Two dimensional GeO2/MoSi2N4 van der Waals heterostructures with robust type-II band alignment

Xueping Li, Peize Yuan, Lin Li, Ting Liu, Chenhai Shen, Yurong Jiang, Xiaohui Song, Congxin Xia

Front. Phys. ›› 2023, Vol. 18 ›› Issue (1) : 13305.

PDF(3766 KB)
Front. Phys. All Journals
PDF(3766 KB)
Front. Phys. ›› 2023, Vol. 18 ›› Issue (1) : 13305. DOI: 10.1007/s11467-022-1216-8
RESEARCH ARTICLE
RESEARCH ARTICLE

Two dimensional GeO2/MoSi2N4 van der Waals heterostructures with robust type-II band alignment

Author information +
History +

Abstract

Constructing two-dimensional (2D) van der Waals heterostructures (vdWHs) can expand the electronic and optoelectronic applications of 2D semiconductors. However, the work on the 2D vdWHs with robust band alignment is still scarce. Here, we employ a global structure search approach to construct the vdWHs with monolayer MoSi2N4 and wide-bandgap GeO2. The studies show that the GeO2/MoSi2N4 vdWHs have the characteristics of direct structures with the band gap of 0.946 eV and type-II band alignment with GeO2 and MoSi2N4 layers as the conduction band minimum (CBM) and valence band maximum (VBM), respectively. Also, the direct-to-indirect band gap transition can be achieved by applying biaxial strain. In particular, the 2D GeO2/MoSi2N4 vdWHs show a robust type-II band alignment under the effects of biaxial strain, interlayer distance and external electric field. The results provide a route to realize the robust type-II band alignment vdWHs, which is helpful for the implementation of optoelectronic nanodevices with stable characteristics.

Graphical abstract

Keywords

van der Waals heterostructures / wide gap material / global structure search / robust type-II band alignment

Cite this article

Download citation ▾
Xueping Li, Peize Yuan, Lin Li, Ting Liu, Chenhai Shen, Yurong Jiang, Xiaohui Song, Congxin Xia. Two dimensional GeO2/MoSi2N4 van der Waals heterostructures with robust type-II band alignment. Front. Phys., 2023, 18(1): 13305 https://doi.org/10.1007/s11467-022-1216-8

1 Introduction

Two-dimensional (2D) semiconductor and related van der Waals heterostructures (vdWHs) have attracted tremendous attention because they possess excellent characteristics [1-5]. Also, the 2D vdWHs can form different band alignments, such as type-I, II and III, which is useful to diverse device applications [6-9]. Among them, type-II band alignment plays a critical role in facilitating the electron and hole separation, resulting in long photogenerated carrier lifetimes and tunable interlayer coupling [10-13]. However, the type-II band alignments of most 2D vdWHs are sensitive to external conditions [14-16], which brings a challenge for nanodevices to achieve stable properties. Therefore, it is valuable to find 2D vdWHs with robust type-II band alignment.
More recently, MoSi2N4 monolayer, an ultrathin transition metal nitride, is in the spotlight for its successful synthesis via passivating MoN2 monolayer with silicon [17]. The MoSi2N4 monolayer is a seven-layer nanosheet consisting of one MoN2 inner layer and two Si-N outer layers [18]. Furthermore, MoSi2N4 is reported to show great potential application due to its superior performance [19-31], including the impressive ambient stability [27], large theoretical electron/hole mobility (~270/1200 cm2·V–1·s–1) [17], excellent thermoelectric property [28,29], and remarkable optical absorption in the visible light range [30,31]. Stimulated by the excellent performance, the 2D vdWHs based on MoSi2N4 have become an emerging field [32-41]. For instance, C3N4/MoSi2N4 vdWHs have achieved a tunable type-II band alignment through interlayer coupling and external electric field [38]. The MoSi2N4/NbS2 vdWHs have robust ultra-low p-type Schottky barrier heights (SBH), while the SBH of MoSi2N4/graphene can be modulated via the interlayer distance and electric fields [32]. Compared to isolated transition metal dichalogens (TMDs), the MoSi2N4/TMDs vdWHs possess higher mobility, improved optical absorption, and tunable band structures [37, 40, 41]. Additionally, Jin et al. [42] have found that the vdWHs based on MoSi2N4 and wide-band gap semiconductors, such as ZnO and GaN, have retained the type-II band alignment when the electric field is smaller, but there is a transition from type-II to type-I band alignment with the application of out-of-plane strain. Among the wide-band gap semiconductors, monolayer GeO2 features an excitonic gap in the deep ultraviolet (UV) region with an energy of 6.24 eV [43]. It is found that the GeO2 has a relatively high work function, which is higher than that of reported results for the thin films of transition metal oxides [44] and traditional monolayers of TMDs [45], suggesting that GeO2 is advisable to be integrated into optical and electronic devices as charge extraction and injection layers that provide improvements in the device ability. Also, the 2D GeO2 material is fabricated by strictly controlling the degree of oxidation at the metal-gas interface [46].
Considering better lattice mismatch and interesting performance between the GeO2 and MoSi2N4 monolayers, and there are few works on GeO2-based vdWHs, so we select GeO2 as representative one of wide-band gap semiconductors to construct the heterojunction with MoSi2N4 and study the corresponding performance. In this work, the most stable structures are obtained by the global structure search method of passing through 12 × 12 displacement grids. Additionally, the robust type-II band arrangement of GeO2/MoSi2N4 vdWHs is insensitive to the regulation of biaxial strain, interlayer distance and external electric field, which opens up more possibilities for developing the stable optoelectronic devices.

2 Computational details

All the calculations are gotten via the Vienna ab initio simulation package (VASP) based on DFT [47]. With the framework of generalized gradient approximation (GGA), the Perdew−Burke−Ernzerhof (PBE) [48] functional is employed to deal with the exchange-correlation energy, and the Heyd−Scuseria−Ernzerhof (HSE06) method is utilized to gain more precise band gap value [49]. In addition, the projector augmented wave (PAW) pseudopotentials are adopted to simulate the electron-ion potential [50]. The vdW interlayer interaction in the GeO2/MoSi2N4 heterostructure is corrected by DFT-D3 correction method. Besides, the out-of-plane vacuum region, about 20 Å, is used, and the plane-wave cutoff is employed as 500 eV. A 11×11×1 k-point mesh is adopted for the sampling in reciprocal space. The structures are relaxed until the Hellmann−Feynman forces and the total energy converge to 0.01 eV/Å and 10−5 eV, respectively. Moreover, we would like to point out that the dipole correction is utilized in our calculations.

3 Results and discussion

3.1 Crystal structure and stability of isolated layer and heterojunctions

The geometric structures, band structures and phonon dispersion curves of GeO2 and MoSi2N4 monolayers are listed in Fig.1. Both GeO2 and MoSi2N4 monolayers have hexagonal structures, as shown in Fig.1(a) and (d), and Tab.1 demonstrates that the optimized lattice parameters are a = b = 2.909 Å and a = b = 2.911 Å, respectively. All these results are consistent with previous reports [22,43]. The bond lengths of monolayers are 1.95 Å for Ge-O, 2.09 Å for Mo-N, and 1.75 Å for Si-N. Fig.1(b) and (e) indicate that the GeO2 and MoSi2N4 monolayers are semiconductors with indirect band structure, and the band gap of monolayer GeO2 and MoSi2N4 is 3.56 (5.32) and 1.79 (2.35) eV for PBE (HSE06) calculation, respectively. In addition, Fig.1(c) and (f) depict the phonon dispersion curves of GeO2 and MoSi2N4 monolayers. Obviously, there are no imaginary phonon modes, which confirms the dynamical stability of the two kinds of 2D materials.
Fig.1 (a, d) The atomic structures, (b, e) band structures, (c, f) phonon dispersion curves of GeO2 and MoSi2N4 monolayers, respectively. The Fermi level is defined as zero. In the diagram of the atomic structures, different colored balls represent different atoms.

Full size|PPT slide

Tab.1 The structural and electronic parameters. The lattice constant (a), interlayer distance (d), total energy (Etotal), and band gap of PBE (EgPBE) and HSE06 (EgHSE06).
a (Å) d (Å) Etotal (eV) EgPBE (eV) EgHSE06 (eV)
MoSi2N4 2.911 −61.603 1.79 2.35
GeO2 2.909 −18.88 3.56 5.32
GeO2/MoSi2N4 2.909 2.78 −82.816 0.275 0.946
In order to obtain the credible structures of 2D GeO2/MoSi2N4 vdWHs, a global structure search method is implemented. The total energy as a function of the in-plane shift is presented in Fig.2(a). We consider 144 stacking structures Am,n passing through a 12 × 12 displacement grids. The Am,n means the displacement of the GeO2 layer relative to the MoSi2N4 layer is (m12)a+(n12)b, where a and b are the lattice vectors of the vdWHs. Among them, the A3,6 stacking pattern is the most stable structure with the interlayer distance between GeO2 and MoSi2N4 layer of 2.78 Å. Also, Fig.2(b) shows the side view of the most stable vdWHs structure consisting of GeO2 and MoSi2N4 layer, in which the lattice mismatch is 0.07%. From the bright red region in Fig.2(a), we find that there are four metastable structures around the A3,6 stacking pattern, which feature that the total energy is close to that of the most stable structure. The total energy of these four metastable structures is shown in Table S1. In addition, Fig. S1 presents the projected band structures of the metastable and most stable structures calculated by the PBE and HSE06 methods. Similar total energy and electronic properties are given for the considered five configurations, and thus only the most stable structure is discussed in the following section. The HSE06 method is adopted for all the next calculations. Fig.2(c) indicates that GeO2 and MoSi2N4 can form a type-II heterojunction with direct band gap of 0.946 eV. To further determine the dynamical stability of the structure with the lowest energy, Fig.2(d) shows the phonon dispersion curves of the vdWHs, and all the phonon frequencies are positive, illustrating that the most stable GeO2/MoSi2N4 vdWHs is also dynamically stable.
Fig.2 (a) Color contour plot of total energy versus in-plane shift for GeO2/MoSi2N4 vdWHs, considering the 12 × 12 grids. (b) The geometric structure, (c) band structure and (d) phonon dispersion curves of the most stable GeO2/MoSi2N4 vdWHs. The Fermi level is defined as zero.

Full size|PPT slide

To further explore the electronic characteristics of the most stable structure GeO2/MoSi2N4 vdWHs, Fig.3 draws the orbital-projected electronic band structures, partial charge densities, plane-averaged differential charge density and electrostatic potential. As illustrated in Fig.3(a) and (b), the CBM is mostly contributed by the Ge_s and O_p orbitals of the GeO2 layer, while the VBM mainly comes from the N_p and Mo_d orbitals in the MoSi2N4 layer. Besides, the band shapes of GeO2 and MoSi2N4 are well preserved after forming the heterojunction due to the vdWHs interaction. The charge distribution of Fig.3(c) shows the CBM is mainly contributed by Ge and O atoms, and the VBM comes from Mo and N atoms, which coincides with the above discussion.
Fig.3 The orbital-projected band structures of GeO2/MoSi2N4 vdWHs on GeO2 (a) and MoSi2N4 (b) layer. Partial charge density (c) and plane-averaged charge density difference (d) of GeO2/MoSi2N4 vdWHs. Yellow (green) colored isosurfaces represent the charge accumulation (depletion), respectively. (e) Electrostatic potential of GeO2/MoSi2N4 vdWHs. The isosurface is set to be 0.008 e/bohr3 for (c) and 0.0001 e/ Å3 for (d).

Full size|PPT slide

The plane-averaged charge density difference is calculated to assess the charge redistribution in the GeO2/MoSi2N4 vdWHs, which can be calculated as Δρ= ρvdWHsρGeO2ρMoSi2N4, where the ρvdWHs, ρGeO2, and ρMoSi2N4 represent the charge density of the vdWHs, isolated GeO2 and MoSi2N4 monolayers, respectively. In the inset of Fig.3(d), the charge transfer mainly occurs at the interface of the vdWHs. The yellow and green region corresponds to the ρ > 0 and ρ < 0, indicating the accumulation and depletion of charge, respectively. Meanwhile, the charge is depleted near the MoSi2N4 layer and accumulated in the vicinity of the GeO2 layer, which implies that the charge can be transferred from the MoSi2N4 layer to the GeO2 layer.
In addition, the electrostatic potential along the perpendicular direction of the vdWHs is explored in Fig.3(e). The potential drop (ΔV) across the two isolated layers is 10.51 eV, which indicates the presence of a strong built-in field across the interface. The electric field strength of GeO2/MoSi2N4 vdWHs exceeds that of other MoSi2N4-based heterostructures, such as the C3N4/MoSi2N4 (6.28 eV) [38] and MoSe2/MoSi2N4 (4.03 eV) [40]. The MoSi2N4 possesses a much deeper potential than GeO2, owing to the stronger electronegativity of the MoSi2N4 monolayer. Thus, the deeper potential may facilitate the separation of electron-hole pairs, which is beneficial for photoelectric detection applications based on the GeO2/MoSi2N4 vdWHs.

3.2 Biaxial strain tunable electronic properties of the vdWHs

To understand the strain influence on the GeO2/MoSi2N4 vdWHs, we define the biaxial strain as ε = (aa0)/a0 on GeO2/MoSi2N4 vdWHs, where a0 and a represent the lattice parameter of heterostructures without and with biaxial strain, respectively. In Fig.4, the negative and positive values are used to denote the compressive and tensile strain, respectively. Fig.4(a−c) display the variation of the total energy, band gap, and interlayer charge transfer of GeO2/MoSi2N4 vdWHs under the biaxial strains. It is obvious from Fig.4(a) that the total energy reaches the minimum value under the pristine condition. The band gap decreases for GeO2/MoSi2N4 vdWHs with increasing tensile, which can be seen in Fig.4(b). On the contrary, as the compressive strain is applied, the band gap value tends to increase. When the tensile strain increases up to 6%, the band gap of the vdWHs decreases from 0.946 eV to 0.124 eV. And the variation trend of the band gap gradually slows down and approaches saturation at 4% strain. Additionally, when the compressive strain range is 0 ~ −6%, the band gap increases linearly from 0.946 eV to 2.806 eV. Moreover, it can realize a transition from direct to indirect band gap when the compressive strain is greater than -3%. As shown in Tab.2, we can see that with increasing the tensile strain for the GeO2/MoSi2N4 vdWHs, the atomic bond lengths of Ge-O, Mo-N and Si-N can be raised, while the bond angles exhibit the reduced trend. In contrast to the case of tensile strain, the increasing compressive strain can make the Ge-O, Mo-N and Si-N bond lengths decrease, and the bond angles consequently enhance. However, it is worth noting that the type-II band arrangement is well preserved in the GeO2/MoSi2N4 vdWHs under all the studied biaxial strain cases.
Tab.2 The bond lengths as a function of biaxial strain for GeO2/MoSi2N4 vdWHs.
Strain dGe-O (Å) dMo-N (Å) dSi-N (Å)
−6% 1.90 2.05 1.68
−5% 1.91 2.06 1.70
−4% 1.91 2.06 1.71
−3% 1.92 2.07 1.72
−2% 1.93 2.08 1.73
−1% 1.94 2.08 1.74
0 1.95 2.09 1.75
1% 1.96 2.10 1.76
2% 1.96 2.11 1.78
3% 1.97 2.12 1.79
4% 1.99 2.13 1.81
5% 2.00 2.14 1.82
6% 2.01 2.15 1.83
Fig.4 The variation of the total energy (a), band gap (b), and interlayer charge transfer (c) of GeO2/MoSi2N4 vdWHs as a function of biaxial strain, respectively. (d−g) Projected band structures of GeO2/MoSi2N4 vdWHs with the biaxial strains of −3%, −2%, 2% and 3%.

Full size|PPT slide

Fig.4(c) exhibits the evolution of interlayer charge transfer as a function of biaxial strain. The results show that the interlayer charge transfer of GeO2/MoSi2N4 vdWHs decreases with increasing tensile and compressive strains. Under the case of tensile and compressive strain, the interlayer charge transfer declines by 0.018 e and 0.005 e, respectively. The charge transfer tends to diminish when the compressive strain is larger than −5%. Furthermore, it reaches a maximum value of 0.031 e under the 1% tensile strain.
To better understand the above-mentioned behavior, Fig.4(d)−(g) present the projected band structures of GeO2/MoSi2N4 vdWHs under different biaxial strains. When tensile strain is applied, the conduction band edge moves to the Fermi level significantly, while the valence band edge barely moves, which is consistent with the decrease in the band gap. On the other side, the compressive strain causes the conduction band edge to move away from the Fermi level substantially, corresponding to the increase of the band gap. As the −2% compressive strain is applied, both CBM and VBM remain at Γ point. However, when the compressive strain is greater than −3%, the CBM of GeO2/MoSi2N4 vdWHs is at Γ point, while the VBM is at K point, thus an indirect band gap can be observed. In particular, for all the studied strain cases, the CBM and VBM of GeO2/MoSi2N4 vdWHs are always formed by the GeO2 and MoSi2N4 layers, respectively, confirming that it always maintains type-II band alignment characteristics. In addition, we notice that the application of strain can induce the band alignment transition from type-II to type-I or type-III in many other 2D vdWHs, such as GaSe/SnX2 (X = S, Se) [14], MoSSe/MBP [51] and BP/β-AsP vdWHs [52].

3.3 Interlayer distance tunable electronic properties of the vdWHs

The interlayer interaction has obvious effects on the electronic structures of the vdWHs. Thus, Fig.5 plots the interlayer distance change on the electronic properties of the GeO2/MoSi2N4 vdWHs. We can see from Fig.5(a) that the curve of the total energy has the minimum energy value when the interlayer distance is 2.78 Å. From Fig.5(b), the band gap increases for GeO2/MoSi2N4 vdWHs with increasing interlayer distance. Also, the band gap changes slightly when the interlayer distance is larger than 3.28 Å. In addition, Fig.5(c) shows that the interlayer charge transfer declines gradually with the increase of the interlayer distance. To reveal the related physical mechanism, in Fig.5(d−g), we give the interlayer distance effects on the projected band structure of GeO2/MoSi2N4 vdWHs. For different distances, the CBM and VBM contributions to GeO2/MoSi2N4 vdWHs are provided by the GeO2 and MoSi2N4 layers separately, showing that this type-II band alignment is insensitive to the interlayer interactions. Furthermore, the variation of the band gap is mainly due to the up and down movement of CBM.
Fig.5 The total energy (a) and band gap (b), interlayer charge transfer (c) and projected band structures (d−g) of GeO2/MoSi2N4 vdWHs at different interlayer distances, respectively.

Full size|PPT slide

3.4 Electric field tunable electronic properties of the vdWHs

The external vertical electric field provides another effective strategy for controlling the electronic structure. In the following, the variation of the band gap for GeO2/MoSi2N4 vdWHs as a function of the electric field is shown in Fig.6 (a). It should be noted that the electric field is perpendicular to the surface of the GeO2/MoSi2N4 vdWHs, and the positive direction is defined from MoSi2N4 layer to GeO2 layer, as illustrated in Fig. S2. When the electric field increases from −0.5 to 0.5 V/Å, the band gap decreases linearly, maintaining the type-II band alignment and direct band structures. Here, the band edge and projected band structures are also shown, and the reasons for the band alignment transition are discussed. In Fig.6(b), with the increase of the external electric field, the CBM of GeO2/MoSi2N4 vdWHs decreases significantly, while the VBM shifts upward slightly. Moreover, when the positive electric field reaches 0.4 V/Å, the VBM decreases more slightly, which causes the value of band gap to decrease slowly as the electric field changes from 0.4 V/Å to 0.5 V/Å. As can be seen from Fig.6(c−f), with the increase of the positive electric field, the CBM of GeO2 layer moves down gradually, while the CBM and VBM of MoSi2N4 keep almost unchanged. Compared with the 0.2 V/Å, the CBM of GeO2 layer moves towards the Fermi level under the electric field of 0.4 V/Å. On the other hand, Fig.6(f) indicates that with increasing the negative electric field, the CBM of GeO2 layer moves away from the Fermi level when the electric field reaches −0.4 V/Å. However, further increased electric field does not change the type of band arrangement. Thus, the external electric field can change the band gap, while the band alignment remains type-II for the GeO2/MoSi2N4 vdWHs. What’s more, for many other vdWHs, the band alignment is sensitive to the external electric field [16, 51, 53-56]. For example, when the positive electric field is larger than 0.2 V/Å, the band alignment of the AlN/VSe2 vdWHs changes from type-II to type-I [16]. The MoXY/WXY (X, Y = S, Se, Te; X ≠ Y) vdWHs maintain robust type-II band alignment under positive electric field, while a reverse electric field can make the band alignment change [53].
Fig.6 The band gap (a) and band alignment (b) of GeO2/MoSi2N4 vdWHs as a function of external electric field, respectively. (c−f) The projected band structures of GeO2/MoSi2N4 vdWHs with different external electric fields. The red (green) line indicates the contribution of the GeO2 (MoSi2N4) layer.

Full size|PPT slide

4 Conclusions

In conclusion, by using a global structure search approach, we design the GeO2/MoSi2N4 vdWHs to exactly explore the band structures, and the corresponding modulations of biaxial strain, electric field, and interlayer distance are considered to improve the performance of GeO2/MoSi2N4 vdWHs. The results show that the monolayer GeO2, MoSi2N4 and GeO2/MoSi2N4 vdWHs are dynamically stable at room temperature. The most stable GeO2/MoSi2N4 vdWHs possess direct band gap of 0.946 eV and type-II band alignment with the CBM and VBM from the GeO2 and MoSi2N4 layers. When the biaxial strain is applied, the direct-to-indirect band gap transition can be achieved. More interestingly, the robust type-II band arrangement of GeO2/MoSi2N4 vdWHs remains unchanged when the external electric field, interlayer distance, and biaxial strain are applied. These results reveal that the GeO2/MoSi2N4 vdWHs can be in promising optoelectronic devices with stable characteristics due to their robust type-II band arrangement.

References

[1]
W. Wang, C. Si, W. Lei, F. Xiao, Y. Liu, C. Autieri, X. Ming. Stacking order and Coulomb correlation effect in the layered charge density wave phase of 1T-NbS2. Phys. Rev. B, 2022, 105(3): 035119
CrossRef ADS Google scholar
[2]
A. Aharon-Steinberg, A. Marguerite, D. J. Perello, K. Bagani, T. Holder, Y. Myasoedov, L. S. Levitov, A. K. Geim, E. Zeldov. Long-range nontopological edge currents in charge-neutral graphene. Nature, 2021, 593(7860): 528
CrossRef ADS Google scholar
[3]
X. Li, P. Yuan, L. Li, M. He, J. Li, C. Xia. Sub-5-nm monolayer GaSe MOSFET with ultralow subthreshold swing and high on-state current: Dielectric layer effect. Phys. Rev. Appl., 2022, 18(4): 044012
CrossRef ADS Google scholar
[4]
T. Wang, A. Dong, X. Zhang, R. K. Hocking, C. Sun. Theoretical study of K3Sb/graphene heterostructure for electrochemical nitrogen reduction reaction. Front. Phys., 2022, 17(2): 23501
CrossRef ADS Google scholar
[5]
Q. Q. Kong, X. G. An, J. Zhang, W. T. Yao, C. H. Sun. Design of heterojunction with components in different dimensions for electrocatalysis applications. Front. Phys., 2022, 17(4): 43601
CrossRef ADS Google scholar
[6]
C. Long, Y. Dai, Z. R. Gong, H. Jin. Robust type-II band alignment in Janus-MoSSe bilayer with extremely long carrier lifetime induced by the intrinsic electric field. Phys. Rev. B, 2019, 99(11): 115316
CrossRef ADS Google scholar
[7]
N. Ubrig, E. Ponomarev, J. Zultak, D. Domaretskiy, V. Zolyomi, D. Terry, J. Howarth, I. Gutierrez-Lezama, A. Zhukov, Z. R. Kudrynskyi, Z. D. Kovalyuk, A. Patane, T. Taniguchi, K. Watanabe, R. V. Gorbachev, V. I. Fal’ko, A. F. Morpurgo. Design of van der Waals interfaces for broad-spectrum optoelectronics. Nat. Mater., 2020, 19(3): 299
CrossRef ADS Google scholar
[8]
J. G. Azadani, S. Lee, H. R. Kim, H. Alsalman, Y. K. Kwon, J. Tersoff, T. Low. Simple linear response model for predicting energy band alignment of two-dimensional vertical heterostructures. Phys. Rev. B, 2021, 103(20): 205129
CrossRef ADS Google scholar
[9]
F. H. Davies, C. J. Price, N. T. Taylor, S. G. Davies, S. P. Hepplestone. Band alignment of transition metal dichalcogenide heterostructures. Phys. Rev. B, 2021, 103(4): 045417
CrossRef ADS Google scholar
[10]
P. Rivera, J. R. Schaibley, A. M. Jones, J. S. Ross, S. Wu, G. Aivazian, P. Klement, K. Seyler, G. Clark, N. J. Ghimire, J. Yan, D. G. Mandrus, W. Yao, X. Xu. Observation of long-lived interlayer excitons in monolayer MoSe2−WSe2 heterostructures. Nat. Commun., 2015, 6(1): 6242
CrossRef ADS Google scholar
[11]
P. Rivera, K. L. Seyler, H. Yu, J. R. Schaibley, J. Yan, D. G. Mandrus, W. Yao, X. Xu. Valley-polarized exciton dynamics in a 2D semiconductor heterostructure. Science, 2016, 351(6274): 688
CrossRef ADS Google scholar
[12]
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., 2020, 16(1): 13501
[13]
X. Li, T. Liu, L. Li, M. He, C. Shen, J. Li, C. Xia. Reconfifigurable band alignment of m-GaS/n-XTe2 (X = Mo, W) multilayer van der Waals heterostructures for photoelectric applications. Phys. Rev. B, 2022, 106(12): 125306
CrossRef ADS Google scholar
[14]
D. Wijethunge, L. Zhang, C. Tang, A. Du. Tuning band alignment and optical properties of 2D van der Waals heterostructure via ferroelectric polarization switching. Front. Phys., 2020, 15(6): 63504
CrossRef ADS Google scholar
[15]
S. Ghosh, A. Varghese, H. Jawa, Y. Yin, N. V. Medhekar, S. Lodha. Polarity-tunable photocurrent through band alignment engineering in a high-speed WSe2/SnSe2 diode with large negative responsivity. ACS Nano, 2022, 16(3): 4578
CrossRef ADS Google scholar
[16]
Y. Zhu, D. Zhang, H. Ye, D. Bai, M. Li, G. P. Zhang, J. Zhang, J. Wang. Magnetic and electronic properties of AlN/VSe2 van der Waals heterostructures from combined first-principles and Schrödinger−Poisson simulations. Phys. Rev. Appl., 2022, 18(2): 024012
CrossRef ADS Google scholar
[17]
Y. L. Hong, Z. Liu, L. Wang, T. Zhou, W. Ma, C. Xu, S. Feng, L. Chen, M. L. Chen, D. M. Sun, X. Q. Chen, H. M. Cheng, W. Ren. Chemical vapor deposition of layered two-dimensional MoSi2N4 materials. Science, 2020, 369(6504): 670
CrossRef ADS Google scholar
[18]
R. Islam, B. Ghosh, C. Autieri, S. Chowdhury, A. Bansil, A. Agarwal, B. Singh. Tunable spin polarization and electronic structure of bottom-up synthesized MoSi2N4 materials. Phys. Rev. B, 2021, 104(20): L201112
CrossRef ADS Google scholar
[19]
A. Bafekry, M. Faraji, M. M. Fadlallah, A. Bagheri Khatibani, A. abdolahzadeh Ziabari, M. Ghergherehchi, S. Nedaei, S. F. Shayesteh, D. Gogova. Tunable electronic and magnetic properties of MoSi2N4 monolayer via vacancy defects, atomic adsorption and atomic doping. Appl. Surf. Sci., 2021, 559: 149862
CrossRef ADS Google scholar
[20]
Q. Wu, L. Cao, Y. S. Ang, L. K. Ang. Semiconductor-to-metal transition in bilayer MoSi2N4 and WSi2N4 with strain and electric field. Appl. Phys. Lett., 2021, 118(11): 113102
CrossRef ADS Google scholar
[21]
H. Zhong, W. Xiong, P. Lv, J. Yu, S. Yuan, Strain-induced semiconductor to metal transition in MA2Z4 bilayers (M=Ti . Mo; A=Si; Z=N, P). Phys. Rev. B, 2021, 103(8): 085124
CrossRef ADS Google scholar
[22]
S. Li, W. Wu, X. Feng, S. Guan, W. Feng, Y. Yao, S. A. Yang, Valley-dependent properties of monolayer MoSi2N4 . WSi2N4, and MoSi2As4. Phys. Rev. B, 2020, 102(23): 235435
CrossRef ADS Google scholar
[23]
T. Zhong, Y. Ren, Z. Zhang, J. Gao, M. Wu. Sliding ferroelectricity in two-dimensional MoA2N4 (A = Si or Ge) bilayers: High polarizations and Moiré potentials. J. Mater. Chem. A, 2021, 9(35): 19659
CrossRef ADS Google scholar
[24]
L. Wang, Y. Shi, M. Liu, A. Zhang, Y. L. Hong, R. Li, Q. Gao, M. Chen, W. Ren, H. M. Cheng, Y. Li, X. Q. Chen. Intercalated architecture of MA2Z4 family layered van der Waals materials with emerging topological, magnetic and superconducting properties. Nat. Commun., 2021, 12(1): 2361
CrossRef ADS Google scholar
[25]
Q. Wu, L. K. Ang. Giant tunneling magnetoresistance in atomically thin VSi2N4/MoSi2N4/VSi2N4 magnetic tunnel junction. Appl. Phys. Lett., 2022, 120(2): 022401
CrossRef ADS Google scholar
[26]
Y. Zang, Q. Wu, W. Du, Y. Dai, B. Huang, Y. Ma. Activating electrocatalytic hydrogen evolution performance of two-dimensional MSi2N4(M=Mo, W): A theoretical prediction. Phys. Rev. Mater., 2021, 5(4): 045801
CrossRef ADS Google scholar
[27]
J. Yuan, Q. Wei, M. Sun, X. Yan, Y. Cai, L. Shen, U. Schwingenschlögl. Protected valley states and generation of valley- and spin-polarized current in monolayer MA2Z4. Phys. Rev. B, 2022, 105(19): 195151
CrossRef ADS Google scholar
[28]
B. Mortazavi, B. Javvaji, F. Shojaei, T. Rabczuk, A. V. Shapeev, X. Y. Zhuang. Exceptional piezoelectricity, high thermal conductivity and stiffness and promising photocatalysis in two-dimensional MoSi2N4 family confirmed by first-principles. Nano Energy, 2021, 82: 105716
CrossRef ADS Google scholar
[29]
Y. Yin, M. Yi, W. Guo. High and anomalous thermal conductivity in monolayer MSi2Z4 semiconductors. ACS Appl. Mater. Interfaces, 2021, 13(38): 45907
CrossRef ADS Google scholar
[30]
C. C. Jian, X. C. Ma, J. Q. Zhang, X. Yong. Strained MoSi2N4 monolayers with excellent solar energy absorption and carrier transport properties. J. Phys. Chem. C, 2021, 125(28): 15185
CrossRef ADS Google scholar
[31]
A. Bafekry, C. Stampfl, M. Naseri, M. M. Fadlallah, M. Faraji, M. Ghergherehchi, D. Gogova, S. A. H. Feghhi. Effect of electric field and vertical strain on the electro−optical properties of the MoSi2N4 bilayer: A first-principles calculation. J. Appl. Phys., 2021, 129(15): 155103
CrossRef ADS Google scholar
[32]
L. Cao, G. Zhou, Q. Wang, L. K. Ang, Y. S. Ang. Two-dimensional van der Waals electrical contact to monolayer MoSi2N4. Appl. Phys. Lett., 2021, 118(1): 013106
CrossRef ADS Google scholar
[33]
J. Zhao, X. H. Jin, H. Zeng, C. Yao, G. Yan. Spin-valley coupling and valley splitting in the MoSi2N4/CrCl3 van der Waals heterostructure. Appl. Phys. Lett., 2021, 119(21): 213101
CrossRef ADS Google scholar
[34]
Y. Ding, Y. Wang. First-principles study of two-dimensional MoN2X2Y2 (X = B~In, Y = N~Te) nanosheets: The III–VI analogues of MoSi2N4 with peculiar electronic and magnetic properties. Appl. Surf. Sci., 2022, 593: 153317
CrossRef ADS Google scholar
[35]
C. Nguyen, N. V. Hoang, H. V. Phuc, A. Y. Sin, C. V. Nguyen. Two-dimensional boron phosphide/MoGe2N4 van der Waals heterostructure: A promising tunable optoelectronic material. J. Phys. Chem. Lett., 2021, 12(21): 5076
CrossRef ADS Google scholar
[36]
Y. Guo, Y. Dong, X. Cai, L. Liu, Y. Jia. Controllable Schottky barriers and contact types of BN intercalation layers in graphene/MoSi2As4 vdW heterostructures via applying an external electrical field. Phys. Chem. Chem. Phys., 2022, 24(30): 18331
CrossRef ADS Google scholar
[37]
Z. Zhang, G. Chen, A. Song, X. Cai, W. Yu, X. Jia, Y. Jia. Optoelectronic properties of bilayer van der Waals WSe2/MoSi2N4 heterostructure: A first-principles study. Physica E, 2022, 144: 115429
CrossRef ADS Google scholar
[38]
C. Q. Nguyen, Y. S. Ang, S. T. Nguyen, N. V. Hoang, N. M. Hung, C. V. Nguyen. Tunable type-II band alignment and electronic structure of C3N4/MoSi2N4 heterostructure: Interlayer coupling and electric field. Phys. Rev. B, 2022, 105(4): 045303
CrossRef ADS Google scholar
[39]
Y. T. Ren, L. Hu, Y. T. Chen, Y. J. Hu, J. L. Wang, P. L. Gong, H. Zhang, L. Huang, X. Q. Shi. Two-dimensional MSi2N4 monolayers and van der Waals heterostructures: Promising spintronic properties and band alignments. Phys. Rev. Mater., 2022, 6(6): 064006
CrossRef ADS Google scholar
[40]
X.CaiZhangZ. ZhuY.LinL.YuW.Wang Q.YangX. JiaX.JiaY., A two-dimensional MoSe2/MoSi2N4 van der Waals heterostructure with high carrier mobility and diversified regulation of its electronic properties, J. Mater. Chem. C 9(31), 10073 (2021)
[41]
A.BafekryM. FarajiA.Abdollahzadeh ZiabariM.M. FadlallahC.V. NguyenM.GhergherehchiS.A. H. Feghhi, A van der Waals heterostructure of MoS2/MoSi2N4: a first-principles study, New J. Chem. 45(18), 8291 (2021)
[42]
J. Q. Ng, Q. Wu, L. K. Ang, Y. S. Ang. Tunable electronic properties and band alignments of MoSi2N4/GaN and MoSi2N4/ZnO van der Waals heterostructures. Appl. Phys. Lett., 2022, 120(10): 103101
CrossRef ADS Google scholar
[43]
Y. Sozen, M. Yagmurcukardes, H. Sahin. Vibrational and optical identification of GeO2 and GeO single layers: A first-principles study. Phys. Chem. Chem. Phys., 2021, 23(37): 21307
CrossRef ADS Google scholar
[44]
S. Chuang, C. Battaglia, A. Azcatl, S. McDonnell, J. S. Kang, X. Yin, M. Tosun, R. Kapadia, H. Fang, R. M. Wallace, A. Javey. MoS2 p-type transistors and diodes enabled by high work function MoOx contacts. Nano Lett., 2014, 14(3): 1337
CrossRef ADS Google scholar
[45]
H. Kim, H. J. Choi. Thickness dependence of work function, ionization energy, and electron affinity of Mo and W dichalcogenides from DFT and GW calculations. Phys. Rev. B, 2021, 103(8): 085404
CrossRef ADS Google scholar
[46]
B. Y. Zhang, K. Xu, Q. Yao, A. Jannat, G. Ren, M. R. Field, X. Wen, C. Zhou, A. Zavabeti, J. Z. Ou. Hexagonal metal oxide monolayers derived from the metal-gas interface. Nat. Mater., 2021, 20(8): 1073
CrossRef ADS Google scholar
[47]
G. Kresse, J. Furthmuller. 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]
J. P. Perdew, K. Burke, M. Ernzerhof. Generalized gradient approximation made simple. Phys. Rev. Lett., 1996, 77(18): 3865
CrossRef ADS Google scholar
[49]
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
[50]
G. Kresse, D. Joubert. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B, 1999, 59(3): 1758
CrossRef ADS Google scholar
[51]
Y. Mogulkoc, R. Caglayan, Y. O. Ciftci. Band alignment in monolayer boron phosphide with Janus MoSSe heterobilayers under strain and electric field. Phys. Rev. Appl., 2021, 16(2): 024001
CrossRef ADS Google scholar
[52]
Y. L. Liu, Y. Shi, H. Yin, C. L. Yang. Two-dimensional BP/β-AsP van der Waals heterostructures as promising photocatalyst for water splitting. Appl. Phys. Lett., 2020, 117(6): 063901
CrossRef ADS Google scholar
[53]
S. Patel, U. Dey, N. P. Adhikari, A. Taraphder. Electric field and strain-induced band-gap engineering and manipulation of the Rashba spin splitting in Janus van der Waals heterostructures. Phys. Rev. B, 2022, 106(3): 035125
CrossRef ADS Google scholar
[54]
M. Yagmurcukardes, E. Torun, R. T. Senger, F. M. Peeters, H. Sahin. Mg(OH)2−WS2 van der Waals heterobilayer: Electric field tunable band-gap crossover. Phys. Rev. B, 2016, 94(19): 195403
CrossRef ADS Google scholar
[55]
K. Iordanidou, J. Wiktor. Two-dimensional MoTe2/SnSe2 van der Waals heterostructures for tunnel-FET applications. Phys. Rev. Mater., 2022, 6(8): 084001
CrossRef ADS Google scholar
[56]
K. Liang, T. Huang, K. Yang, Y. Si, H. Y. Wu, J. C. Lian, W. Q. Huang, W. Y. Hu, G. F. Huang. Dipole engineering of two-dimensional van der Waals heterostructures for enhanced power-conversion efficiency: The case of Janus Ga2SeTe/InS. Phys. Rev. Appl., 2021, 16(5): 054043
CrossRef ADS Google scholar

Electronic supplementary material

Supplementary materials are available in the online version of this article at https://doi.org/10.1007/s11467-022-1216-8 and https://journal.hep.com.cn/fop/EN/10.1007/s11467-022-1216-8 and are accessible for authorized users.

Acknowledgements

This work was supported by the National Natural Science Foundation of China under Grant Nos. 11904085 and 12074103. Program for Outstanding Youth of Henan Province under Grant No. 202300410221. Henan Normal University Innovative Science and Technology Team under Grant No. 20200185. The calculations were also supported by the High Performance Computing Center of Henan Normal University.

RIGHTS & PERMISSIONS

2023 Higher Education Press
AI Summary AI Mindmap
PDF(3766 KB)

Supplementary files

fop-21216-OF-xiacongxin_suppl_1 (520 KB)

1044

Accesses

5

Citations

Detail

Sections
Recommended

/