakatake M Xian L Rubio A Le Lay G 2018 2D Mater. 5 025002
[68] Nagao T Sadowski J Saito M Yaginuma S Fujikawa Y Kogure T Ohno T Hasegawa Y Hasegawa S Sakurai T 2004 Phys. Rev. Lett. 93 105501
[69] Xu J P Zhang J Q Tian H Xu H Ho W Xie M 2017 Phys. Rev. Mater. 1 061002
[70] Zhuang J Liu C Gao Q Liu Y Feng H Xu X Wang J Zhao J Dou S X Hu Z Du Y 2018 ACS Nano 12 5059
[71] Lu W Nan H Hong J Chen Y Zhu C Liang Z Ma X Ni Z Jin C Zhang Z 2014 Nano Res. 7 853
[72] Zhu S Y Shao Y Wang E Cao L Li X Y Liu Z L Liu C Liu L W Wang J O Ibrahim K Sun J T Wang Y L Du S Gao H J 2019 Nano Lett. 19 6323
[73] Lu Y Xu W Zeng M Yao G Shen L Yang M Luo Z Pan F Wu K Das T He P Jiang J Martin J Feng Y P Lin H Wang X 2015 Nano Lett. 15 80
[74] Zhang S Guo S Chen Z Wang Y Gao H Gómez-Herrero J Ares P Zamora F Zhu Z Zeng H 2018 Chem. Soc. Rev. 47 982
[75] Hasan M Z Kane C L 2010 Rev. Mod. Phys. 82 3045
[76] Dai X Le C C Wu X X Qin S S Lin Z P Hu J P 2016 Chin. Phys. Lett. 33 127301
[77] Fu L Kane C L 2007 Phys. Rev. 76 045302
[78] Lu Y Zhou D Chang G Guan S Chen W Jiang Y Jiang J Wang X Yang S A Feng Y P Kawazoe Y Lin H 2016 npj Comput. Mater. 2 16011
[79] Kamal C Ezawa M 2015 Phys. Rev. 91 085423
[80] Liu Q Zhang X Abdalla L B Fazzio A Zunger A 2015 Nano Lett. 15 1222
[81] Kim J Baik S S Ryu S H Sohn Y Park S Park B G Denlinger J Yi Y Choi H J Kim K S 2015 Science 349 723
[82] Märkl T Kowalczyk P J Le Ster M Mahajan I V Pirie H Ahmed Z Bian G Wang X Chiang T C Brown S A 2017 2D Mater. 5 011002
[83] Hirahara T Nagao T Matsuda I Bihlmayer G Chulkov E V Koroteev Y M Echenique P M Saito M Hasegawa S 2006 Phys. Rev. Lett. 97 146803
[84] Drozdov I K Alexandradinata A Jeon S Nadj-Perge S Ji H Cava R J Andrei Bernevig B Yazdani A 2014 Nat. Phys. 10 664
[85] Zhang H Ma Y Chen Z 2015 Nanoscale 7 19152
[86] Zhao M Zhang X Li L 2015 Sci. Rep. 5 16108
[87] Zhang P Liu Z Duan W Liu F Wu J 2012 Phys. Rev. 85 201410
[88] Yang G Xu Z Liu Z Jin S Zhang H Ding Z 2017 J. Phys. Chem. 121 12945
[89] Drozdov I K Alexandradinata A Jeon S Nadj-Perge S Ji H Cava R J Bernevig B A Yazdani A 2014
[90] Guo S Zhang Y Ge Y Zhang S Zeng H Zhang H 2019 Adv. Mater. 31 1902352
[91] Nie Y Rahman M Wang D Wang C Guo G 2016 Sci. Rep. 5 17980
[92] Zhu L Wang S S Guan S Liu Y Zhang T Chen G Yang S A 2016 Nano Lett. 16 6548
[93] Hua C Li S Xu Z Zheng Y Yang S A Lu Y 2020 Adv. Sci. 7 1901939
[94] Zhao J Li Y Ma J 2016 Nanoscale 8 9657
[95] Wang Y Ji W Zhang C Li P Li F Ren M Chen X L Yuan M Wang P 2016 Sci. Rep. 6 20342
[96] Song Z Liu C C Yang J Han J Ye M Fu B Yang Y Niu Q Lu J Yao Y 2014 NPG Asia Mater. 6 e147
[97] Zhou T Zhang J Xue Y Zhao B Zhang H Jiang H Yang Z 2016 Phys. Rev. 94 235449
[98] Jin K H Huang H Wang Z Liu F 2019 Nanoscale 11 7256
[99] Wang Y Xiao C Chen M Hua C Zou J Wu C Jiang J Yang S A Lu Y Ji W 2018 Mater. Horizons 5 521
[100] Chang K Liu J Lin H Wang N Zhao K Zhang A Jin F Zhong Y Hu X Duan W Zhang Q Fu L Xue Q K Chen X Ji S H 2016 Science 353 274
[101] Liu F You L Seyler K L Li X Yu P Lin J Wang X Zhou J Wang H He H Pantelides S T Zhou W Sharma P Xu X Ajayan P M Wang J Liu Z 2016 Nat. Commun. 7 12357
[102] Xiao J Zhu H Wang Y Feng W Hu Y Dasgupta A Han Y Wang Y Muller D A Martin L W Hu P Zhang X 2018 Phys. Rev. Lett. 120 227601
[103] Sharma P Xiang F X Shao D F Zhang D Tsymbal E Y Hamilton A R Seidel J 2019 Sci. Adv. 5 eaax5080
[104] Fei Z Zhao W Palomaki T A Sun B Miller M K Zhao Z Yan J Xu X Cobden D H 2018 Nature 560 336
[105] Tu Z Wu M Zeng X C 2017 J. Phys. Chem. Lett. 8 1973
[106] Kooi B J Noheda B 2016 Science 353 221
[107] Chen P Zhang X J Liu B G 2019 ACS Appl. Nano Mater. 2 58
[108] Liu C Wan W Ma J Guo W Yao Y 2018 Nanoscale 10 7984
[109] Shen S Liu C Ma Y Huang B Dai Y 2019 Nanoscale 11 11864
[110] Yang Q Xiong W Zhu L Gao G Wu M 2017 J. Am. Chem. Soc. 139 11506
[111] Xiao D Liu G B Feng W Xu X Yao W 2012 Phys. Rev. Lett. 108 196802
[112] Di Sante D Barone P Bertacco R Picozzi S 2013 Adv. Mater. 25 509
[113] Kou L Ma Y Liao T Du A Chen C 2018 Phys. Rev. Appl. 10 024043
[114] Wu M Burton J D Tsymbal E Y Zeng X C Jena P 2012 J. Am. Chem. Soc. 134 14423
[115] Monserrat B Bennett J W Rabe K M Vanderbilt D 2017 Phys. Rev. Lett. 119 036802
[116] Di Sante D Barone P Stroppa A Garrity K F Vanderbilt D Picozzi S 2016 Phys. Rev. Lett. 117 076401
[117] He J Di Sante D Li R Chen X Q Rondinelli J M Franchini C 2018 Nat. Commun. 9 492
[118] Lau A Ortix C 2019 Phys. Rev. Lett. 122 186801
[119] Sun L Wan S L 2015 Chin. Phys. Lett. 32 057501
[120] Qi X Li R Zang J Zhang S 2009 Science 323 1143f
[121] Hu X Pang Z Zhang C Wang P Li P Ji W 2019 J. Mater. Chem. 7 9406
[122] Kou L Fu H Ma Y Yan B Liao T Du A Chen C 2018 Phys. Rev. 97 075429
[123] Kang K Li T Sohn E Shan J Mak K F 2019 Nat. Mater. 18 324
[124] Du Z Z Wang C M Li S Lu H Z Xie X C 2019 Nat. Commun. 10 3047
[125] Sun H H Wang M X Zhu F Wang G Y Ma H Y Xu Z A Liao Q Lu Y Gao C L Li Y Y Liu C Qian D Guan D Jia J F 2017 Nano Lett. 17 3035