Reversible doping polarity and ultrahigh carrier density in two-dimensional van der Waals ferroelectric heterostructures

Yanyan Li, Mingjun Yang, Yanan Lu, Dan Cao, Xiaoshuang Chen, Haibo Shu

Front. Phys. ›› 2023, Vol. 18 ›› Issue (3) : 33307.

PDF(8013 KB)
Front. Phys. All Journals
PDF(8013 KB)
Front. Phys. ›› 2023, Vol. 18 ›› Issue (3) : 33307. DOI: 10.1007/s11467-022-1244-4
RESEARCH ARTICLE
RESEARCH ARTICLE

Reversible doping polarity and ultrahigh carrier density in two-dimensional van der Waals ferroelectric heterostructures

Author information +
History +

Abstract

Van der Waals semiconductor heterostructures (VSHs) composed of two or more two-dimensional (2D) materials with different band gaps exhibit huge potential for exploiting high-performance multifunctional devices. The application of 2D VSHs in atomically thin devices highly depends on the control of their carrier type and density. Herein, on the basis of comprehensive first-principles calculations, we report a new strategy to manipulate the doping polarity and carrier density in a class of 2D VSHs consisting of atomically thin transition metal dichalcogenides (TMDs) and α-In2X3 (X = S, Se) ferroelectrics via switchable polarization field. Our calculated results indicate that the band bending of In2X3 layer driven by the FE polarization can be utilized for engineering the band alignment and doping polarity of TMD/In2X3 VSHs, which enables us to control their carrier density and type of the VSHs by the orientation and magnitude of local FE polarization field. Inspired by these findings, we demonstrate that doping-free p−n junctions achieved in MoTe2/In2Se3 VSHs exhibit high carrier density (1013−1014 cm−2), and the inversion of the VHSs from n−p junctions to p−i−n junctions has been realized by the polarization switching from upward to downward states. This work provides a nonvolatile and nondestructive doping strategy for obtaining programmable p−n van der Waals (vdW) junctions and opens the possibilities for self-powered and multifunctional device applications.

Graphical abstract

Keywords

van der Waals heterostructures / ferroelectric polarization / carrier type / band alignment / density-functional theory

Cite this article

Download citation ▾
Yanyan Li, Mingjun Yang, Yanan Lu, Dan Cao, Xiaoshuang Chen, Haibo Shu. Reversible doping polarity and ultrahigh carrier density in two-dimensional van der Waals ferroelectric heterostructures. Front. Phys., 2023, 18(3): 33307 https://doi.org/10.1007/s11467-022-1244-4

1 Introduction

Semiconductor heterostructures (SHSs) are a basic building block for electronic and optoelectronic devices since they not only combine the properties and functionalities of constituent materials but also exhibit various exotic effects due to the interfacial interactions [1-3]. Traditional synthetic methods of SHSs, such as chemical and physical epitaxial growth, suffer from the limitation of lattice mismatch that easily leads to the formation of interface defects and lattice disorders, which greatly restricts their device applications [4, 5]. Emerging 2D layered semiconductor materials, such as transition metal dichalcogenides (TMD) [6,7], black phosphorus (BP) [8], and group-VA elements and compounds [9], provide a state-of-the-art approach to design SHSs via van der Waals (vdW) integration. Owing to the dangling-bond-free surface with weak interlayer interactions, the integration of these 2D materials into vdW semiconductor heterostructures (VSHs) without the limitation of lattice mismatch. Hence, VSHs offer a large degree of freedom for the device design with desired interface structures and band alignments and become an ideal platform for exploring novel physical properties [10,11]. A typical example is twisted TMD VSHs that possess unconventional moiré excitons and charge density wave (CDW) state [12,13]. Great achievements have been made in the last years on the synthesis of 2D VSHs, such as MoS2/WS2 [14], MoSe2/WSe2 [15], SnSe/MoS2 [16], and SnS/SnS2 [17], and BP/MoS2 [18]. 2D VSHs exhibit huge potential for nanoelectronics and optoelectronics, including light emitting diodes, field-effect transistors, photodetectors, memories, lasers, and solar cells, etc. [19,20].
The device applications of 2D VSHs depend on the manipulation of the type, density, and spatial distribution of charge carriers. The impurity doping is the most common way to produce charge carriers by means of intentional introduction of impurity atoms into the host lattice. However, this approach implemented in 2D semiconductors remains a large challenge due to their ultrathin atomic structures and limited physical space [21]. In particular for the heavy doping, excessive incorporation of impurities into 2D host lattices easily causes severe structural damage and lattice disorder [22]. Therefore, various new doping strategies were developed to program the carrier type and concentration of 2D materials, including chemical intercalation [23], surface charge transfer [24], electron-beam irradiation [25], laser light illumination [26], and electrostatic doping [27, 28]. However, most of these doping methods inevitably introduce chemical species or induce the formation of lattice defects [23-26], consequently resulting in the degeneration of device performance. Moreover, the polarity and transfer characteristics of p−n junctions are difficult to be changed once these doping methods have been implemented in 2D semiconductors and VSHs [29]. Although the gate-tuned electrostatic doping method provides a flexible way to tune the charge carriers, it requires an additional voltage to maintain the type and density of carriers therefore is volatile [30]. Hence, the development of nonvolatile and nondestructive doping ways to realize p-n junctions in atomically thin semiconductor materials is highly desired.
Because of atomic-scale thickness and ultraflat surface, the band structures of 2D vdW crystals and heterostructures are sensitive to the modulation of external and local electric fields. Considering that an external electric field cannot provide a nonvolatile doping way for the control of charge carriers, the creation of a local electric field is an alternative strategy to achieve this goal. Ferroelectric materials possess a spontaneous polarization that can be reversed by an electric field stimulus. Therefore, the integration of 2D semiconductors on ferroelectric (FE) substrates create a local and nonvolatile electric polarization field at the interface [31], which can be used to modulate the carrier type and density of 2D channel materials without sacrifice of their intrinsic physical properties. For example, the selective p- and n-type doping in atomically thin TMDs such as MoS2, MoTe2, and WSe2 have been demonstrated on a series of FE substrates, including BiFeO3 [32], PbTiZrO3 [33], LiNbO3 [34], Na0.5Bi4.5Ti4O15 [35], and P(VDF-FrFE) [36,37]. However, most of FE materials require a large thin-film thickness to maintain their polarization due to the depolarization field effect [38], which is incompatible with the trend of device minimization. Moreover, the insulator nature of FE oxides with large band gap (> 3.5 eV) [39] makes them difficult to be integrated into SHSs as the channel materials for device applications. Instead, the rapid development of 2D layered FE materials such as CuInP2S6 [40], In2Se3 [41], SnS [42], and MoTe2 [43] brings new possibility for the realization of nonvolatile doping by means of a local electric polarization field. Some 2D ferroelectrics exhibit robust spontaneous polarization down to single-layer limit [42-44]. Especially for 2D α-In2X3 (X = S, Se) ferroelectrics, they are semiconductors with the band gap range of 1.4−2.8 eV [45]. Therefore, the combination of 2D semiconductors (e.g., TMDs and BP) with α-In2X3 ferroelectrics into FE VSHs can induce several aspects of advantages compared with other types of SHSs. (i) The dangling-bond-free surfaces lead to the formation of VSHs with a sharp atomic interface that is conductive to the doping modulation by the local polarization-field. (ii) The switchable FE field can be used to engineer the band alignment of the VSHs, including type-I (straddling gap), type-II (staggered gap), and type-III (broken gap) [46], which is essential for exploiting multifunctional devices. (iii) The ultrathin structure of FE VSHs is adapted to the trend of device minimization. Although the FE polarization field has been confirmed as a feasible way for regulating the carrier density and type in 2D semiconductors, the FE polarization effect on the band alignment, doping polarity, and carrier density in 2D FE VSHs remains unexplored.
In this work, we report a nonvolatile doping strategy in 2D ferroelectric VSHs consisted of TMD monolayers (MoSe2 and MoTe2) and experimentally available α-In2X3 (X = S, Se) nanosheets. The local FE polarization-field effect on the band alignment and carrier density of TMD/In2X3 VSHs is systematically studied by the first-principles calculations. Different from conventional 2D semiconductors, there is a layer-by-layer band shift in α-In2X3 semiconductors induced by the FE polarization field, which creates a favorable condition for the modulation of band alignment and carrier density in TMD/In2X3 VSHs. We demonstrate that free-doping and nonvolatile p−n or p−i−n junctions with ultrahigh carrier concentration (1013−1014 cm−2) can be achieved in MoTe2/In2Se3 VSHs, and their carrier type and spatial distribution can be modulated by the control of FE polarization states. These findings suggest huge potential of 2D MoTe2/In2Se3 VSHs for self-powered multifunctional device applications.

2 Methods

2.1 Computational details

All density functional theory (DFT) calculations were performed in the Vienna ab initio Simulation Package (VASP) code with the projector augmented wave (PAW) method [47]. The electronic exchange-correlation energy was treated by the generalized-gradient approximation (GGA) functional in the form of Perdew−Burke−Ernzerh (PBE) [48]. The vdW interactions were treated with the DFT-D3 correction in Grimme’s scheme [49]. A kinetic energy cutoff was set to 500 eV for the plane-wave expansion set. A vacuum layer of more than 20 Å along the z-axis (or out-of-plane) direction was added to avoid the interactions between adjacent periodic slabs. Since the PBE functional usually underestimates the band gap of semiconductors, electronic structures of all 2D structures were predicted by the hybrid functional of Heyd−Scuseria−Ernzerhof (HSE06) [50]. The dipole correction along the z-axis direction was adopted to meet the convergence criterion. The Monkhorst−Pack scheme was used for the k-point sampling in the Brillouin zone with the grid of 16 × 16 × 1 for 2D materials and 12 × 12 × 1 for 2D VSHs, respectively. All geometry optimization was carried out using the conjugate-gradient method and the convergence criteria of energy and force was set to 10−4 eV/atom and 10−2 eV/Å, respectively. The out-of-plane electric polarization was evaluated by using the Berry phase method [51]. Bader charge analysis was performed to examine the interfacial charge transfers [52].

2.2 Heterostructure models

Prior to constructing In2X3/TMD (X = S, Se) VSHs, the geometrics of 2D freestanding TMD monolayers (i.e., MoSe2 and MoTe2) and α-In2X3 (X = S, Se) were firstly optimized. For hexagonal H-phase TMD monolayers, the optimized in-plane lattice constant a is 3.32 Å and 3.55 Å for the monolayer MoSe2 and MoTe2, respectively. α-In2X3 (X = S, Se) nanosheets also possess hexagonal lattice with the atom sequence of Se−In−Se−In−Se in each layer. The optimized lattice constant a is 3.95 Å and 4.11 Å for α-In2S3 and α-In2Se3, respectively. The increase of layer number from the monolayer to tetralayer brings few changes of the in-plane lattice constants (< 0.1 Å). Based on the optimized 2D freestanding structures, 2D In2X3/TMD VSHs were created by vertically stacking a (2 × 2) TMD monolayer on a (3 × 3) α-In2X3 nanosheet, which makes that the lattice mismatch of all considered VSHs is less than 2.7%. According to the polarization orientation, TMD/In2X3 VSHs were classed into TMD/In2X3(↑) and TMD/In2X3(↓), respectively. The optimal interfacial configuration of each In2X3/TMD VSH was obtained by the energy comparison [see Fig. S1 in the Electronic Supplementary Material (ESM) for details]. The interfacial interactions of the VSHs were determined by calculating interfacial binding energies (γint) as follows:
γint=(EtotETMDEIn2X3)/SA,
where Etot, ETMD, EIn2X3 are the total energies of a given In2X3/TMD VSH, monolayer TMD monolayer, and α-In2X3 nanosheet, respectively, SA is the interfacial area. The optimized lattice parameters and energetic data of MoTe2/In2Se3 and MoSe2/In2S3 VSHs were listed in Tables S1 and S2 of the ESM, respectively.

2.3 Calculation of carrier density

The carrier density of electrons (σn) and holes (σp) in 2D semiconductors and VSHs was determined as follows [46]:
σn=4πmekTh2ln[1+exp(EFECkT)],
σp=4πmhkTh2ln[1+exp(EVEFkT)],
where me* (mh*) is electron (hole) effective mass, T is the temperature, k is Boltzmann constant, h is Plank constant. EFEC and EVEF are the separation from Fermi level to CBM and VBM, respectively, and they can be directly extracted from the calculated band structures. The me* and mh* were calculated from the band curvature at the CBM and VBM, respectively, by using the following formula [46]:
1m=12d2E(k)dk2,
where E(k) is the energy of band-edge states (VBM and CBM) at wave-vector k, and is the reduced Plank constant.

3 Results and discussion

3.1 Polarization effect on electronic properties of 2D In2X3 ferroelectrics

Before addressing the FE polarization effect on the band alignment and carrier density of 2D TMD/In2X3 VSHs, it is necessary to first understand ferroelectric and electronic properties of 2D layered α-In2X3 (X = S, Se) nanosheets. Fig.1(a) shows atomic structures of In2X3 nanosheets in α and β phases. The α-In2X3 nanosheets have a hexagonal lattice with asymmetric distribution of X−In−X−In−X atomic layers in each In2X3 sheet that causes spontaneously aligned electric dipoles along the out-of-plane direction, which is responsible for the formation of FE polarization. According to the polar-axis orientation, the FE polarization can be divided into downward and upward states (Pdown and Pup), respectively. In contrast, the X−In−X−In−X atomic arrangement in β-In2X3 nanosheets is spatially symmetric along the out-of-plane direction [Fig.1(a)], which makes β-In2X3 nanosheets as paraelectric (PE) phase without spontaneous polarization. When α-In2X3 nanosheets are loaded by an external electric field, their polarization orientation can be switched between the Pdown and Pup states via the PE phase. The intrinsic FE polarization leads to a build-in electrostatic field in α-In2X3 along the out-of-plane direction, which can be confirmed by electrostatic potential distributions (EPDs). The electrostatic potential (Φ) of a trilayer (3L) α-In2Se3 nanosheet shows an asymmetric distribution [Fig.1(b)], resulting in an electrostatic potential difference (ΔΦ) of 3.20 eV between two surface terminations and electric dipole moment (μ) of 1.01 Deby. The similar results have been appeared in other 2D α-In2X3 (Fig. S2 in the ESM). For β-In2Se3 nanosheets, their Φ shows completely a symmetric distribution along the out-of-plane direction (Fig. S3 in the ESM) and corresponding ΔΦ and μ values are zero. Hence, ΔΦ and μ values reflect the magnitude of spontaneous polarization in In2X3 ferroelectrics. We plot the μ of eight 2D α-In2X3 as a function of ΔΦ in Fig.1(c). There is a nearly linear relationship between ΔΦ and μ, and it can be fitted by an expression μ = 0.33ΔΦ. So a large ΔΦ means the accumulation of more surface dipole charges, which corresponds to a large polarization field. Hence, the intrinsic polarization of α-In2S3 is correspondingly stronger than that of α-In2Se3. For example, based on Berry-phase method [51], the intrinsic polarization (P) is 1.41 μC/cm2 and 1.09 μC/cm2 for monolayer α-In2S3 and α-In2Se3, respectively. In addition, the increase of layer number leads to the increase of ΔΦ and μ values in 2D α-In2X3 ferroelectrics. Nevertheless, the ΔΦ and μ values are gradually saturated when the layer number is beyond 3L (Fig. S4 in the ESM), which agrees well with previous report about the thickness-dependent polarization in 2D α-In2Se3 [53].
Fig.1 Atomic structures and intrinsic polarization of 2D α-In2X3 ferroelectrics. (a) Top and side views of the atomic structure of trilayer In2X3 nanosheets in FE and PE phases. The arrows ↓ and ↑ denote the ferroelectric α-In2X3 nanosheet in Pdown and Pup states, respectively. The rhombic frames represent the range of a unit cell. The blue and orange balls denote In and X atoms, respectively. (b) The average EPD (Φ) of a trilayer α-In2Se3 along the out-of-plane direction. Here ΔΦ denotes the electrostatic potential difference between two surface terminations. (c) The μ of 2D α-In2X3 ferroelectrics as a function of ΔΦ.

Full size|PPT slide

The local FE polarization-field effect on electronic properties of 2D α-In2X3 is evaluated by their band structures and density of states (DOS) obtained by the HSE06 functional. Fig.2(a) presents the projected band structure and DOS of a trilayer α-In2Se3. It seems that it is a metal due to the lack of band gap. Actually, each In2Se3 layer in the trilayer structure remains semiconducting characteristic with the band gap values of 1.4−1.5 eV, but there is a layer-by-layer band shift [Fig.2(a)]. More specifically, the top-layer bands indicate a notable upward shift relative to the bottom-layer ones when the polarization direction points from the bottom layer to the top layer. The charge-density isosurface distributions of the trilayer α-In2Se3 display that its conduction band minimum (CBM) and valence band maximum (VBM) are dominated by the bottom-layer and top-layer electronic states [Fig.2(b)], respectively. The result confirms the layer-by-layer band shift induced by the polarization field [i.e., so-called quantum-confined Stark (QCS) effect] that drives the separation of electron and hole wave functions. Hence, there is an inclined band structure or band bending in the trilayer α-In2Se3 nanosheet [Fig.2(c)]. Similar results can be also found in other 2D α-In2X3 structures (Fig. S5 in the ESM).
Fig.2 Polarization effect on electronic properties of 2D α-In2X3 nanosheets. (a) Projected band structure and DOS of a trilayer α-In2Se3 nanosheet obtained by the HSE06 functional. The Fermi level is set to the energy zero point. (b) Charge density isosurfaces of CBM and VBM states in the trilayer α-In2Se3 nanosheet. (c) Schematic of the polarization-induced layer-by-layer band shift in a trilayer α-In2X3 nanosheet. (d) Band-edge energies of eight 2D α-In2X3 structures relative to the vacuum level. EC and EV correspond to the energy of CBM and VBM, respectively. The arrows indicate the polarization direction.

Full size|PPT slide

In order to offer a comprehensive understanding of the polarization effect on the band alignment of 2D α-In2X3 ferroelectrics, their band-edge energies (EC and EV) relative to the vacuum level are plotted in Fig.2(d). Different from the conventional 2D semiconductors (e.g., TMDs), all 2D α-In2X3 (X = S, Se) structures possess inclined band alignment along the direction of polar axis. Their band gaps (Eg) show a rapid reduction with increasing layer number due to the cooperation of quantum size effect and QCS effect. For example, the Eg is 1.43, 0.11, and 0 eV for the monolayer (1L), bilayer (2L), and trilayer (3L) α-In2Se3, respectively. Moreover, the VBM energy of top-layer In2X3 is higher than the CBM energy of bottom-layer In2X3 when the thickness is beyond 2L due to the layer-by-layer band shift, consequently leading to the formation of self-doped p−n junctions in 2D α-In2X3 [54]. Moreover, the self-doping characteristics can be stabilized even if the layer thickness is further increased due to the Fermi-level pinning caused by the crossover of band-edge states between the top and bottom layers.

3.2 Band alignments and tunable carrier types in TMD/In2X3 VSHs

Thanks to the polarization-induced band bending (or layer-by-layer band shift) in 2D α-In2X3 (X = S, Se) ferroelectrics, the integration of them with the conventional 2D semiconductors (e.g., MoSe2 and MoTe2) creates the possibility to control their band alignments and carrier distribution via switchable FE polarization. We selected 2D MoTe2/In2Se3 VSHs as an example to demonstrate the polarization-field effect on their electronic structures and doping polarity, and the thickness of α-In2Se3 layer in the VHSs is considered from 1L to 4L. The VSHs were constructed by vertically stacking a 2 × 2 MoTe2 monolayer on 3 × 3 α-In2Se3 nanosheets for the minimization of lattice mismatch. Based on the polar-axis orientation (Pdown and Pup), the VSHs were classified into MoTe2/In2Se3(↓) and MoTe2/In2Se3(↑) (Fig.3). The interfacial structure of each VSH was determined by the energy comparison among six potential interfacial configurations (see Fig. S1 in the ESM for details), and the most stable one was selected for the study of electronic properties. The optimized lattice and energetic parameters of the lowest-energy VSHs in Pdown and Pup states are listed in Tables S1 and S2. All MoTe2/In2Se3 VSHs have negative interfacial binding energies (γint), indicating that the formation of vdW heterojunctions is energetically favorable. Moreover, the γint values of these VSHs (−0.32− −0.45 J/m2) are comparable with other 2D vdW crystals (−0.2− −0.5 J/m2) [55], indicating the heterointerfaces are interacted by the vdW force. By comparison, the γint of MoTe2/In2Se3(↓) VSHs is correspondingly larger than that of MoTe2/In2Se3(↑) ones, resulting in relatively small interlayer distances (d = 3.39−3.44 Å) in MoTe2/In2Se3(↓) VSHs due to stronger interfacial interactions. The difference of interfacial interactions between the two types of VSHs plays an important role in the determinization of their band alignments, which will be discussed later.
Fig.3 Top and side views of the atomic structure of 2D MoTe2/In2Se3 VSHs in Pdown and Pup states. The solid frames denote the supercells of α-In2Se3 and MoTe2 used for the construction of VSHs, and the dash frames denote the primitive cell of α-In2Se3 and MoTe2. T-In2Se3 and B-In2Se3 denote the top-layer and bottom-layer In2Se3, respectively.

Full size|PPT slide

We now turn to reveal the local polarization-field effect on the band alignment of 2D MoTe2/In2Se3 VSHs. Fig.4 shows the projected band structures and DOS of MoTe2/In2Se3(↑) VSHs calculated by the HSE06 functional. For MoTe2/1L-In2Se3(↑) VSH [Fig.4(a)], it exhibits a typical type-II band alignment in which the CBM and VBM states are contributed by the α-In2Se3 and MoTe2 layers, respectively. The thickness increase in In2Se3 layer brings two notable changes in band structures [Fig.4(a)−(c)]: (i) the band gap of In2Se3 layer rapidly decrease and disappear due to the size and QCS effects, (ii) the MoTe2-layer bands show a downward shift towards the low-energy direction. It seems that the increase of In2Se3 layer induces the band-alignment transition from the type-II to type-I, to metallic, but the results of projected DOS indicate that the interfacial band alignment of all the VSHs possesses the type-II characteristic. For example, for MoTe2/2L-In2Se3(↑) VSH, the band-edge states of the top-layer In2Se3 (T-In2Se3) and MoTe2 layer exhibit a staggered distribution, indicating the type-II band alignment [Fig.4(b)]. Similar result has also been appeared in other VSHs in Pup state. The band bending of In2Se3 layer induced by the polarization field makes the electron accumulation at T-In2Se3 near the interface of VSHs, especially for the light irradiation condition [Fig.4(d)]. Owing to the existence of build-in electric field, the thickness increase of In2Se3 layer will drive the separation of electrons and holes onto T-In2Se3 and B-In2Se3, respectively. Moreover, the carrier density of the VSHs can be enhanced by the thickness increase of In2Se3 layer, resulting in the formation of self-doped p−n junctions in the In2Se3 layer. Nevertheless, the distribution of band-edge states in MoTe2/In2Se3(↑) VSHs has few changes when the thickness of In2Se3 layer is beyond 3L (Fig. S6 in the ESM), which arises from the crossover of band-edge states. Therefore, the band alignment and doping polarity of the VSHs will be similar to that of MoTe2/3L-In2Se3(↑) VSH [Fig.4(c)] even if the thickness of In2Se3 layer is further increased.
Fig.4 Band alignments of 2D MoTe2/In2Se3(↑) VSHs with the thickness range of In2Se3 layer from 1L to 3L. (a−c) Projected band structure and DOS of VSH with (a) 1L-In2Se3, (b) 2L-In2Se3, and (c) 3L-In2Se3. The red and blue bands denote the contribution from In2Se3 and MoTe2 layers, respectively. The horizontal dash lines denote the position of Fermi level. T-In2Se3, M-In2Se3, and B-In2Se3 represent the top-layer, middle-layer, and bottom-layer In2Se3 in the VSHs, respectively. (d) Schematic of band alignments of MoTe2/In2Se3(↑) VSHs. The black and gray arrows denote the carrier transport direction and polarization direction, respectively.

Full size|PPT slide

Fig.5 displays the project band structures and DOS of MoTe2/In2Se3(↓) VSHs. Contrary to the case of MoTe2/In2Se3(↑) VSHs, the CB- and VB-edge states of MoTe2/In2Se3(↓) VSHs are contributed by the B-In2Se3 and MoTe2 layers, respectively. The CBM positions of all these VSHs are lower than the Fermi level and their VBM positions are higher than the Fermi level, leading to the formation of p-MoTe2/n-In2Se3 vdW junctions. Interestingly, all these VSHs have type-II interface band alignment based on the result of projected DOS. Different from the conventional type-II SHSs, the electrons and holes can be rapidly separated into two surface terminations of the VSHs due to the effect of FE polarization field. Hence, the VSHs behave as self-doped p−i−n junction along the polarization direction [Fig.5 (d)]. The tunable band alignments and self-doping characteristic have been also found in 2D MoSe2/In2S3 VSHs (see Figs. S7 and S8 in the ESM), suggesting that the FE polarization field can be regarded as an efficient way to tune the carrier type and density in 2D ferroelectric VSHs.
Fig.5 Band alignments of 2D MoTe2/In2Se3(↓) VSHs with the thickness range of In2Se3 layer from 1L to 3L. (a−c) Projected band structure and DOS of VSH with (a) 1L-In2Se3, (b) 2L-In2Se3, and (c) 3L-In2Se3. The red and blue bands represent the contribution of In2Se3 and MoTe2 layers, respectively. The horizontal dash lines denote the position of Fermi level. T-In2Se3, M-In2Se3, and B-In2Se3 represent the top-layer, middle-layer, and bottom-layer In2Se3 in the VSHs, respectively. (d) Schematic of band alignments of MoTe2/In2Se3(↓) VSHs. The arrows indicate the polarization direction.

Full size|PPT slide

In order to reveal the effect of FE polarization on the band alignment and doping polarity of MoTe2/In2Se3 VSHs, their EPDs and interface charge transfers have been investigated. Fig.6(a) indicates the EPDs of MoTe2/3L-In2Se3 VSH in two polarization states. The electrostatic potentials (Φ) at two surface terminations of the VSHs are different due to the existence of a polarization field. The calculated ΔΦ of MoTe2/3L-In2Se3(↑) VSH (3.26 eV) largely approaches that of freestanding 3L-In2Se3 (3.20 eV) but is far large than that of MoTe2/3L-In2Se3(↑) VSH (1.61 eV). Similar results can be also found in other VSHs [Fig.6(b)]. The reduction of ΔΦ implies that the polarized charges of In2Se3 layer have been partially counteracted by interfacial charge transfers, weakening the FE polarization field [56]. To examine the interfacial charge transfers, the differential charge densities of MoTe2/3L-In2Se3 VSHs in two polarization states have been calculated [Fig.6(c)]. The red and green charge-density isosurfaces denote the charge accumulation and depletion, respectively. It can be clearly identified the charge localization at T-In2Se3 and the charge depletion at the MoTe2 layer for the VSH in Pdown state, corresponding to the electron transfer (~0.27 e) from the MoTe2 layer to T-In2Se3 based on the analysis of Bader charges. In contrast, the charge accumulation and depletion at the interface of MoTe2/3L-In2Se3 in Pup state are relatively small. Bader analysis shows few charge transfers (~0.1 e) from the T-In2Se3 layer to the MoTe2 layer. In addition to the charge transfers, the interfacial electron tunneling is another important factor for the change of ΔΦ in the VSHs. For the VSHs in Pup state, they present the type-II band alignment [Fig.4(d)]. Hence, the valence electrons of the MoTe2 layer cannot spontaneously transfer into the In2Se3 layer due to the interfacial energy barriers. This is the reason why the layer-dependent ΔΦ of MoTe2/In2Se3(↑) VSHs is very similar to that of the isolated α-In2Se3. In contrast, for the VSHs in Pdown state, the CBM position of the bottom-layer In2Se3 is lower than the VBM position of the MoTe2 layer when the layer number of In2Se3 is beyond 1L [Fig.5(d)], resulting in a type-III band alignment between the bottom-layer In2Se3 and the MoTe2 layer. Hence, the electrons can be tunneled from the MoTe2 layer into the In2Se3 layer, which weakens the FE polarization field. Therefore, the ΔΦ values of MoTe2/In2Se3(↓) VSHs are far smaller than those of the isolated α-In2Se3. The similar results have been also found in MoSe2/In2S3 VSHs.
Fig.6 Average electrostatic potential (Φ) and interfacial charge transfer of MoTe2/In2Se3 VSHs. (a) Electrostatic potential distribution of MoTe2/3L-In2Se3 VSH in Pup (red line) and Pdown (blue line) states. (b) Electrostatic potential differences (ΔΦ) of MoTe2/In2Se3 VSHs and isolated In2Se3 nanosheets as a function of In2Se3 layer thickness. (c) The charge-density difference and interfacial charge transfer (Q) of MoTe2/3L-In2Se3 VSH in Pdown (left) and Pup (right) state.

Full size|PPT slide

3.3 Polarization-tuned carrier density and spatial distribution in TMD/In2X3 VSHs

To qualitatively describe the polarization-induced self-doping effect in TMD/In2X3 VSHs, the carrier density of MoTe2/In2Se3 VSHs in two polarization states have been calculated. The density of electrons (σn) and holes (σp) of the VSHs in Pup and Pdown states as a function of temperature are plotted in Fig.7(a) and (b), respectively. The intrinsic carrier density of MoTe2 monolayer has been also provided for the comparison. For MoTe2/In2Se3(↑) VSHs [Fig.7(a)], the σn and σp are sensitive to the variation of In2Se3 layer. For instance, σn increases from ~10−2 to 4.3 × 1013 cm−2 and σp increases from ~106 to 1.2 × 1013 cm−2 with the thickness increase of In2Se3 layer from 1L to 3L at the room temperature (298 K). Here it needs to be mentioned that the further increase of In2Se3 layer (e.g., >3L) does not significantly change the σn and σp values due to the Fermi-level pinning induced by the crossover of band-edge states mentioned above. In contrast, the carrier density of MoTe2/In2Se3(↓) VSHs displays a relatively smaller change with the variation of In2Se3 layer. Moreover, the carrier density of the VSHs is not very sensitive to the temperature change [Fig.7(b)]. For example, σn largely keeps at 2.1 × 1012, 4.2 × 1013, and 6.7 × 1013 cm−2 and σn basically maintains at 1.6 × 1011, 4.8 × 1013, and 8.3 × 1013 cm−2 for the VSH with 1L-, 2L-, and 3L-In2Se3, respectively. The results originate from the formation of tunneling p−n junctions [Fig.5(d)] where the carrier transport is dominated by the tunneling mechanism [57]. The similar result has been demonstrated in the tunneling BP/ReS2 p−n heterojunctions [58]. Moreover, the carrier density of MoTe2/In2Se3 VSHs is far larger than that of the pristine MoTe2 (< 106 cm−2). Importantly, the doping level in TMD/In2X3 VSHs induced by the local FE polarization exceeds most of electrically and chemically modified TMD junctions, including lateral MoTe2 homojunction [(3−5) × 1012 cm−2] [59], MoTe2/MoS2 VSHs (1011−1012 cm−2) [60], lateral WSe2/MoSe2 heterojunctions (1011 cm−2) [61], and WSe2/MoS2 VSHs (~1012 cm−2) [62].
Fig.7 Polarization-dependent carrier density in 2D MoTe2/In2Se3 VSHs. The σn (solid lines) and σp (dash lines) of MoTe2/In2Se3 VSHs in (a) Pup state and (b) Pdown state. The VSH-nL (n = 1−3) denote the MoTe2/In2Se3 VSH with nL-In2Se3 layer. The spatial distribution of (c) σn and (d) σp in MoTe2/3L-In2Se3 VSH along the FE switching pathway from the Pup state, PE phase, to Pdown state.

Full size|PPT slide

In addition to the promotion of carrier density, the polarization switching can be also used to manipulate the carrier spatial distribution of TMD/In2X3 VSHs. For example, the electrons and holes of MoTe2/In2Se3(↑) tend to be localized at T-In2Se3 and B-In2Se3 layers, respectively. While the polarization switching into Pdown state leads to the carrier separation into the MoTe2 and B-In2Se3 layers. In order to better insight into the evolution of carrier spatial distribution in TMD/In2X3 VSHs during the polarization switching, we used MoTe2/3L-In2Se3 VSH as an example and plotted its spatial distribution of carrier density isosurfaces along the phase transition pathway from the Pup state to the PE state to the Pdown state, as shown in Fig.7(c) and (d). The corresponding band structure evolution of MoTe2 and In2Se3 layers in the VSH associated to the sequential phase switching has been presented in Fig. S9. When the polarization state transforms from the Pup state to the PE state, the FE polarization of In2Se3 layer will be gradually reduced to zero. The layer-by-layer band shift in the In2Se3 layer is weakened and the band-edge-states of MoTe2 layer indicate an upward shift in this process (Fig. S9 in the ESM). Hence, we observe the electron delocalization from the T-In2Se3 layer to the whole In2Se3 layer [Fig.7(c)] and the hole localization from the B-In2Se3 layer to the MoTe2 layer [Fig.7(d)], respectively. The decreased FE polarization leads to the reduction of σn and σn in this process due to the carrier delocalization and band shift. When the polarization state transforms from the PE state to the Pdown state, the FE polarization of In2Se3 layer will be gradually enhanced. The band-edge states of In2Se3 layer are changed from degenerate into dispersed with a layer-by-layer shift (Fig. S9 in the ESM), promoting the electron localization in the B-In2Se3 layer. The enhanced polarization and band bending are responsible for the increase of σn and σn in the process [Fig.7(c) and (d)]. The results suggest that the carrier density and spatial distribution of TMD/In2X3 VSHs can be simultaneously modulated by the FE polarization switching.

4 Conclusions

In summary, we have theoretically demonstrated the feasibility to modulate the band alignment and spatial carrier density of 2D TMD/In2X3 VSHs via switchable FE polarization. Our calculated results indicate that the layer-by-layer band shift of In2X3 layer driven by the FE polarization field induces TMD/In2X3 VSHs with tunable band alignments and doping polarities, which enables us to manipulate the carrier type and density in the VSHs by the control of polarization orientation and magnitude. Based on this strategy, nonvolatile vdW p−n junctions have been achieved in MoTe2/In2Se3 VSHs with ultrahigh carrier density (1013−1014 cm−2). Moreover, the switching of FE polarization from Pup to Pdown state can lead to the transition of the VSHs from n−p junctions to p−i−n junctions. This work provides a new doping strategy for modulating the carrier type and density of 2D VSHs without any structural damages, paving the way for exploiting versatile 2D electronics and optoelectronics.

References

[1]
J. Simon , V. Protasenko , C. Lian , H. Xing , D. Jena . Polarization-induced hole doping in wide-band-gap uniaxial semiconductor heterostructures. Science, 2010, 327(5961): 60
CrossRef ADS Google scholar
[2]
J. D. Sau , R. M. Lutchyn , S. Tewari , S. Das Sarma . Generic new platform for topological quantum computation using semiconductor heterostructures. Phys. Rev. Lett., 2010, 104(4): 040502
CrossRef ADS Google scholar
[3]
C. Siegert , A. Ghosh , M. Pepper , I. Farrer , D. A. Ritchie . The possibility of an intrinsic spin lattice in high-mobility semiconductor heterostructures. Nat. Phys., 2007, 3(5): 315
CrossRef ADS Google scholar
[4]
J. Narayan , S. Oktyabrsky . Formation of misfit dislocations in thin film heterostructures. J. Appl. Phys., 2002, 92(12): 7122
CrossRef ADS Google scholar
[5]
X. Liu , D. Cao , Y. Yao , P. Tang , M. Zhang , X. Chen , H. Shu . Heteroepitaxial growth and interface band alignment in a large-mismatch CsPbI3/GaN heterojunction. J. Mater. Chem. C, 2022, 10(6): 1984
CrossRef ADS Google scholar
[6]
R. Yang , J. Fan , M. Sun . Transition metal dichalcogenides (TMDCs) heterostructures: Optoelectric properties. Front. Phys., 2022, 17(4): 43202
CrossRef ADS Google scholar
[7]
K. Mak , J. Shan . Photonics and Optoelectronics of 2D semiconductor transition metal dichalcogenides. Nat. Photonics, 2016, 10(4): 216
CrossRef ADS Google scholar
[8]
H. Liu , Y. Du , Y. Deng , P. D. Ye . Semiconducting black phosphorus: synthesis, transport properties and electronic applications. Chem. Soc. Rev., 2015, 44(9): 2732
CrossRef ADS Google scholar
[9]
S. Zhang , S. Guo , Z. Chen , Y. Wang , H. Gao , J. Gómez-Herrero , P. Ares , F. Zamora , Z. Zhu , H. Zeng . Recent progress in 2D group-VA semiconductors: from theory to experiment. Chem. Soc. Rev., 2018, 47(3): 982
CrossRef ADS Google scholar
[10]
Y. Liu , N. O. Weiss , X. Duan , H. C. Cheng , Y. Huang , X. Duan . Van der Waals heterostructures and devices. Nat. Rev. Mater., 2016, 1(9): 16042
CrossRef ADS Google scholar
[11]
Y. Y. Wang , F. P. Li , W. Wei , B. B. Huang , Y. Dai . Interlayer coupling effect in van der Waals heterostructures of transition metal dichalcogenides. Front. Phys., 2021, 16(1): 13501
CrossRef ADS Google scholar
[12]
L. Zhang , Z. Zhang , F. Wu , D. Wang , R. Gogna , S. Hou , K. Watanabe , K. Taniguchi , K. Kulkarni , T. Kuo , S. R. Forrest , H. Deng . Twist-angle dependence of moiré excitons in WS2/MoSe2 heterobilayers. Nat. Commun., 2020, 11(1): 5888
CrossRef ADS Google scholar
[13]
M. R. Rosenberger , H. J. Chuang , M. Phillips , V. P. Oleshko , K. M. McCreary , S. V. Sivaram , C. S. Hellberg , B. T. Jonker . Twist angle-dependent atomic reconstruction and moiré patterns in transition metal dichalcogenide heterostructures. ACS Nano, 2020, 14(4): 4550
CrossRef ADS Google scholar
[14]
H. Chen , X. Wen , J. Zhang , T. Wu , Y. Gong , X. Zhang , J. Yuan , C. Yi , J. Lou , P. M. Ajayan , W. Zhuang , G. Zhang , J. Zheng . Ultrafast formation of interlayer hot excitons in atomically thin MoS2/WS2 heterostructures. Nat. Commun., 2016, 7(1): 12512
CrossRef ADS Google scholar
[15]
A. F. Rigosi , H. M. Hill , Y. Li , A. Chernikov , T. F. Heinz . Probing interlayer interactions in transition metal dichalcogenide heterostructures by optical spectroscopy: MoS2/WS2 and MoSe2/WSe2. Nano Lett., 2015, 15(8): 5033
CrossRef ADS Google scholar
[16]
J. Guo , L. Wang , Y. Yu , P. Wang , Y. Huang , X. Duan . SnSe/MoS2 van der Waals heterostructure junction field-effect transistors with nearly ideal subthreshold slope. Adv. Mater., 2019, 31(49): 1902962
CrossRef ADS Google scholar
[17]
Y. Cheng , P. Tang , P. Liang , X. Liu , D. Cao , X. Chen , H. Shu . Sulfur-driven transition from vertical to lateral growth of 2D SnS−SnS2 heterostructures and their band alignments. J. Phys. Chem. C, 2020, 124(50): 27820
CrossRef ADS Google scholar
[18]
J. Xu , J. Jia , S. Lai , J. Ju , S. Lee . Tunneling field effect transistor integrated with black phosphorus-MoS2 junction and ion gel dielectric. Appl. Phys. Lett., 2017, 110(3): 033103
CrossRef ADS Google scholar
[19]
S. J. Liang , B. Cheng , X. Cui , F. Miao . Van der Waals heterostructures for high-performance device applications: Challenges and opportunities. Adv. Mater., 2020, 32: 1903800
[20]
R. Cheng , F. Wang , L. Yin , Z. Wang , Y. Wen , T. A. Shifa , J. He . High-performance, multifunctional devices based on asymmetric van der Waals heterostructures. Nat. Electron., 2018, 1(6): 356
CrossRef ADS Google scholar
[21]
H. P. Komsa , J. Kotakoski , S. Kurasch , O. Lehtinen , U. Kaiser , A. V. Krasheninnikov . Two-dimensional transition metal dichalcogenides under electron irradiation: Defect production and doping. Phys. Rev. Lett., 2012, 109(3): 035503
CrossRef ADS Google scholar
[22]
Q. Zhang , H. Ying , X. Li , R. Xiang , Y. Zheng , H. Wang , J. Su , M. Xu , X. Zheng , S. Maruyama , X. Zhang . Controlled doping engineering in 2D MoS2 crystals toward performance augmentation of optoelectronic devices. ACS Appl. Mater. Interfaces, 2021, 13(27): 31861
CrossRef ADS Google scholar
[23]
Y. Gong , H. Yuan , C. L. Wu , P. Tang , S. Z. Yang , A. Yang , G. Li , B. Liu , J. van de Groep , M. L. Brongersma , M. F. Chisholm , S. C. Zhang , W. Zhou , Y. Cui . Spatial controlled doping of two-dimensional SnS2 through intercalation for electronics. Nat. Nanotechnol., 2018, 13(4): 294
CrossRef ADS Google scholar
[24]
D. Kiriya , M. Tosun , P. Zhao , J. S. Kang , A. Javey . Air-stable surface charge transfer doping of MoS2 by benzyl viologen. J. Am. Chem. Soc., 2014, 136(22): 7853
CrossRef ADS Google scholar
[25]
W. Shi , S. Kahn , L. Jiang , S. Y. Wang , H. Z. Tsai , D. Wong , T. Taniguchi , K. Watanabe , F. Wang , M. F. Crommie , A. Zettl . Reversible writing of high mobility and high-carrier density doping patterns in two-dimensional van der Waals heterostructures. Nat. Electron., 2020, 3(2): 99
CrossRef ADS Google scholar
[26]
R. Zhang , Z. Xie , C. An , S. Fan , Q. Zhang , S. Wu , L. Xu , X. Hu , D. Zhang , D. Sun , J. Chen , J. Liu . Ultraviolet light-induced persistent and degenerated doping in MoS2 for potential photocontrollable electronics applications. ACS Appl. Mater. Interfaces, 2018, 10(33): 27840
CrossRef ADS Google scholar
[27]
M. Buscema , D. J. Groenendijk , G. A. Steele , H. S. J. van der Zant , A. Castellanos-Gomez . Photovoltaic effect in few-layer phosphorus PN junctions defined local electrostatic gating. Nat. Commun., 2014, 5(1): 4651
CrossRef ADS Google scholar
[28]
P. Agnihotri , P. Dhakras , J. U. Lee . Bipolar junction transistors in two-dimensional WSe2 with large current and photocurrent grains. Nano Lett., 2016, 16(7): 4355
CrossRef ADS Google scholar
[29]
S. J. Lee , Z. Lin , X. Duan , Y. Huang . Doping on demand in 2D devices. Nat. Electron., 2020, 3(2): 77
CrossRef ADS Google scholar
[30]
L. Kong , X. Zhang , Q. Tao , M. Zhang , W. Dang , Z. Li , L. Feng , L. Liao , X. Duan , Y. Liu . Doping-free complementary WSe2 circuit via van der Waals metal integration. Nat. Commun., 2020, 11(1): 1866
CrossRef ADS Google scholar
[31]
D. Wijethunge , L. Zhang , C. Tang , A. Du . Tunning band alignment and optical properites of 2D van der Waals heterostructure via ferroelectric polarization switching. Front. Phys., 2020, 15(6): 63504
CrossRef ADS Google scholar
[32]
J.W. ChenS.T. LoS.C. HoS.S. WongT.H. Y. VuX.Q. ZhangY.D. LiuY.Y. ChiouY.X. ChenJ.C. YangY.C. ChenY.H. ChuY.H. LeeC.J. ChungT.M. ChenC.H. ChenC.L. Wu, A gate-free monolayer WSe2 PN diode, Nat. Commun. 9(1), 3143 (2018)
[33]
Z. Lu , C. Serrao , A. I. Khan , L. You , J. C. Wong , Y. Ye , H. Zhu , X. Zhang , S. Salahuddin . Nonvolatile MoS2 field effect transistors directly gated by single crystalline epitaxial ferroelectric. Appl. Phys. Lett., 2017, 111(2): 023104
CrossRef ADS Google scholar
[34]
A. Nguyen , P. Sharma , T. Scott , E. Preciado , V. Klee , D. Sun , I. H. D. Lu , D. Barroso , S. H. Kim , V. Y. Shur , A. R. Akhmatkhanov , A. Gruverman , L. Bartels , P. A. Dowben . Toward ferroelectric control of monolayer MoS2. Nano Lett., 2015, 15(5): 3364
CrossRef ADS Google scholar
[35]
X. Liu , X. Zhou , Y. Pan , J. Yang , H. Xiang , Y. Yuan , S. Liu , H. Luo , D. Zhang , J. Sun . Charge–ferroelectric transition in ultrathin Na0.5Bi4.5Ti4O15 flakes probed via a dual-gated full van der Waals transistor. Adv. Mater., 2020, 32(49): 2004813
CrossRef ADS Google scholar
[36]
G. Wu , X. Wang , Y. Chen , S. Wu , B. Wu , Y. Jiang , S. Shen , T. Lin , Q. Liu , X. Wang , P. Zhou , S. Zhang , W. Hu , X. Meng , J. Chu , J. Wang . MoTe2 p–n homojunctions defined by ferroelectric polarization. Adv. Mater., 2020, 32(16): 1907937
CrossRef ADS Google scholar
[37]
G. Wu , B. Tian , L. Liu , W. Lv , S. Wu , X. Wang , Y. Chen , J. Li , Z. Wang , S. Wu , H. Shen , T. Lin , P. Zhou , Q. Liu , C. Duan , S. Zhang , X. Meng , S. Wu , W. Hu , X. Wang , J. Chu , J. Wang . Programmable transition metal dichalcogenide homojunctions controlled by nonvolatile ferroelectric domains. Nat. Electron., 2020, 3(1): 43
CrossRef ADS Google scholar
[38]
N. A. Spaldin . Fundamental size limits in ferroelectricity. Science, 2004, 304(5677): 1606
CrossRef ADS Google scholar
[39]
M. Dawber , K. M. Rabe , J. F. Scott . Physics of thin-film ferroelectric oxides. Rev. Mod. Phys., 2005, 77(4): 1083
CrossRef ADS Google scholar
[40]
A. Belianinov , Q. He , A. Dziaugys , P. Maksymovych , E. Eliseev , A. Borisevich , A. Morozovska , J. Banys , Y. Vysochanskii , S. V. Kalinin . CuInP2S6 room temperature layered ferroelectric. Nano Lett., 2015, 15(6): 3808
CrossRef ADS Google scholar
[41]
W. Ding , J. Zhu , J. Wang , Y. Gao , D. Xiao , Y. Gu , Z. Zhang , W. Zhu . Prediction of intrinsic two-dimensional frroelectrics in In2Se3 and other III2−VI3 van der Waals materials. Nat. Commun., 2017, 8(1): 14956
CrossRef ADS Google scholar
[42]
N. Higashitarumizu , H. Kawamoto , C. J. Lee , B. H. Lin , F. H. Chu , I. Yonemori , T. Nishimura , K. Wakabayashi , W. Chang , K. Nagashio . Purely in-plane ferroelectricity in monolayer SnS at room temperature. Nat. Commun., 2020, 11(1): 2428
CrossRef ADS Google scholar
[43]
S. Yuan , X. Luo , H. L. Chan , C. Xiao , Y. Dai , M. Xie , J. Hao . Room-temperature ferroelectricity in MoTe2 down to the atomic monolayer limit. Nat. Commun., 2019, 10(1): 1775
CrossRef ADS Google scholar
[44]
F. Xue , W. Hu , K. C. Lee , L. S. Lu , J. Zhang , H. L. Tang , A. Han , W. T. Hsu , S. Tu , W. H. Chang , C. H. Lien , J. H. He , Z. Zhang , L. J. Li , X. Zhang . Room-temperature ferroelectricity in hexagonally layered α-In2Se3 nanoflakes down to the monolayer limit. Adv. Funct. Mater., 2018, 28(50): 1803738
CrossRef ADS Google scholar
[45]
J. Quereda , R. Biele , G. Rubio-Bollinger , N. Agrait , R. D’Agosta , A. Castellanos-Gomez . Strong quantum confinement effect in the optical properties of ultrathinα-In2Se3. Adv. Opt. Mater., 2016, 4(12): 1939
CrossRef ADS Google scholar
[46]
M. Yang , H. Shu , Y. Li , D. Cao , X. Chen . Polarization-induced band alignment transition and nonvolatile p−n junctions in 2D van der Waals heterostructures. Adv. Electron. Mater., 2022, 8(3): 2101022
CrossRef ADS Google scholar
[47]
G. Kresse , J. Furthmüller . Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B, 1996, 54(16): 11169
CrossRef ADS Google scholar
[48]
M. C. Payne , M. P. Teter , D. C. Allan , T. A. Arias , J. D. Joannopoulos . Iterative minimization techniques for ab initio total-energy calculations: Molecular dynamics and conjugate gradients. Rev. Mod. Phys., 1992, 64(4): 1045
CrossRef ADS Google scholar
[49]
S. Grimme . Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem., 2006, 27(15): 1787
CrossRef ADS Google scholar
[50]
J. Heyd , G. E. Scuseria , M. Ernzerhof . Hybrid functionals based on a screened coulomb potential. J. Chem. Phys., 2003, 118(18): 8207
CrossRef ADS Google scholar
[51]
R. D. King-Smith , D. Vanderbilt . Theory of polarization of crystalline solids. Phys. Rev. B, 1993, 47(3): 1651
CrossRef ADS Google scholar
[52]
R.F. Bader, A quantum theory of molecular structure and its applications, Chem. Rev. 91(5), 893 (1991)
[53]
W. F. Io , S. Yuan , S. Y. Pang , L. W. Wong , J. Zhao , J. Hao . Temperature- and thickness-dependence of robust out-of-plane ferroelectricity in CVD grown ultrathin van der Waals α-In2Se3 layers. Nano Res., 2020, 13(7): 1897
CrossRef ADS Google scholar
[54]
R. Peng , Y. Ma , S. Zhang , B. Huang , L. Kou , Y. Dai . Self-doped p–n junctions in two-dimensional In2X3 van der Waals materials. Mater. Horiz., 2020, 7(2): 504
CrossRef ADS Google scholar
[55]
T. Björkman , A. Gulans , A. V. Krasheninnikov , R. M. Nieminen . Van der Waals bonding in layered compounds from advanced density-functional first-principles calculations. Phys. Rev. Lett., 2012, 108(23): 235502
CrossRef ADS Google scholar
[56]
M. Yang , H. Shu , P. Tang , P. Liang , D. Cao , X. Chen . Intrinsic polarization-induced enhanced ferromagnetism and self-doped p–n junctions in CrBr3/GaN van der Waals heterostructures. ACS Appl. Mater. Interfaces, 2021, 13(7): 8764
CrossRef ADS Google scholar
[57]
P. J. Jeon , Y. T. Lee , J. Y. Lim , J. S. Kim , D. K. Hwang , S. Im . Black phosphorus−zinc oxide nanomaterial heterojunction for p−n diode and junction field-effect transistor. Nano Lett., 2016, 16(2): 1293
CrossRef ADS Google scholar
[58]
P. K. Srivastava , Y. Hassan , Y. Gebredingle , J. Jung , B. Kang , W. J. Yoo , B. Singh , C. Lee . Van der waals broken-gap p−n heterojunction tunnel diode based on black Phosphorus and rhenium disulfide. ACS Appl. Mater. Interfaces, 2019, 11(8): 8266
CrossRef ADS Google scholar
[59]
D. Qu , X. Liu , M. Huang , C. Lee , F. Ahmed , H. Kim , R. S. Ruoff , J. Hone , W. J. Yoo . Carrier-type modulation and mobility improvement of thin MoTe2. Adv. Mater., 2017, 29(39): 1606433
CrossRef ADS Google scholar
[60]
Y. Xie , E. Wu , S. Fan , G. Geng , X. Hu , L. Xu , S. Wu , J. Liu , D. Zhang . Modulation of MoTe2/MoS2 van der Waals heterojunctions for multifunctional devices using N2O plasma with an opposite doping effect. Nanoscale, 2021, 13(16): 7851
CrossRef ADS Google scholar
[61]
J. E. Kim , W. T. Kang , V. Tu Vu , Y. R. Kim , Y. S. Shin , I. Lee , U. Y. Won , B. H. Lee , K. Kim , T. L. Phan , Y. H. Lee , W. J. Yu . Ideal PN photodiode using doping controlled WSe2−MoSe2 lateral heterostructure. J. Mater. Chem. C, 2021, 9(10): 3504
CrossRef ADS Google scholar
[62]
C. H. Lee , G. H. Lee , A. M. van der Zande , W. Chen , Y. Li , M. Han , X. Cui , G. Arefe , C. Nuckolls , T. F. Heinz , J. Guo , J. Hone , P. Kim . Atomically thin p–n junctions with van der Waals heterointerfaces. Nat. Nanotechnol., 2014, 9(9): 676
CrossRef ADS Google scholar

Electronic supplementary materials

Supplementary materials (interfacial configurations and stabilities of 2D TMD/In2X3 VSHs, electrostatic potentials and band structures of 2D α-In2X3 nanosheets, thickness effect on projected band structures of MoTe2/In2Se3 VSHs, band alignments of MoSe2/In2S3 VSHs, and the band evoltion of In2Se3 and MoTe2 layer in MoTe2/In2Se3 VSHs associated to the ferroelectric phase transition) are available in the online version of this article at http://doi.org/10.1007/s11467-022-1244-4 and https://journal.hep.com.cn/fop/EN/10.1007/s11467-022-1244-4.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 62174151 and 61775201) and the Fund of Zhejiang Provincial Natural Science Foundation of China (Grant No. LZ22F040003 and LY22A040002). Computational resources from the Shanghai Supercomputer Center are acknowledged.

RIGHTS & PERMISSIONS

2023 Higher Education Press
AI Summary AI Mindmap
PDF(8013 KB)

Supplementary files

fop-21244-OF-shuhaibo_suppl_1 (1842 KB)

819

Accesses

7

Citations

Detail

Sections
Recommended

/