† Corresponding author. E-mail:
Heterostructures (HSs) have attracted significant attention because of their interlayer van der Waals interactions. The electronic structures and optical properties of stacked GaN–MoS2 HSs under strain have been explored in this work using density functional theory. The results indicate that the direct band gap (1.95 eV) of the GaN–MoS2 HS is lower than the individual band gaps of both the GaN layer (3.48 eV) and the MoS2 layer (2.03 eV) based on HSE06 hybrid functional calculations. Specifically, the GaN–MoS2 HS is a typical type-II band HS semiconductor that provides an effective approach to enhance the charge separation efficiency for improved photocatalytic degradation activity and water splitting efficiency. Under tensile or compressive strain, the direct band gap of the GaN–MoS2 HS undergoes redshifts. Additionally, the GaN–MoS2 HS maintains its direct band gap semiconductor behavior even when the tensile or compressive strain reaches 5% or -5%. Therefore, the results reported above can be used to expand the application of GaN–MoS2 HSs to photovoltaic cells and photocatalysts.
Stacked van der Waals (vdW) heterostructures (HSs) often lead to new physics and to promising opportunities for the design of unusual electronic devices.[1] Therefore, two-dimensional (2D) vertically stacked vdW HSs have drawn the attention of many researchers because of their tunable band gaps and thicknesses of several atomic layers.[2] For example, MoS2–ZnO,[3] MoS2–WSe2,[4] BP–BN,[5] MoSe2–graphene,[6] and AlGaN–GaN[7] have provided novel material platforms to explore a range of innovative applications, including ultrathin photodetectors,[8] solar cells,[9] and tunneling transistor-based memory devices.[10] However, the application of these 2D materials is enabled by their outstanding physical properties. In addition, the electronic structures of these materials are highly sensitive to both thickness and external strain.[11–15] Miwa et al. demonstrated that MoS2–graphene HSs could be grown epitaxially on a semiconducting silicon carbide substrate under ultrahigh vacuum conditions.[16] In a recent work by He et al., the band gap of the MoSe2–WSe2 HS was reduced to 1.48 eV, which was lower than the band gaps of the individual WSe2 (1.60 eV) and MoSe2 (1.50 eV) layers, and the direct or indirect band gaps of HSs have shown redshifts under tensile strains.[17] Overall, strain can be feasibly used to tune the band gaps of these 2D materials.
GaN has recently drawn considerable interest and 2D GaN has been used in optoelectronic applications. Unlike other 2D materials, wurtzite structured bulk GaN is unsuitable for 2D structure formation through exfoliation.[18] Fortunately, 2D GaN has been experimentally synthesized directly via a graphene encapsulation process.[19] Sun et al. systematically investigated the electronic properties of a graphene-g-GaN HS based on first-principles calculations and discovered that an increase in the interlayer distance from 2.5 Å to 4.5 Å led to a transition from a p-type to an n-type Schottky contact.[20] Subsequently, the stability properties of 2D GaN have also been confirmed theoretically.[21–23] Sanders et al. investigated the electronic and optical properties of both monolayer and bilayer 2D GaN.[24] Accurate band gaps, excitonic properties, and the luminescence energies of 2D GaN under various strains were studied via first-principles calculations based on density functional theory (DFT) and many-body perturbation theory, and it was revealed that the quantum confined Stark effect was amplified by the strong polarization field of the material. With its excellent properties, 2D GaN could be applied in fields including energy-efficient display applications, nonlinear optics, and water purification.[24] In addition, 2D GaN and transition metal dichalcogenides can be combined to provide superior physical properties.
2D MoS2, a typical direct band gap (1.8 eV) semiconductor, has also received significant attention.[25,26] Monolayer MoS2 can absorb light in the 510 nm–780 nm range with absorbance of 5%–10%.[26] Therefore, the exploration of the novel structures and excellent properties of GaN–MoS2 HSs is of great importance. To the best of our knowledge, the electronic and optical properties of GaN–MoS2 HS have not been investigated previously. The study of the structural, electronic, and optical properties of GaN–MoS2 HSs is very valuable because of the potential applications of these structures in the fields of photovoltaic cells and photocatalysis.
All the properties of GaN–MoS2 HSs were studied using first-principles calculations based on DFT. The generalized gradient approximation of the Perdew–Burke–Ernzerhof functional[27] was adopted and implemented in Vienna ab initio simulation package software.[28–31] The projector augmented wave method was used to solve the Kohn–Sham equations.[32] An energy cutoff of 600 eV and Brillouin zone sampling of a K-mesh of 11 ×11×1 were set. Additionally, a K-mesh of 15×15×1 was used for the energy band structure calculations. In the relaxation process, the energy criterion was set at 10−5 eV and the residual force was less than 10−3 eV/Å. The vacuum layer thickness was more than 20 Å to prevent spurious interactions with the neighboring image structures. The long-range vdW interactions were important in holding the 2D HS together.[33] The vdW-DF2 functional was selected here to describe the long-range electron correlation effects.[34]
In the optical properties calculations, the Fourier transform of the dielectric function was given as[35]
The wurtzite GaN structure is cleaved along [0001] to form a planar structure, which was previously predicted to be stable.[19] After optimization, the lattice constants of monolayer GaN and MoS2 were determined to be 3.18 Å and 3.20 Å, respectively; these values are consistent with previously reported results.[12,24] The structure of the GaN–MoS2 HS is depicted in Fig.
To assess the vdW interactions between the GaN and MoS2 layers, the distance D between the Ga and S atoms and the binding energy Δ E are calculated for the GaN–MoS2 HS as follows:
Next, the band structure of the GaN–MoS2 HS is derived, as shown in Fig.
To gain an insight into the electronic properties of the GaN–MoS2 HS, the total density of states and the projected density of states are presented in Fig.
For further clarification of the interactions, the electronic band structure of the GaN–MoS2 HS is calculated under various biaxial strains. Strain is known to be one of most effective means to tailor the properties of 2D materials.[11–13,17] Yu et al. calculated the electronic properties of RbGeI3 under strains ranging from −6% to 6% and found that the band gap of the material decreased monotonically.[11] In Fig.
The imaginary part of the dielectric constant ε2 for the GaN–MoS2 HS from −5% to 5% strain is then calculated, as shown in Fig.
In summary, the electronic and optical properties of the GaN–MoS2 HS have been determined from first-principles calculations. The lattice parameter of the GaN–MoS2 HS is 3.19 Å and the structure of this HS is stable. The GaN–MoS2 HS is found to be a typical type-II band structure semiconductor. The application of tensile or compressive strain leads to a reduction of the direct band gap of the GaN–MoS2 HS. These results can offer insights into the interactions between the stacked GaN and MoS2 monolayers in the GaN–MoS2 HS and will help to guide the design of future applications of these HSs in solar cells and photocatalysts.
1 | |
2 | |
3 | |
4 | |
5 | |
6 | |
7 | |
8 | |
9 | |
10 | |
11 | |
12 | |
13 | |
14 | |
15 | |
16 | |
17 | |
18 | |
19 | |
20 | |
21 | |
22 | |
23 | |
24 | |
25 | |
26 | |
27 | |
28 | |
29 | |
30 | |
31 | |
32 | |
33 | |
34 | |
35 | |
36 | |
37 | |
38 |