Next Article in Journal
Preparation, Characterization, and In Vitro Evaluation of the Biological Activity of Several Metal-Based Complexes with Two Widely Used Fluoroquinolone Antibiotics: Lomefloxacin and Pefloxacin Drugs
Previous Article in Journal
Temperature-Dependent Optical Behaviors and Demonstration of Carrier Localization in Polar and Semipolar AlGaN Multiple Quantum Wells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Theoretical Investigation of the Structural and Electronic Properties of P/SnBr2 Heterojunctions

1
School of Materials Science and Engineering, Anhui University, Hefei 230601, China
2
School of Physics and Optoelectronics Engineering, Anhui University, Hefei 230601, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(7), 1077; https://doi.org/10.3390/cryst13071077
Submission received: 14 June 2023 / Revised: 6 July 2023 / Accepted: 6 July 2023 / Published: 8 July 2023

Abstract

:
In this paper, the structural and electronic properties of P/SnBr2 heterojunctions were investigated using the first-principles calculation method based on the density functional theory (DFT). The band alignment of the P/SnBr2 heterojunction was type I. The bandgap value was 0.71 eV in the DFT calculation. Furthermore, the bandgap of the heterojunction could be efficiently tuned by controlling an electric field and biaxial strain. The bandgap changed linearly with the electric field in a certain range; when the electric field was greater than 0.8 V/Å, the heterojunction was metallic. The bandgap could also be tuned when a biaxial strain was applied. Under tensile or compressive stress, significant effects such as the band alignment shift from type I to type III, and the transition from indirect to direct bandgap occurred. In conclusion, these research findings provide theoretical guidance for designing new heterojunctions based on SnBr2.

1. Introduction

The discovery of graphene has led to extensive research on the physical and chemical properties of two-dimensional (2D) materials [1,2]. Nowadays, various kinds of 2D materials, such as the transition metal dichalcogenides (TMDCs) [3,4], hexagonal boron nitride (h-BN) [5] and phosphorylenes [6,7,8], are common platforms used in optoelectronic and electronic applications due to their unique material properties [9,10], and have attracted much attention. Due to their valuable optical, electronic, chemical, magnetic and mechanical properties, which are significantly different from the conventional bulk materials, 2D materials show great potential for the development of various nanodevices [11,12].
Unlike most multilayer TMDCs with indirect bandgaps, phosphorene of different thicknesses always have a direct bandgap, which is a significant advantage for optoelectronic applications. Moreover, the carrier mobility of phosphorene is notably greater than that of other 2D materials, such as the transition metal dichalcogenides (TMDCs) [13,14]. When the thickness of phosphorene is about 10 nm, the ground carrier mobility of black phosphorus can reach 1000 cm2·V−1·s−1 [15]. Compared with graphene, phosphorene can conduct electrons at a rate similar to graphene, and has a considerable bandgap, which makes it significantly better than graphene with zero bandgap. The outstanding feature of layered phosphorene is its in-plane anisotropy, which results in an angle-dependent optical conductivity and carrier mobility [16,17]. Due to its unique structure, excellent physical properties and anisotropy, phosphorene has great potential in field-effect transistors (FETs), photodetectors, batteries, thermoelectric applications and other fields [18,19,20], while it is rare in other 2D materials. The monolayer structure of phosphorene is the second basic monolayer material after graphene, which can be exfoliated from the bulk crystal to produce unique properties. Tin halide compounds and related materials are being explored as promising lead-free hybrid materials. There has been significant progress in developing organic tin halide compounds with varying structures and controllable dimensions on a molecular level. Recently, lead-free metal halide perovskite and hybrid power mines have attracted more and more interest in various optoelectronic applications, from solar cells to light-emitting diodes and sensors [21,22].
In recent years, the discovery of more and more 2D materials has triggered the explosive expansion in the study of van der Waals (vdW) heterostructures [23,24,25]. Due to suspension-free bonding surfaces and weak vdW interlayer interactions, different 2D materials can be stacked vertically to form different heterostructures. These heterostructures are not limited to the properties of a single 2D material, and can combine the advantages of the individual components to achieve different functions [26,27]. By forming heterostructures, one can not only break the limitations of monolayer materials, but also the new structures of multilayer materials often show excellent optoelectronic properties, which expands the application of this material in the new generation of optoelectronic nanodevices. For example, a heterostructure with the vertical stacking of MoSe2/PtSe2 was synthesized by Zhou et al. The direct bandgap structure with type-II energy band arrangement between PtSe2 and MoSe2 was confirmed by first-nature principle calculations [28]. He et al. synthesized In2SeS/g-C3N4 and could effectively tune the energy band structure of the system by controlling the stacking pattern of the 2D heterojunction, thus, obtaining various electronic properties [29]. Zhang et al. demonstrated that the MoSe2/WSe2 heterostructure is a type-II vdW using the application of an electric field [30]. So far, there have been many studies based on 2D heterojunctions. However, the problems related to the bandgap in 2D materials are still a hot topic of research. For example, the zero-bandgap property of graphene limits its application in electronic and optoelectronic devices [31]. h-BN has a large bandgap and is an insulator [32]. Researchers have been searching for new 2D materials with moderate bandgaps that are highly advantageous for fabricating electronic and optoelectronic devices, and the tuning of their electronic structure and properties is of great importance. Phosphole is considered as a promising candidate for new electronic devices due to its direct energy gap and high carrier mobility at room temperature [33]. At the same time, low-dimensional tin halide compounds are more stable than 3D tin halide perovskite, because organic cations have better protection for photo active tin halides [34,35].
Considering the significant characteristics of the 2D tin bromide and phosphorene monolayer, we suggested creating a novel heterojunction of P/SnBr2 to enhance their benefits in optoelectronic and solar devices. In this paper, the structure and electronic properties of P/SnBr2 heterojunctions were mainly discussed using the first principle method based on DFT. The results showed that the heterojunction had an atypical type I band arrangement. The calculation results showed that the bandgap of P/SnBr2 heterojunction was −0.08~0.83 eV under the applied electric field. The bandgap varied from −0.015 to 1.04 eV when a biaxial strain was applied, which can be applied to solar energy and photoelectric devices. This suggested that the new heterojunction could be applied to solar devices and optoelectronic devices.

2. The Computational Details

In order to systematically study the electronic properties of P/SnBr2, the first principle method based on the density functional theory (DFT) was adopted in this paper, and the properties of the heterojunction were calculated by the Vienna Ab-initio Simulation Package (VASP) [36,37]. The exchange correlation potential was described by Predew Bourke Ernzerhof (PBE) under the generalized gradient approximation (GGA), which has been proven to be suitable for 2D material heterojunction systems [38,39,40]. The conjugate gradient method was used to optimize the atomic positions and lattice parameters. The convergence criterion of the total energy was set to 10−5 eV, the force convergence criterion for individual atoms was 0.01 eV/Å, and the plane-wave cutoff energy was set to 500 eV. The Brillouin zone was sampled as a whole using a 3 × 11 × 1 grid centered on the gamma point [41], taking into account the effect of the interlayer vdW forces; the DFT-2 was used for correction. The impact of spin–orbit coupling on the heavy atoms was considered throughout the entire calculation process. In the construction of the heterojunction model, a vacuum layer of 20 Å was added along the Z-axis to avoid the influence of periodicity on the structure.

3. Results and Discussion

3.1. The Atomic Structure of P/SnBr2 Heterojunctions

Prior to computing the pertinent characteristics of the heterojunction, we conducted a structural optimization of both the phosphorene and SnBr2 monolayers. After the structural optimization, the lattice parameters of phosphorene were a = 3.30 Å and b = 4.60 Å [42]. The initial configuration of SnBr2 was 1T-phase SnBr2 (space group P3m1) with structurally optimized lattice parameters a = 4.26 Å and b = 4.36 Å. In order to reduce the lattice mismatch during the construction of the heterojunction, we expanded the phosphorene into a 5 × 1 supercell and transformed the initial configuration of SnBr2 from a hexagonal unit into a tetragonal unit, and then expanded it into a 2 × 1 supercell. The lattice mismatch ratio of the heterojunction formed by two monolayers was 5.2%. The lattice parameters of the heterojunction were a = 15.87 Å and b = 4.50 Å. The structure of the heterojunction is shown in Figure 1a.
In order to determine the most stable interlayer distance of the heterojunction, we calculated the binding energy, which was given by:
E b = E P / S n B r 2 E P E S n B r 2 / n
where E P / S n B r 2 , E P and E S n B r 2 represent the total energy of the heterojunction, phosphorene and SnBr2 monolayer, respectively. n indicates the number of P in the unit cell in the heterojunction, and a smaller binding energy indicates a more stable material structure. To ensure that the P/SnBr2 heterojunction had the lowest binding energy, we gradually increased the layer spacing from 2.8 Å to 4.2 Å and calculated their corresponding binding energies separately to obtain the relationship between the binding energy and the layer spacing, as shown in Figure 1b. It can be seen from Figure 1b that the binding energy of the heterojunction system was the lowest at the layer spacing of 3.45 Å. Therefore, the layer spacing was taken as 3.45 Å when the structure was constructed, and this layer spacing was used in the subsequent electronic characteristic analysis in this paper.

3.2. The Electronic Properties of P/SnBr2 Heterojunctions

Based on the most stable heterojunction system, we further calculated the energy band structures of the phosphorene (shown in red) and SnBr2 (shown in blue) monolayers and the projected energy band structures of the heterojunctions, as shown in Figure 2a–c, respectively.
The monolayer phosphorene exhibited direct bandgap semiconductor properties at the high symmetry point gamma with a bandgap of 0.89 eV. The indirect bandgap of the SnBr2 monolayer was 2.13 eV, with the conduction band minimum (CBM) occurring at the gamma point and the valence band maximum (VBM) located between the high symmetry point Y and the gamma point. The P/SnBr2 heterojunction was an indirect bandgap semiconductor with a bandgap of 0.71 eV. In the projected energy band structure of the heterojunction, the red triangle represents the part contributed by phosphorene and the blue square represents the part contributed by SnBr2. It can be seen that its CBM and VBM were located on SnBr2 and phosphorus, respectively. VBM was mainly contributed by phosphorene, while CBM was mainly contributed by a mixture of SnBr2 and a small amount of phosphorene. The total density of states (TDOSs) and partial density of state (PDOS) of the heterojunction system are shown in Figure 2d. It can be observed that the VBM of the P/SnBr2 heterojunction primarily stemmed from the 2s and 2p orbitals of the P atom, whereas the CBM originated from the 5p orbitals of the Sn atom in SnBr2, along with the 2p orbitals of the P atom.
In order to fully understand the electronic structure properties of the P/SnBr2 heterostructure, we performed partial charge density calculations and analysis for the CBM and VBM, as shown in Figure 3a. It can be clearly seen that the VBM came from phosphorene, and the CBM mainly came from SnBr2, with some coming from phosphorene, which is consistent with the conclusion of the energy band structure diagram of the heterojunction.
We also calculated the energy band shifts of the P/SnBr2 heterojunction by using the vacuum energy level as a standard reference point for the different systems. The VBM and CBM band energy alignments were plotted in Figure 3b. It was found that the energy band edges of phosphorene and SnBr2 monolayer were (−0.228, 0.525) eV and (−1.507, 0.577) eV, respectively, based on our calculations. This indicated that the CBM of phosphorene was close to that of SnBr2, while its VBM was higher than that of SnBr2. In addition, we also calculated the conduction band and valence band shifts after the contact of isolated single molecular layers using the core-level alignment method. After constructing the heterojunction system, the energy differences between the VBM and CBM of the two monolayers were 0.0526 eV and 1.279 eV, respectively. This conclusion was consistent with the previous projected energy band diagram of the P/SnBr2 heterojunction, which further indicated the type I band arrangement formed by the P/SnBr2 heterojunction. The calculated work function of the P/SnBr2 heterojunction was 4.73 eV. Our findings indicated that the Fermi energy level of the heterojunction was −0.02 eV. The variance in the energy band shift before and after contact was primarily due to the charge redistribution induced by the formation of the heterojunction. Type I heterojunctions have a unique band-edge arrangement that prevents electron–hole recombination and spontaneously separates free electrons and holes. It can be used in the field of optoelectronics and can achieve efficient photoelectron and solar energy conversion.
The average planar electrostatic potential of the P/SnBr2 heterojunction, normal to the heterostructure (along the Z-direction), is illustrated in Figure 4a. The outcomes demonstrated that the potential of the phosphorene monolayer was more profound compared to that of the SnBr2 monolayer. The heterojunction bilayer exhibited an electric potential difference of 13.18 eV, which implied the existence of an electrostatic field perpendicular to the interface direction. Moreover, due to the higher electron negativity of phosphorene compared to the SnBr2 monolayer, the former has a more profound electrostatic potential compared to the latter. Therefore, when the P/SnBr2 heterostructure was formed, electrons were spontaneously transferred from the SnBr2 layer to the phosphorene layer, which generated a strong built-in electric field through the heterostructure. The inherent electric field generated in the heterostructure enhanced the segregation efficiency of the photogenerated charge carriers and reduced their recombination, leading to the development of a P/SnBr2 heterostructure that exhibited excellent optoelectronic properties. In order to display the charge transfer of heterostructures more clearly, we calculated the average electron density difference in the heterostructures along the Z-direction. As shown in Figure 4b, the yellow part represented the aggregation of electrons and the cyan part represented the dissipation of electrons; the positive values represented the gain of electrons and the negative values represented the loss of electrons. We can see that after the formation of the heterojunction, the charges of the two layers were redistributed, with the phosphorene layer losing electrons and the SnBr2 monolayer gaining electrons. In order to quantitatively estimate the electron transfer, we performed a Bader charge analysis. The amount of charge transferred from the phosphorene monolayer to SnBr2 monolayer was 0.149 e.

3.3. The Regulation of the Electronic Structure Properties of P/SnBr2 Heterojunctions by Electric Fields

The application of external electric fields has been shown to be an effective strategy for regulating the electronic structure of heterojunctions [43]. Thus, we conducted a systematic investigation of the impact of the vertical electric fields on the energy band structure of the P/SnBr2 heterojunction. The positive direction of the electric field was designated as the direction from the phosphorene layer to the SnBr2 layer along the Z-axis. Figure 5a shows the interaction between the P/SnBr2 heterojunction and the external vertical electric field. Here, the applied vertical electric field ranged from −0.8/Å to 0.8 V/Å. With the increase in the positive electric field, the bandgap of the heterojunction gradually decreased, while with the increase in the negative electric field, the bandgap correspondingly increased. It showed that the bandgap of the heterojunction was positively related with the external electric field; the bandgap values varied between −0.08 and 0.82 eV. The heterojunction was metallic when the electric field was greater than 0.8 V/Å. When the electric field was less than −0.4 V/Å, the modulation of the bandgap by the electric field became insignificant and the electric field no longer played an important role in regulating the bandgap.
In order to better understand the effect of the electric field on the interfacial properties of the P/SnBr2 heterostructures, we calculated the difference in the average planar charge density of the heterostructures along the vertical z-axis under different positive and negative electric fields. From Figure 5b, it can be clearly seen that the electron transfer increased rapidly when a negative electric field was applied, while the electron transfer from phosphorene to SnBr2 increased gradually under a positive electric field as the electric field increased to 0.7 V/Å. Our findings demonstrated that the intensity and orientation of the electric field significantly influenced the electronic and interfacial properties of the P/SnBr2 heterostructures. This effect could efficiently adjust the energy band structure and interfacial interactions, thereby, enhancing their potential applications in nanoscale optoelectronic devices.

3.4. Effect and Regulation of Biaxial Strain on P/SnBr2 Heterojunctions

During the formation of heterostructures, strains are difficult to remove due to the influence of external forces or substrates. Meanwhile, mechanical strain has been proved to be an effective method for regulating the electronic properties in 2D applications [44,45]. Therefore, we studied the electronic properties of the P/SnBr2 heterosjunctions under the biaxial strain conditions of −8%~8%.
To simulate the mechanical strain, we modified the lattice constants along the x- and y-directions, followed by a full optimization of atomic positions based on these changes. The strain was defined as ε = (L−L0)/L0, wherein L and L0 were the values of the lattice constants in the strained and original states, respectively. The positive and negative values of strain applied to the P/SnBr2 heterojunction (ε > 0) indicated the tensile and compressive strains applied to the system, respectively. We changed the biaxial strain applied on the P/SnBr2 heterojunction from −8% to +8%, with a step of 2%. The evolution of the energy band structure of the P/SnBr2 heterojunction is shown in Figure 6. The projected energy band diagrams for applying −8%, −6%, 4% and 6% strains are shown in Figure 7a–d, respectively.
Figure 6 and Figure 7 show that the bandgap of the P/SnBr2 heterojunction decreased from 0.71 eV to 0.53 eV (when the strain was −2%) under the effect of compressive strain. With the further increase in compressive strain, when the strain was −6%, the CBM value was gradually lower than the VBM value, and the bandgap of the heterojunction was 0.06 eV. When the strain was -8%, the value of the CBM was lower than that of the VBM. Since the CBM and VBM come from P and SnBr2, respectively, the energy band arrangement of the heterojunction underwent a type I to type III transition when the strain was less than −8%. Then, the bandgap of the heterojunction was closed and the system appeared metallic.
Under the tensile strain, the CBM shifted slightly upward and the VBM shifted slightly downward when the tensile strain increased to 2%, resulting in an increase in the bandgap of the P/SnBr2 heterojunction from 0.71 eV to 0.89 eV. With the tensile strain increased to 4%, the bandgap of the P/SnBr2 heterojunction reached a maximum of 1.04 eV. After that, with the increase in the tensile strain, the bandgap of the heterojunction decreased gradually, but the value of the bandgap was still greater than the initial bandgap without strain. Meanwhile, we observed that at 6% and 8% tensile strain, the VBM shifted from a position between Г and X to the Г point, leading to a transition in the P/SnBr2 heterojunction from an indirect bandgap to a direct bandgap.
It can be inferred that the application of biaxial strains is an effective means to regulate the energy bandgap and energy band alignment of the heterostructures. There exists a range of linear regulation of biaxial strain on the bandgap of P/SnBr2 heterojunctions; the strain range was −8%~4%, and the bandgap changed from metallicity to 1.04 eV. The transition from indirect to direct bandgap could be achieved at 6% and 8% strains.

4. Conclusions

In summary, the paper presented a systematic study of heterojunctions composed of phosphorene and SnBr2 using the DFT method. We found that the external electric field and biaxial strain could effectively regulate the electronic properties of the P/SnBr2 heterojunctions. In a certain range of electric field and strain regulation, the bandgap changed linearly. When the biaxial strain increased to −8% or the electric field strength was greater than 0.8 V/Å, the heterojunction became a metallic system. When the compression strain was greater than 8%, the band alignment shifted from type I to type III. An indirect to direct band gap transition occurred when the tensile strain was greater than 6%. The energy band alignment of the heterojunction was type I. Such structures allows the separation of electrons and holes in real space. Thus, the recombination of carriers is reduced, which is very beneficial to the application of materials in solar cells. The results of the paper could provide certain valuable parameters and guidance for the research and related experiments in the application of SnBr2-based heterojunctions for the manufacture of solar cells and optoelectronic devices.

Author Contributions

Conceptualization, K.Y. and Z.D.; methodology, K.Y. and Z.D.; software, K.Y. and Z.D.; validation, K.Y. and Z.D.; formal analysis, K.Y., J.S. and Z.D.; investigation, K.Y., Z.D. and Q.H.; resources, K.Y. and Z.D.; data curation, K.Y. and Z.D.; writing—original draft preparation, K.Y. and Z.D.; writing—review and editing, K.Y. and Z.D.; visualization, K.Y. and Z.D.; supervision, K.Y. and Z.D.; project administration, K.Y., Q.L. and Z.D.; funding acquisition, Z.D., J.S. and Q.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key Project of the Foundation of Anhui Education Committee, China (2022AH050066, KJ2019A0020), the National Natural Science Foundation of China (21773001), Anhui Project (Z010118169), the Natural Science Foundation of Anhui Province (2108085MA16).

Data Availability Statement

Not applicable.

Acknowledgments

Computational resources from the Shanghai Supercomputer Center and Hefei Advanced Computing Center are acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  2. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Yang, R.; Fan, J.; Sun, M. Transition metal dichalcogenides (TMDCs) heterostructures: Optoelectric properties. Front. Phys. 2022, 17, 43202. [Google Scholar] [CrossRef]
  4. Cianci, S.; Blundo, E.; Felici, M.; Polimeni, A.; Pettinari, G. Tailoring the optical properties of 2D transition metal dichalcogenides by strain. Opt. Mater. 2022, 125, 112087. [Google Scholar] [CrossRef]
  5. Ashhadi, M.; Hadavi, M.S.; Sarri, Z. Electronic transport properties and first-principles study of graphene/h-BN and h-BN bilayers. Phys. E Low Dimens. Syst. Nanostruct. 2017, 87, 312–316. [Google Scholar] [CrossRef]
  6. Brent, J.R.; Savjani, N.; Lewis, E.A.; Haigh, S.J.; Lewis, D.J.; O’Brien, P. Production of few-layer phosphorene by liquid exfoliation of black phosphorus. Chem. Commun. 2014, 50, 13338–13341. [Google Scholar] [CrossRef] [Green Version]
  7. Carvalho, A.; Wang, M.; Zhu, X.; Rodin, A.S.; Su, H.; Neto, A.H.C. Phosphorene: From theory to applications. Nat. Rev. Mater. 2016, 1, 16061. [Google Scholar] [CrossRef]
  8. Novoselov, K.S.; Mishchenko, A.; Carvalho, A.; Neto, A.H.C. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Chen, Y.; Wang, X.; Wu, G.; Wang, Z.; Fang, H.; Lin, T.; Sun, S.; Shen, H.; Hu, W.; Wang, J.; et al. High-performance photovoltaic detector based on MoTe2/MoS2 van der Waals heterostructure. Small 2018, 14, 1703293. [Google Scholar] [CrossRef]
  10. Yang, Y.; Wang, Z. A two-dimensional MoS2/C3N broken-gap heterostructure, a first principles study. RSC Adv. 2019, 9, 19837–19843. [Google Scholar] [CrossRef]
  11. Cui, X.; Troadec, C.; Wee, A.T.S.; Huang, Y.L. Surface nanostructure formation and atomic-scale templates for nanodevices. ACS Omega 2018, 3, 3285–3293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zhao, K.; He, D.; Fu, S.; Bai, Z.; Miao, Q.; Huang, M.; Wang, Y.; Zhang, X. Interfacial Coupling and Modulation of van der Waals Heterostructures for Nanodevices. Nanomaterials 2022, 12, 3418. [Google Scholar] [CrossRef]
  13. Zhang, J.L.; Han, C.; Hu, Z.; Wang, L.; Liu, L.; Wee, A.T.; Chen, W. 2D phosphorene: Epitaxial growth and interface engineering for electronic devices. Adv. Mater. 2018, 30, 1802207. [Google Scholar] [CrossRef] [PubMed]
  14. Tapia, M.A.; Gusmão, R.; Serrano, N.; Sofer, Z.; Ariño, C.; Díaz-Cruz, J.M.; Esteban, M. Phosphorene and other layered pnictogens as a new source of 2D materials for electrochemical sensors. TrAC Trends Anal. Chem. 2021, 139, 116249. [Google Scholar] [CrossRef]
  15. Zhang, W.; Li, F.; Hu, J.; Zhang, P.; Yin, J.; Tang, X.; Jiang, Y.; Wu, B.; Ding, Y. Strain engineering on transmission carriers of monolayer phosphorene. J. Phys. Condens. Matter 2017, 29, 465501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yang, J.; Lu, Y. Optical properties of phosphorene. Chin. Phys. B 2017, 26, 034201. [Google Scholar] [CrossRef]
  17. Liang, J.; Hu, Y.; Zhang, K.; Wang, Y.; Song, X.; Tao, A.; Liu, Y.; Jin, Z. 2D layered black arsenic-phosphorus materials: Synthesis, properties, and device applications. Nano Res. 2022, 15, 3737–3752. [Google Scholar] [CrossRef]
  18. Jiang, Z.T.; Lv, Z.T.; Zhang, X.D. Electromechanical field effect transistors based on multilayer phosphorene nanoribbons. Phys. Lett. A 2017, 381, 1962–1966. [Google Scholar] [CrossRef]
  19. Luo, Y.; Xie, Y.; Ye, X.; Wang, Y. A self-powered phosphorene photodetector with excellent spin-filtering and spin-valve effects. Phys. Chem. Chem. Phys. 2019, 21, 7613–7617. [Google Scholar] [CrossRef]
  20. Li, J.; Guo, C.; Li, C.M. Recent Advances of Two-Dimensional (2D) MXenes and Phosphorene for High-Performance Rechargeable Batteries. ChemSusChem 2020, 13, 1047–1070. [Google Scholar] [CrossRef]
  21. Hao, F.; Stoumpos, C.C.; Cao, D.H.; Chang, R.P.H.; Kanatzidis, M.G. Lead-free solid-state organic–inorganic halide perovskite solar cells. Nat. Photonics 2014, 8, 489–494. [Google Scholar] [CrossRef]
  22. Fu, P.; Huang, M.; Shang, Y.; Yu, N.; Zhou, H.-L.; Zhang, Y.-B.; Chen, S.; Gong, J.; Ning, Z. Organic-Inorganic layered and hollow tin bromide perovskite with tunable broadband emission. ACS Appl. Mater. Interfaces 2018, 10, 34363–34369. [Google Scholar] [CrossRef] [PubMed]
  23. Dorrell, W.; Pirie, H.; Gardezi, S.M.; Drucker, N.C.; Hoffman, J.E. van der Waals metamaterials. Phys. Rev. B 2020, 101, 121103. [Google Scholar] [CrossRef] [Green Version]
  24. Wang, Q.; Zhang, L.; Liu, X.; Li, S. Two-Dimensional Semiconductor Heterojunctions for Optoelectronics and Electronics. Front. Energy Res. 2021, 9, 802055. [Google Scholar] [CrossRef]
  25. Di Bartolomeo, A. Emerging 2D materials and their van der Waals heterostructures. Nanomaterials 2020, 3, 579–588. [Google Scholar] [CrossRef] [Green Version]
  26. Song, D.Y.; Chu, D.; Lee, S.K.; Pak, S.W.; Kim, E.K. High photoresponsivity from multilayer MoS2/Si heterojunction diodes formed by vertically stacking. J. Appl. Phys. 2017, 122, 124505. [Google Scholar] [CrossRef]
  27. Zhu, W.; Xue, Z.; Wang, G.; Zhao, M.; Chen, D.; Guo, Q.; Liu, Z.; Feng, X.; Ding, G.; Chu, P.K.; et al. Graphene quantum dot-decorated vertically oriented graphene/germanium heterojunctions for near-infrared photodetectors. ACS Appl. Nano Mater. 2020, 3, 6915–6924. [Google Scholar] [CrossRef]
  28. Zhou, J.; Kong, X.; Sekhar, M.C.; Lin, J.; Le Goualher, F.; Xu, R.; Wang, X.; Chen, Y.; Zhou, Y.; Zhu, C.; et al. Epitaxial synthesis of monolayer PtSe2 single crystal on MoSe2 with strong interlayer coupling. ACS Nano 2019, 13, 10929–10938. [Google Scholar] [CrossRef]
  29. He, C.; Han, F.S.; Zhang, J.H.; Zhang, W.X. The In 2 SeS/gC3N4 heterostructure: A new two-dimensional material for photocatalytic water splitting. J. Mater. Chem. C 2020, 8, 6923–6930. [Google Scholar] [CrossRef]
  30. Zhang, F.; Li, W.; Dai, X. Modulation of electronic structures of MoSe2/WSe2 van der Waals heterostructure by external electric field. Solid State Commun. 2017, 266, 11–15. [Google Scholar] [CrossRef]
  31. Barhoumi, M.; Lazaar, K.; Said, M. Electronic and vibrational properties of TMDs heterogeneous bilayers, nontwisted bilayers silicene/TMDs heterostructures and photovoltaic heterojunctions of fullerenes with TMDs monolayers. Phys. E Low Dimens. Syst. Nanostruct. 2018, 104, 155–164. [Google Scholar] [CrossRef]
  32. Cassabois, G.; Valvin, P.; Gil, B. Hexagonal boron nitride is an indirect bandgap semiconductor. Nat. Photonics 2016, 10, 262–266. [Google Scholar] [CrossRef] [Green Version]
  33. Liu, H.; Neal, A.T.; Zhu, Z.; Luo, Z.; Xu, X.; Tomanek, D.; Ye, P.D. Phosphorene: An unexplored 2D semiconductor with a high hole mobility. ACS Nano 2014, 8, 4033–4041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zhou, C.; Tian, Y.; Wang, M.; Rose, A.; Besara, T.; Doyle, N.K.; Yuan, Z.; Wang, J.C.; Clark, R.; Hu, Y.; et al. Low-dimensional organic tin bromide perovskites and their photoinduced structural transformation. Angew. Chem. Int. Ed. 2017, 56, 9018–9022. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, Y.; Jiang, K.-J.; Wang, P.; Gu, W.-M.; Yu, G.-H.; Zhou, X.; Song, Y. Highly oriented quasi-2D layered tin halide perovskites with 2-thiopheneethylammonium iodide for efficient and stable tin perovskite solar cells. New J. Chem. 2022, 46, 2259–2265. [Google Scholar] [CrossRef]
  36. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  37. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  38. Ma, Y.; Zhao, X.; Niu, M.; Dai, X.; Li, W.; Wang, X.; Zhao, M.; Wang, T.; Tang, Y. Modulation of interfacial electronic properties in PbI2 and BN van der Waals heterobilayer via external electric field. Appl. Surf. Sci. 2017, 411, 46–52. [Google Scholar] [CrossRef]
  39. Su, J.; Liu, H.; Jia, Z. Electric field tunable electronic properties of antimonene/graphyne van der Waals heterostructure. J. Alloys Compd. 2022, 909, 164653. [Google Scholar] [CrossRef]
  40. Sharma, N.K.; Sahoo, S.; Sahu, M.C.; Mallik, S.K.; Jena, A.K.; Sharma, H.; Gupta, S.K.; Ahuja, R.; Sahoo, S. Electronic bandstructure modulation of MoX2/ZnO (X: S, Se) heterostructure by applying external electric field. Surf. Interfaces 2022, 29, 101817. [Google Scholar] [CrossRef]
  41. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  42. Wei, Y.; Wang, F.; Zhang, W.; Zhang, X. The electric field modulation of electronic properties in a type-II phosphorene/PbI2 van der Waals heterojunction. Phys. Chem. Chem. Phys. 2019, 21, 7765–7772. [Google Scholar] [CrossRef] [PubMed]
  43. Baskoutas, S.; Paspalakis, E.; Terzis, A.F. Electronic structure and nonlinear optical rectification in a quantum dot: Effects of impurities and external electric field. J. Phys. Condens. Matter 2007, 19, 395024. [Google Scholar] [CrossRef]
  44. Wang, Y.; Wang, B.; Huang, R.; Gao, B.; Kong, F.; Zhang, Q. First-principles study of transition-metal atoms adsorption on MoS2 monolayer. Phys. E Low Dimens. Syst. Nanostruct. 2014, 63, 276–282. [Google Scholar] [CrossRef]
  45. Li, X.H.; Wang, B.J.; Cai, X.L.; Zhang, L.W.; Wang, G.D.; Ke, S.H. Tunable electronic properties of arsenene/GaS van der Waals heterostructures. RSC Adv. 2017, 7, 28393–28398. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) The front view of P/SnBr2 heterojunction; (b) curve of system energy changing with layer spacing.
Figure 1. (a) The front view of P/SnBr2 heterojunction; (b) curve of system energy changing with layer spacing.
Crystals 13 01077 g001
Figure 2. (a) Band structures of phosphorene; (b) band structures of SnBr2; (c) projected band structures of the P/SnBr2; (d) state density diagram of the heterojunction.
Figure 2. (a) Band structures of phosphorene; (b) band structures of SnBr2; (c) projected band structures of the P/SnBr2; (d) state density diagram of the heterojunction.
Crystals 13 01077 g002
Figure 3. (a) The real-space charge densities of the VBM and CBMfor P/SnBr2; (b) the energy band alignment of the P/SnBr2 heterojunction.
Figure 3. (a) The real-space charge densities of the VBM and CBMfor P/SnBr2; (b) the energy band alignment of the P/SnBr2 heterojunction.
Crystals 13 01077 g003
Figure 4. (a) Average electrostatic potential distribution of P/SnBr2 heterojunction in the z-direction; (b) differential electron density of P/SnBr2 along the Z direction.
Figure 4. (a) Average electrostatic potential distribution of P/SnBr2 heterojunction in the z-direction; (b) differential electron density of P/SnBr2 along the Z direction.
Crystals 13 01077 g004aCrystals 13 01077 g004b
Figure 5. (a) the curve illustrating the change in bandgap for the P/SnBr2 heterojunction under distinct electric fields; (b) the average charge density difference across the plane of the P/SnBr2 heterostructure in the z-direction under diverse electric fields.
Figure 5. (a) the curve illustrating the change in bandgap for the P/SnBr2 heterojunction under distinct electric fields; (b) the average charge density difference across the plane of the P/SnBr2 heterostructure in the z-direction under diverse electric fields.
Crystals 13 01077 g005aCrystals 13 01077 g005b
Figure 6. The curve depicts the alteration of the bandgap for the P/SnBr2 heterojunction under biaxial strains.
Figure 6. The curve depicts the alteration of the bandgap for the P/SnBr2 heterojunction under biaxial strains.
Crystals 13 01077 g006
Figure 7. Projected band structure diagrams of the heterojunction at a −8% strain (a), at a −6% strain (b), at a 4% strain (c) and at a 6% strain (d).
Figure 7. Projected band structure diagrams of the heterojunction at a −8% strain (a), at a −6% strain (b), at a 4% strain (c) and at a 6% strain (d).
Crystals 13 01077 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, K.; Ding, Z.; Hu, Q.; Sun, J.; Li, Q. A Theoretical Investigation of the Structural and Electronic Properties of P/SnBr2 Heterojunctions. Crystals 2023, 13, 1077. https://doi.org/10.3390/cryst13071077

AMA Style

Yang K, Ding Z, Hu Q, Sun J, Li Q. A Theoretical Investigation of the Structural and Electronic Properties of P/SnBr2 Heterojunctions. Crystals. 2023; 13(7):1077. https://doi.org/10.3390/cryst13071077

Chicago/Turabian Style

Yang, Kun, Zongling Ding, Qi Hu, Jin Sun, and Qiuju Li. 2023. "A Theoretical Investigation of the Structural and Electronic Properties of P/SnBr2 Heterojunctions" Crystals 13, no. 7: 1077. https://doi.org/10.3390/cryst13071077

APA Style

Yang, K., Ding, Z., Hu, Q., Sun, J., & Li, Q. (2023). A Theoretical Investigation of the Structural and Electronic Properties of P/SnBr2 Heterojunctions. Crystals, 13(7), 1077. https://doi.org/10.3390/cryst13071077

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop