Next Article in Journal
In Situ Formation of Surface-Induced Oxygen Vacancies in Co9S8/CoO/NC as a Bifunctional Electrocatalyst for Improved Oxygen and Hydrogen Evolution Reactions
Next Article in Special Issue
The Low-Temperature Photocurrent Spectrum of Monolayer MoSe2: Excitonic Features and Gate Voltage Dependence
Previous Article in Journal
Using Gold-Nanorod-Filled Mesoporous Silica Nanobeads for Enhanced Radiotherapy of Oral Squamous Carcinoma
Previous Article in Special Issue
Photosensing and Characterizing of the Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets Fabricated by Thermal V–S Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic and Optical Properties of Atomic-Scale Heterostructure Based on MXene and MN (M = Al, Ga): A DFT Investigation

1
School of Mechanical and Electronic Engineering, Nanjing Forestry University, Nanjing 210037, China
2
School of Materials Science and Engineering, Southeast University, Nanjing 211189, China
3
Center for More-Electric-Aircraft Power System, Nanjing University of Aeronautics and Astronautics, Nanjing 211106, China
*
Authors to whom correspondence should be addressed.
Submission received: 7 July 2021 / Revised: 16 August 2021 / Accepted: 26 August 2021 / Published: 30 August 2021
(This article belongs to the Special Issue Optoelectronic Properties and Applications of Nanomaterials)

Abstract

:
After the discovery of graphene, a lot of research has been conducted on two-dimensional (2D) materials. In order to increase the performance of 2D materials and expand their applications, two different layered materials are usually combined by van der Waals (vdW) interactions to form a heterostructure. In this work, based on first-principles calculation, some charming properties of the heterostructure constructed by Hf2CO2, AlN and GaN are addressed. The results show that Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures can keep their original band structure shape and have strong thermal stability at 300 K. In addition, the Hf2CO2/MN heterostructure has I-type band alignment structure, which can be used as a promising light-emitting device material. The charge transfer between the Hf2CO2 and AlN (or GaN) monolayers is 0.1513 (or 0.0414) |e|. The potential of Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures decreases by 6.445 eV and 3.752 eV, respectively, across the interface. Furthermore, both Hf2CO2/AlN and Hf2CO2/GaN heterostructures have remarkable optical absorption capacity, which further shows the application prospect of the Hf2CO2/MN heterostructure. The study of this work provides theoretical guidance for the design of heterostructures for use as photocatalytic and photovoltaic devices.

Graphical Abstract

1. Introduction

Since 2004, Novoselov and Geim prepared graphene from graphite by the mechanical exfoliation method [1], and its remarkable physical and chemical properties were explored [2,3,4,5,6,7,8,9,10,11], which also attracted extensive interest and attention on other two-dimensional (2D) materials, and they all show fantastic properties [12,13,14,15,16,17]. For example, black phosphorene is a honeycomb-like folded layered material that can achieve transistor performance with a thickness of less than 7.5 nm, and the highest carrier mobility can be obtained by 1000 cm2/V·s when the thickness is 10 nm at room temperature [18,19,20,21]. Puckered arsenene possesses the ability to adjust its bandgap by applying the external strain on its surface. Interestingly, arsenene can even be transformed into a straight gap semiconductor by applying 1% strain [22,23,24,25,26,27]. Transition metal dichalcogenides (TMDs) materials are layered materials with excellent thermal [28,29], electronic [30] and optical properties [31]. For instance, MoS2 has high broadband gain (up to 13.3), detection rate (up to 1010 cm Hz1/2/W) and high thermal stability when using it as an optoelectronic device [32]. In addition, there are Janus TMDs materials that destroy the symmetry of the original structure and make its carrier mobility increase from 28 to 606 cm2/V·s [33]. The Janus MoSSe material is able to separate light-generated electrons and holes while also exhibiting perfect light absorption capabilities, which lays the foundation for promoting water redox reactions, and it is a remarkable water decomposition light catalyst [34]. The novel properties of these 2D materials can provide unprecedented help for the development of nano-devices and solar cells [35,36].
In order to expand the application of 2D materials and build more special performances, superposing different layered materials to construct a heterostructure is usually realized [37,38,39,40,41,42,43]. Especially, two-layer materials construct a heterostructure by van der Waals (vdW) forces, which can induce novel interfacial [44], optical [45] and electronic properties [46]. The SiC/TMDs vdW can be used as a water decomposition catalyst to completely separate hydrogen and oxygen under the condition of light [47]. The PbSe/CdSe heterostructure has near infrared emission characteristics, which is closely related to its type-I alignment band [48]. The average carrier value of the type-I PbI2/WS2 layered heterostructure is 0.039 cm2·s−1, and it is found that the interlayer diffusion behavior between electrons and holes is similar [49]. These investigations have demonstrated that type-I heterostructures possess promising applications in photocatalytic, photovoltaic and optical devices [50,51]. Recently, layered MAX phases have been exfoliated into monolayer and multilayers, named MXenes, which has attracted wide attention [52]. The charming electrochemical [53], conductive [54] and stable capacity [55] characteristics provide potential applications in electrocatalysts, photocatalysts and energy storage devices [56,57,58]. Although most of the MXenes are metallic, some MXenes are semiconductors with a desirable bandgap [59,60]. In particular, the Cr2TiC2 monolayer behaves as a novel bipolar antiferromagnetic semiconductor showing opposite spin directions, which can be used as antiferromagnetic spin field effect transistor [61]. Hf2CO2 possesses excellent electronic and thermoelectric properties and carrier mobility (about 1531.48 cm2/V·s for electrons), suggesting an efficient photocatalyst for water splitting and nano-electronic devices [62,63,64,65,66], and this novel electronic characteristic can even be tuned by external strain [67]. Hf2CO2 is also sensitive to NH3, which can sharply enhance electronic conductivity [68]. More recently, MN (M = Al, Ga) has been reported to have remarkable optical, electronic and mechanical properties, and it can be considered as a candidate for future optical and photovoltaic devices [69,70,71,72,73,74,75]. Interestingly, the prepared 2D AlN shows great promise in deep-ultraviolet optoelectronic applications, ultraviolet LEDs and laser diodes [76,77]. In addition, 2D GaN is also fabricated by epitaxial graphene using a migration-enhanced encapsulated growth method [75], and is studied as a decent semiconductor for heterostructures [78], photocatalysts [79] and photocathodes [80]. Moreover, some Hf2CO2 and MN-based heterostructures have been reported, such as Hf2CO2/WS2 [62], Hf2CO2/blue phosphorene [65], MoS2/MN [81], GeC/GaN [82], etc., while studies of the Hf2CO2/MN heterostructure are still limited. Therefore, considering such fantastic electronic properties of Hf2CO2, and the synthesized MN (M = Al, Ga), it is worth constructing the heterostructure by Hf2CO2 and MN to explore the novel performances and the potential applications.
In this study, the first-principles method was utilized to investigate the formed heterostructure based on Hf2CO2 and MN (M = Al, Ga). The binding energy and the ab initio molecular dynamics (AIMD) calculations were conducted to check the stability of the heterostructure. Furthermore, the type-I band alignment of the Hf2CO2/MN heterostructure was addressed, which demonstrates the potential applications of the light-emitting devices. In addition, the charge difference between the MXene and the MX layers was studied and the potential drop was also calculated to develop the interfacial properties of the heterostructure. Moreover, the light absorption capacity of the Hf2CO2/MN heterostructure was obtained by calculating the intrinsic optical absorption spectrum.

2. Computing Method

The method of calculations in this work was based on density functional theory (DFT), implemented by first-principles simulation under the circumstances of the Vienna ab initio simulation package (VASP) [83]. Based on generalized gradient approximation (GGA), the Perdew–Burke–Ernzerhof (PBE) functional was employed for the explanation of the exchange correlation functional [84,85,86]. For the more precise bandgap results, we used the hybrid Heyd–Scuseria–Ernzerhof (HSE06; screening parameter 0.2 Å−1, mixing parameter 0.25) simulations [87] and the DFT-D3 method of Grimme, and the dipole corrections were also used to correct the weak dispersion forces. A tested 500 eV cutoff energy was considered. After the convergence test for the k-point (seen in Table S1 of Supplementary Materials), the Monkhorst–Pack k-point of 7 × 7 × 1 was adopted to relax the structure, while the static and optical calculations were conducted by the 11 × 11 × 1 k-point. For the prevention of the interaction of the adjacent atomic layers, the vacuum slab was controlled by 25 Å. Furthermore, the energy of the calculated materials in this work was set within 1 × 105 eV, while the Hellmann–Feynman forces on the atoms were set to less than 0.01 eV·Å1.

3. Results and Discussions

We first optimized the structures of Hf2CO2, AlN and GaN, and the top and side views of the crystal structure and band energy for the Hf2CO2, AlN and GaN monolayers are shown in Figure 1a–c, respectively. The lattice constants of monolayered Hf2CO2 AlN and GaN are obtained by 3.363, 3.127 and 3.255 Å, respectively. In addition, the bond lengths of Hf−C, Hf−O, Al−N and Ga−N in Hf2CO2, AlN and GaN monolayers are 2.369, 2.132, 1.805 and 1.895 Å, respectively. In addition, the HSE06 method-calculated energy band structures of Hf2CO2, AlN and GaN monolayers show that all these layered materials have semiconductor features with the bandgaps of 1.820, 4.042 and 3.203 eV, respectively. For the Hf2CO2 monolayer, the conduction band minimum (CBM) is located at the M point, while the valence band maximum (VBM) appears at the Γ point. The CBM and VBM of AlN (or GaN) are generated at the Γ point and K point, respectively. All these calculated results are almost the same as the previous investigation results [65,66,88].
To construct the heterostructure by Hf2CO2 and MN (N = Al, Ga) monolayers, the six most representative stacking configurations, shown in Figure 2, should be taken into consideration. In the six Hf2CO2/MN heterostructures, the binding energy (Ebinding) of the Hf2CO2/MN heterostructure is decided by:
E binding = E MXene / MN E MXene E MN
where EMXene/MN, EMXene and EMN show the total energy of the Hf2CO2/MN heterostructure, monolayered Hf2CO2 and MN, respectively. The smaller the binding energy, the more stable the structure of the heterostructure [89], and thus the most stable structure of those six stacking configurations of the heterostructure is decided as the lowest binding energy, which is demonstrated in the AA stacking style of the Hf2CO2/MN heterostructure. The calculated Ebinding of the Hf2CO2/AlN and Hf2CO2/GaN heterostructures with the most stable configuration is −56.98 and −52.44 meV/Å2, respectively, in Table 1. It is worth noting that the framework of the quantum theory of atoms in molecules (QTAIM) functional is not considered here, which is also a popular method for simulations [90,91,92,93,94,95,96,97,98,99]. The results show that the Hf2CO2/MN heterostructure is formed by vdW interactions [100,101]. For the most stable Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures, the bond lengths of Hf−C, Hf−O and M−N are slightly changed compared with original layered materials, which further proves the weak vdW forces between the interface of the heterostructures. In addition, the interface distances of the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures are 1.924 and 2.235 Å, respectively. Additionally, in the following sections, we only discuss the most stable Hf2CO2/MN heterostructure stacking structure.
In order to further investigate the thermal stability of the Hf2CO2/MN vdW heterostructure, AIMD simulations were explored for the Hf2CO2/MN vdW heterostructure by the Nosé–Hoover heat bath scheme [102]. To consider the constraints of the lattice translation, we constructed a 6 × 6 × 1 supercell for the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures in the AIMD simulation, which contained 252 atoms in total. The ambient temperature of the simulation was set as 300 K, and the structures of the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures after relaxation for 5 ps are shown in Figure 3a,c, respectively. The simulation results of AIMD show that the structures of the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures still remain intact after 5 ps under 300 K, revealing the robust thermal stability of the heterostructure. In addition, as shown in Figure 3b,d, the total energy fluctuation and simulation time for Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures in AIMD calculation are demonstrated, respectively, and they all show the convergence state by time, ensuring the reliability of the results.
Figure 4a,c show the projected band structure of the MXene/MN vdW heterostructure. It is obvious that Hf2CO2/AlN and Hf2CO2/GaN heterostructures have an indirect bandgap of 2.006 eV and 1.899 eV, respectively. The gray and red marks represent the band contribution of AlN (or GaN) and Hf2CO2 layers, respectively. Therefore, we can see that the CBM and VBM of both MXene/MN vdW heterostructures are donated from the Hf2CO2 layer, and the MXene/MN vdW heterostructure shows I-type band structure. In addition, we also investigated the partial density calculation of MXene/MN vdW heterostructures, as shown in Figure 4b,d, which further proves the intrinsic type-I band structure characteristic. The band structures of the MXene/AlN and MXene/GaN vdW heterostructures by all six different stacking configurations are calculated in Figures S1 and S2, respectively, in the Supplementary Materials.
In the Hf2CO2/MN vdW heterostructure, the bandgap of the Hf2CO2 layer is smaller than that of the AlN (or GaN) layer. Additionally, the CBM and VBM of the MXene/MN vdW heterostructure are fixed in the band gap of the Hf2CO2 layer, as shown in Figure 5. When some external conditions are applied, the electrons in the wide bandgap of MN will be excited and move to the CBM of MN. At the same time, holes will be induced in the VBM of MN. With the help of conduction band shift (CBO) and valence band shift (VBO), the electrons and holes are excited from the MN layer to the Hf2CO2 layer at CBM and VBM, respectively, as shown in Figure 5a. The CBO and VBO of the Hf2CO2/AlN (or Hf2CO2/GaN) vdW heterostructure are obtained as 2.496 eV (or 0.432eV) and 1.744 eV (or 0.383 eV), respectively. Due to the lower energy, the electrons and holes excited in the Hf2CO2 narrow bandgap are prevented from transferring to the MN layer, as shown in Figure 5b [50], suggesting the potential usage of the light-emitting device.
The interesting properties of the interface for the MXene/MN heterostructure were induced by vdW forces, such as the charge difference density (Δρ), which is evaluated by:
Δ ρ = ρ MXene / MN ρ MXene ρ MN
where ρMXene/MN, ρMXene and ρMN show the total charge density of the Hf2CO2/AlN (or Hf2CO2/GaN) vdW heterostructure, monolayered Hf2CO2 and AlN (or GaN), respectively. Figure 6a,b show the difference in charge density between the interface in Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures. It is obvious that AlN (or GaN) acts as an electron donor in contact with Hf2CO2. In addition, the charge density is redistributed in Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures, which contributes to the formation of electron-rich and hole-rich regions. It is found that there is charge transfer between the two monolayers. The electron transfer of 0.1513 (or 0.0414) |e| in the Hf2CO2/AlN (or Hf2CO2/GaN) vdW heterostructure is calculated by the Bader-charge analysis method [103]. In addition, we observed that when Hf2CO2 and MN come into contact and reach the equilibrium position, the potential across the interface of the Hf2CO2/MN vdW heterostructure decreases to varying degrees due to the charge transfer. The potential of Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures decreases by 6.445 eV and 3.752 eV, respectively, which can also be used as an effective driving force to promote carriers in Figure 5. The charge density differences of the MXene/AlN and MXene/GaN vdW heterostructures compared to the other five stacking configurations are obtained by Figures S3 and S4, respectively, in the Supplementary Materials. It is worth noting that the AlN (or GaN) layer still acts as an electron donor for the Hf2CO2 layer in the other five stacking configuration heterostructures, and the transferred electrons are calculated in Table S1 in the Supplementary Materials.
The optical absorption capacity of the Hf2CO2/MN vdW heterostructure was also investigated. The light absorption capacity of the MXene, MN and the MXene/MN vdW heterostructure is obtained in Figure 7 by the optical absorption spectrum, which is calculated as:
α ( ω ) = 2 ω c { [ ε 1 2 ( ω ) + ε 2 2 ( ω ) ] 1 2 ε 1 ( ω ) } 1 2
where ω is the angular frequency, α shows the absorption coefficient and c is the speed of light. In addition, ε1(ω) is used to explain the dielectric constant for real parts, and the imaginary one is demonstrated by ε1(ω). It is obvious that the MXene/MN vdW heterostructure possesses the ability to absorb sunlight over a wide range in the visible and NIR regions, which considerably overlaps with the wavelength range of the solar spectrum. Importantly, one can see that the optical performance of the heterostructures is much better than that of AlN and GaN monolayers. Near the wavelength range of visible light, the calculated optical absorption peaks of the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures are 3.627 × 105 cm−1 and 3.778 ×105 cm−1, respectively. Furthermore, the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures also possess another peak value obtained by 1.113 × 105 cm−1 and 0.962 × 105 cm−1, located at 405 nm and 410 nm, which is higher than that of 0.853 × 105 cm−1 for the Hf2CO2 monolayer. All these results reveal that both Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures have novel optical characteristics. It is worth noting that the calculated optical spectra of these 2D materials in this work do not consider the electron–hole interaction. At present, the GW+BSE method has been regarded to be a very crediable method for including the electron–hole interaction, which has been applied in other low-dimensional materials [104,105].

4. Conclusions

The structural and electronic properties of Hf2CO2, AlN and GaN monolayers and their heterostructures are investigated by the DFT method. Both Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures have strong thermal stability and maintain the original structure at 300 K. Importantly, it is found that both Hf2CO2/AlN and Hf2CO2/GaN heterostructures are formed by vdW interactions, showing a type-I band structure with a bandgap of 2.006 eV and 1.899 eV, respectively, and they are ideal candidates for light-emitting devices. In addition, the potential of Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures is reduced by 6.445 eV and 3.752 eV, respectively. Furthermore, Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures have excellent light absorption ability, which can provide theoretical support and technical guidance for future light-emitting device materials. For the efficiency of the light-emitting device, some properties play an important role, such as carrier mobility, lifetime, diffusion and light emission capability, which can be studied further.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano11092236/s1, Table S1: The tested results of the energy for the Hf2CO2/AlN and Hf2CO2/GaN systems; Figure S1 and Figure S2: The band structure of the MXene/MX heterostructure for other stacking styles; Figure S3 and Figure S4: The charge difference of the MXene/MX heterostructure for other stacking styles; Table S2: The electron transfer between the interface of the Hf2CO2/AlN and Hf2CO2/GaN vdW heterostructures.

Author Contributions

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

Funding

This investigation was supported by the Open Fund Project of Maanshan Engineering Technology, Research Center of Advanced Design for Automotive Stamping Dies (Grant number: QMSG202105).

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geim, A.K.; Novoselov, K. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  2. Jariwala, D.; Sangwan, V.K.; Lauhon, L.; Marks, T.J.; Hersam, M.C. Carbon nanomaterials for electronics, optoelectronics, photovoltaics, and sensing. Chem. Soc. Rev. 2013, 42, 2824–2860. [Google Scholar] [CrossRef] [Green Version]
  3. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-Based Ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, J.; Deng, S.; Liu, Z.; Liu, Z. The rare two-dimensional materials with Dirac cones. Natl. Sci. Rev. 2015, 2, 22–39. [Google Scholar] [CrossRef] [Green Version]
  5. Sun, M.; Tang, W.; Li, S.; Chou, J.-P.; Hu, A.; Schwingenschlögl, U. Molecular doping of blue phosphorene: A first-principles investigation. J. Phys. Condens. Matter 2019, 32, 055501. [Google Scholar] [CrossRef] [Green Version]
  6. Sun, M.; Schwingenschlögl, U. Unique Omnidirectional Negative Poisson’s Ratio in δ-Phase Carbon Monochalcogenides. J. Phys. Chem. C 2021, 125, 4133–4138. [Google Scholar] [CrossRef]
  7. Sun, M.; Schwingenschlögl, U. δ-CS: A Direct-Band-Gap Semiconductor Combining Auxeticity, Ferroelasticity, and Potential for High-Efficiency Solar Cells. Phys. Rev. Appl. 2020, 14, 044015. [Google Scholar] [CrossRef]
  8. Palummo, M.; Bernardi, M.; Grossman, J.C. Exciton Radiative Lifetimes in Two-Dimensional Transition Metal Dichalcogenides. Nano Lett. 2015, 15, 2794–2800. [Google Scholar] [CrossRef] [Green Version]
  9. Bernardi, M.; Palummo, M.; Grossman, J.C. Semiconducting Monolayer Materials as a Tunable Platform for Excitonic Solar Cells. ACS Nano 2012, 6, 10082–10089. [Google Scholar] [CrossRef] [Green Version]
  10. Chen, H.-Y.; Palummo, M.; Sangalli, D.; Bernardi, M. Theory and Ab Initio Computation of the Anisotropic Light Emission in Monolayer Transition Metal Dichalcogenides. Nano Lett. 2018, 18, 3839–3843. [Google Scholar] [CrossRef] [Green Version]
  11. Sun, M.; Luo, Y.; Yan, Y.; Schwingenschlögl, U. Ultrahigh Carrier Mobility in the Two-Dimensional Semiconductors B8Si4, B8Ge4, and B8Sn4. Chem. Mater. 2021, 33, 6475–6483. [Google Scholar] [CrossRef]
  12. 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] [Green Version]
  13. Butler, S.Z.; Hollen, S.; Cao, L.; Cui, Y.; Gupta, J.A.; Gutiérrez, H.R.; Heinz, T.F.; Hong, S.S.; Huang, J.; Ismach, A.F.; et al. Progress, Challenges, and Opportunities in Two-Dimensional Materials Beyond Graphene. ACS Nano 2013, 7, 2898–2926. [Google Scholar] [CrossRef]
  14. Zheng, Z.; Ren, K.; Huang, Z.; Zhu, Z.; Wang, K.; Shen, Z.; Yu, J. Remarkably improved Curie temperature for two-dimensional CrI3 by gas molecular adsorption: A DFT study. Semicond. Sci. Technol. 2021, 36, 075015. [Google Scholar] [CrossRef]
  15. Sun, M.; Yan, Y.; Schwingenschlögl, U. Beryllene: A Promising Anode Material for Na- and K-Ion Batteries with Ultrafast Charge/Discharge and High Specific Capacity. J. Phys. Chem. Lett. 2020, 11, 9051–9056. [Google Scholar] [CrossRef]
  16. Sun, M.; Chou, J.-P.; Hu, A.; Schwingenschlögl, U. Point Defects in Blue Phosphorene. Chem. Mater. 2019, 31, 8129–8135. [Google Scholar] [CrossRef]
  17. Sun, M.; Schwingenschlögl, U. Structure Prototype Outperforming MXenes in Stability and Performance in Metal-Ion Batteries: A High Throughput Study. Adv. Energy Mater. 2021, 11, 2003633. [Google Scholar] [CrossRef]
  18. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black phosphorus field-effect transistors. Nat. Nanotechnol. 2014, 9, 372–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wu, J.; Mao, N.; Xie, L.; Xu, H.; Zhang, J. Identifying the Crystalline Orientation of Black Phosphorus Using Angle-Resolved Polarized Raman Spectroscopy. Angew. Chem. Int. Ed. 2015, 54, 2366–2369. [Google Scholar] [CrossRef] [PubMed]
  20. Li, Y.; Yang, S.; Li, J. Modulation of the Electronic Properties of Ultrathin Black Phosphorus by Strain and Electrical Field. J. Phys. Chem. C 2014, 118, 23970–23976. [Google Scholar] [CrossRef]
  21. Hong, T.; Chamlagain, B.; Lin, W.; Chuang, H.-J.; Pan, M.; Zhou, Z.; Xu, Y.-Q. Polarized photocurrent response in black phosphorus field-effect transistors. Nanoscale 2014, 6, 8978–8983. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, S.; Yan, Z.; Shengli, Z.; Chen, Z.; Zeng, H. Atomically Thin Arsenene and Antimonene: Semimetal-Semiconductor and Indirect-Direct Band-Gap Transitions. Angew. Chem. Int. Ed. 2015, 54, 3112–3115. [Google Scholar] [CrossRef] [PubMed]
  23. Shu, H.; Li, Y.; Niu, X.; Guo, J. Electronic structures and optical properties of arsenene and antimonene under strain and an electric field. J. Mater. Chem. C 2017, 6, 83–90. [Google Scholar] [CrossRef]
  24. Xu, Y.; Peng, B.; Zhang, H.; Shao, H.; Zhang, R.; Zhu, H. First-principle calculations of optical properties of monolayer arsenene and antimonene allotropes. Ann. Phys. 2017, 529. [Google Scholar] [CrossRef] [Green Version]
  25. Ersan, F.; Akturk, E.; Ciraci, S. Interaction of Adatoms and Molecules with Single-Layer Arsenene Phases. J. Phys. Chem. C 2016, 120, 14345–14355. [Google Scholar] [CrossRef]
  26. Dong, M.M.; He, C.; Zhang, W.X. Tunable Electronic Properties of Arsenene and Transition-Metal Dichalcogenide Heterostructures: A First-Principles Calculation. J. Phys. Chem. C 2017, 121, 22040–22048. [Google Scholar] [CrossRef]
  27. Kamal, C.; Ezawa, M. Arsenene: Two-dimensional buckled and puckered honeycomb arsenic systems. Phys. Rev. B 2015, 91. [Google Scholar] [CrossRef] [Green Version]
  28. Mogulkoc, Y.; Modarresi, M.; Mogulkoc, A.; Ciftci, Y. Electronic and optical properties of bilayer blue phosphorus. Comput. Mater. Sci. 2016, 124, 23–29. [Google Scholar] [CrossRef] [Green Version]
  29. Sun, H.; Liu, G.; Li, Q.; Wan, X. First-principles study of thermal expansion and thermomechanics of single-layer black and blue phosphorus. Phys. Lett. A 2016, 380, 2098–2104. [Google Scholar] [CrossRef]
  30. Mogulkoc, Y.; Modarresi, M.; Mogulkoc, A.; Alkan, B. Electronic and optical properties of boron phosphide/blue phosphorus heterostructures. Phys. Chem. Chem. Phys. 2018, 20, 12053–12060. [Google Scholar] [CrossRef]
  31. Sun, M.; Hao, Y.; Ren, Q.; Zhao, Y.; Du, Y.; Tang, W. Tuning electronic and magnetic properties of blue phosphorene by doping Al, Si, As and Sb atom: A DFT calculation. Solid State Commun. 2016, 242, 36–40. [Google Scholar] [CrossRef]
  32. Tsai, D.-S.; Liu, K.-K.; Lien, D.-H.; Tsai, M.-L.; Kang, C.-F.; Lin, C.-A.; Li, L.-J.; He, J.-H. Few-Layer MoS2with High Broadband Photogain and Fast Optical Switching for Use in Harsh Environments. ACS Nano 2013, 7, 3905–3911. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, J.; Shu, H.; Zhao, T.; Liang, P.; Wang, N.; Cao, D.; Chen, X. Intriguing electronic and optical properties of two-dimensional Janus transition metal dichalcogenides. Phys. Chem. Chem. Phys. 2018, 20, 18571–18578. [Google Scholar] [CrossRef]
  34. Ma, X.; Wu, X.; Wang, H.; Wang, Y. A Janus MoSSe monolayer: A potential wide solar-spectrum water-splitting photocatalyst with a low carrier recombination rate. J. Mater. Chem. A 2018, 6, 2295–2301. [Google Scholar] [CrossRef]
  35. Liao, C.; Tang, L.; Wang, L.; Li, Y.; Xu, J.; Jia, Y. Low-threshold near-infrared lasing at room temperature using low-toxicity Ag2Se quantum dots. Nanoscale 2020, 12, 21879–21884. [Google Scholar] [CrossRef] [PubMed]
  36. Jiang, D.; Ni, C.; Tang, W.; Xiang, N. Numerical simulation of elasto-inertial focusing of particles in straight microchannels. J. Phys. D Appl. Phys. 2021, 54, 065401. [Google Scholar] [CrossRef]
  37. Sun, M.; Schwingenschlögl, U. B2P6: A Two-Dimensional Anisotropic Janus Material with Potential in Photocatalytic Water Splitting and Metal-Ion Batteries. Chem. Mater. 2020, 32, 4795–4800. [Google Scholar] [CrossRef]
  38. Wang, S.; Ren, C.; Tian, H.; Yu, J.; Sun, M. MoS2/ZnO van der Waals heterostructure as a high-efficiency water splitting photocatalyst: A first-principles study. Phys. Chem. Chem. Phys. 2018, 20, 13394–13399. [Google Scholar] [CrossRef]
  39. Kumar, R.; Das, D.; Singh, A.K. C2N/WS2 van der Waals type-II heterostructure as a promising water splitting photocatalyst. J. Catal. 2018, 359, 143–150. [Google Scholar] [CrossRef]
  40. Gong, Y.; Lin, J.; Wang, X.; Shi, G.; Lei, S.; Lin, Z.; Zou, X.; Ye, G.; Vajtai, R.; Yakobson, B.I.; et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 2014, 13, 1135–1142. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, Z.; Zhang, Y.; Xie, Z.; Wei, X.; Guo, T.; Fan, J.; Ni, L.; Tian, Y.; Liu, J.; Duan, L. Tunable electronic properties of an Sb/InSe van der Waals heterostructure by electric field effects. Phys. Chem. Chem. Phys. 2019, 21, 5627–5633. [Google Scholar] [CrossRef] [PubMed]
  42. Ren, K.; Tang, W.; Sun, M.; Cai, Y.; Cheng, Y.; Zhang, G. A direct Z-scheme PtS2/arsenene van der Waals heterostructure with high photocatalytic water splitting efficiency. Nanoscale 2020, 12, 17281–17289. [Google Scholar] [CrossRef]
  43. Ren, K.; Sun, M.; Luo, Y.; Wang, S.; Yu, J.; Tang, W. First-principle study of electronic and optical properties of two-dimensional materials-based heterostructures based on transition metal dichalcogenides and boron phosphide. Appl. Surf. Sci. 2019, 476, 70–75. [Google Scholar] [CrossRef]
  44. Ren, K.; Wang, S.; Luo, Y.; Xu, Y.; Sun, M.; Yu, J.; Tang, W. Strain-enhanced properties of van der Waals heterostructure based on blue phosphorus and g-GaN as a visible-light-driven photocatalyst for water splitting. RSC Adv. 2019, 9, 4816–4823. [Google Scholar] [CrossRef] [Green Version]
  45. Ren, K.; Wang, K.; Cheng, Y.; Tang, W.; Zhang, G. Two-dimensional heterostructures for photocatalytic water splitting: A review of recent progress. Nano Futur. 2020, 4, 032006. [Google Scholar] [CrossRef]
  46. Li, J.; Huang, Z.; Ke, W.; Yu, J.; Ren, K.; Dong, Z. High solar-to-hydrogen efficiency in Arsenene/GaX (X = S, Se) van der Waals heterostructure for photocatalytic water splitting. J. Alloy. Compd. 2021, 866, 158774. [Google Scholar] [CrossRef]
  47. Din, H.U.; Idrees, M.; Rehman, G.; Nguyen, C.V.; Gan, L.-Y.; Ahmad, I.; Maqbool, M.; Amin, B. Electronic structure, optical and photocatalytic performance of SiC–MX2 (M = Mo, W and X = S, Se) van der Waals heterostructures. Phys. Chem. Chem. Phys. 2018, 20, 24168–24175. [Google Scholar] [CrossRef]
  48. Williams, K.R.; Diroll, B.T.; Watkins, N.E.; Rui, X.; Brumberg, A.; Klie, R.F.; Schaller, R.D. Synthesis of Type I PbSe/CdSe Dot-on-Plate Heterostructures with Near-Infrared Emission. J. Am. Chem. Soc. 2019, 141, 5092–5096. [Google Scholar] [CrossRef]
  49. Zheng, W.; Zheng, B.; Jiang, Y.; Yan, C.; Chen, S.; Liu, Y.; Sun, X.; Zhu, C.; Qi, Z.; Yang, T.; et al. Probing and Manipulating Carrier Interlayer Diffusion in van der Waals Multilayer by Constructing Type-I Heterostructure. Nano Lett. 2019, 19, 7217–7225. [Google Scholar] [CrossRef] [PubMed]
  50. Bellus, M.; Li, M.; Lane, S.; Ceballos, F.; Cui, Q.; Zeng, X.C.; Zhao, H. Type-I van der Waals heterostructure formed by MoS2 and ReS2 monolayers. Nanoscale Horiz. 2016, 2, 31–36. [Google Scholar] [CrossRef] [PubMed]
  51. Dorfs, D.; Franzl, T.; Osovsky, R.; Brumer, M.; Lifshitz, E.; Klar, T.; Eychmüller, A. Type-I and Type-II Nanoscale Heterostructures Based on CdTe Nanocrystals: A Comparative Study. Small 2008, 4, 1148–1152. [Google Scholar] [CrossRef]
  52. Lei, J.-C.; Zhang, X.; Zhou, Z. Recent advances in MXene: Preparation, properties, and applications. Front. Phys. 2015, 10, 276–286. [Google Scholar] [CrossRef]
  53. Jiang, Q.; Lei, Y.; Liang, H.; Xi, K.; Xia, C.; Alshareef, H.N. Review of MXene electrochemical microsupercapacitors. Energy Storage Mater. 2020, 27, 78–95. [Google Scholar] [CrossRef]
  54. Ling, Z.; Ren, C.E.; Zhao, M.-Q.; Yang, J.; Giammarco, J.M.; Qiu, J.; Barsoum, M.W.; Gogotsi, Y. Flexible and conductive MXene films and nanocomposites with high capacitance. Proc. Natl. Acad. Sci. USA 2014, 111, 16676–16681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Naguib, M.; Come, J.; Dyatkin, B.; Presser, V.; Taberna, P.-L.; Simon, P.; Barsoum, M.W.; Gogotsi, Y. MXene: A promising transition metal carbide anode for lithium-ion batteries. Electrochem. Commun. 2012, 16, 61–64. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, S.; Li, B.; Li, L.; Tian, Z.; Zhang, Q.; Chen, L.; Zeng, X.C. Highly efficient N2 fixation catalysts: Transition-metal carbides M2C (MXenes). Nanoscale 2019, 12, 538–547. [Google Scholar] [CrossRef] [PubMed]
  57. Xiao, R.; Zhao, C.; Zou, Z.; Chen, Z.; Tian, L.; Xu, H.; Tang, H.; Liu, Q.; Lin, Z.; Yang, X. In situ fabrication of 1D CdS nanorod/2D Ti3C2 MXene nanosheet Schottky heterojunction toward enhanced photocatalytic hydrogen evolution. Appl. Catal. B: Environ. 2019, 268, 118382. [Google Scholar] [CrossRef]
  58. Zhang, X.; Zhang, Z.; Zhou, Z. MXene-based materials for electrochemical energy storage. J. Energy Chem. 2018, 27, 73–85. [Google Scholar] [CrossRef] [Green Version]
  59. Khazaei, M.; Arai, M.; Sasaki, T.; Chung, C.-Y.; Venkataramanan, N.S.; Estili, M.; Sakka, Y.; Kawazoe, Y. Novel Electronic and Magnetic Properties of Two-Dimensional Transition Metal Carbides and Nitrides. Adv. Funct. Mater. 2012, 23, 2185–2192. [Google Scholar] [CrossRef]
  60. Zhan, X.; Si, C.; Zhou, J.; Sun, Z. MXene and MXene-based composites: Synthesis, properties and environment-related applications. Nanoscale Horiz. 2019, 5, 235–258. [Google Scholar] [CrossRef]
  61. He, J.; Ding, G.; Zhong, C.; Li, S.; Li, D.; Zhang, G. Cr2TiC2-based double MXenes: Novel 2D bipolar antiferromagnetic semiconductor with gate-controllable spin orientation toward antiferromagnetic spintronics. Nanoscale 2018, 11, 356–364. [Google Scholar] [CrossRef] [PubMed]
  62. Zhu, B.; Zhang, F.; Qiu, J.; Chen, X.; Zheng, K.; Guo, H.; Yu, J.; Bao, J. A novel Hf2CO2/WS2 van der Waals heterostructure as a potential candidate for overall water splitting photocatalyst. Mater. Sci. Semicond. Process. 2021, 133, 105947. [Google Scholar] [CrossRef]
  63. Gandi, A.N.; Alshareef, H.N.; Schwingenschlögl, U. Thermoelectric Performance of the MXenes M2CO2 (M = Ti, Zr, or Hf). Chem. Mater. 2016, 28, 1647–1652. [Google Scholar] [CrossRef] [Green Version]
  64. Sharma, V.; Kumar, A.; Krishnan, V. Two-dimensional MXene-based heterostructures for photocatalysis. In Handbook of Smart Photocatalytic Materials; Elsevier: Amsterdam, The Netherlands, 2020; pp. 247–267. [Google Scholar] [CrossRef]
  65. Guo, Z.; Miao, N.; Zhou, J.; Sa, B.; Sun, Z. Strain-mediated type-I/type-II transition in MXene/Blue phosphorene van der Waals heterostructures for flexible optical/electronic devices. J. Mater. Chem. C 2016, 5, 978–984. [Google Scholar] [CrossRef]
  66. Fu, C.-F.; Li, X.; Luo, Q.; Yang, J. Two-dimensional multilayer M2CO2(M = Sc, Zr, Hf) as photocatalysts for hydrogen production from water splitting: A first principles study. J. Mater. Chem. A 2017, 5, 24972–24980. [Google Scholar] [CrossRef]
  67. Li, S.; Li, X.; Zhang, R.; Cui, H. Strain-tunable electronic properties and optical properties of Hf2CO2 MXene. Int. J. Quantum Chem. 2020, 120. [Google Scholar] [CrossRef]
  68. Wang, Y.; Ma, S.; Wang, L.; Jiao, Z. A novel highly selective and sensitive NH3 gas sensor based on monolayer Hf2CO2. Appl. Surf. Sci. 2019, 492, 116–124. [Google Scholar] [CrossRef]
  69. Bacaksiz, C.; Sahin, H.; Ozaydin, H.D.; Horzum, S.; Senger, T.; Peeters, F.M. Hexagonal AlN: Dimensional-crossover-driven band-gap transition. Phys. Rev. B 2015, 91, 085430. [Google Scholar] [CrossRef] [Green Version]
  70. Bai, Y.; Deng, K.; Kan, E. Electronic and magnetic properties of an AlN monolayer doped with first-row elements: A first-principles study. RSC Adv. 2015, 5, 18352–18358. [Google Scholar] [CrossRef]
  71. Zhang, C.-W. First-principles study on electronic structures and magnetic properties of AlN nanosheets and nanoribbons. J. Appl. Phys. 2012, 111, 43702. [Google Scholar] [CrossRef]
  72. Zhang, C.-W.; Zheng, F.-B. First-principles prediction on electronic and magnetic properties of hydrogenated AlN nanosheets. J. Comput. Chem. 2011, 32, 3122–3128. [Google Scholar] [CrossRef]
  73. Xu, C.; Xue, L.; Yin, C.; Wang, G. Formation and photoluminescence properties of AlN nanowires. Phys. Status solidi (a) 2003, 198, 329–335. [Google Scholar] [CrossRef]
  74. Xu, D.; He, H.; Pandey, R.; Karna, S.P. Stacking and electric field effects in atomically thin layers of GaN. J. Phys. Condens. Matter 2013, 25, 345302. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, Q.; Hu, H.; Chen, X.; Wang, J. Tailoring band gap in GaN sheet by chemical modification and electric field: Ab initio calculations. Appl. Phys. Lett. 2011, 98, 053102. [Google Scholar] [CrossRef]
  76. Wang, W.; Zheng, Y.; Li, X.; Li, Y.; Zhao, H.; Huang, L.; Yang, Z.; Zhang, X.; Li, G. 2D AlN Layers Sandwiched Between Graphene and Si Substrates. Adv. Mater. 2018, 31, e1803448. [Google Scholar] [CrossRef]
  77. Taniyasu, Y.; Kasu, M.; Makimoto, T. An aluminium nitride light-emitting diode with a wavelength of 210 nanometres. Nat. Cell Biol. 2006, 441, 325–328. [Google Scholar] [CrossRef]
  78. Lee, S.; Kim, D.Y. Characteristics of ZnO/GaN heterostructure formed on GaN substrate by sputtering deposition of ZnO. Mater. Sci. Eng. B 2007, 137, 80–84. [Google Scholar] [CrossRef]
  79. McDermott, E.J.; Kurmaev, E.Z.; Boyko, T.D.; Finkelstein, L.D.; Green, R.J.; Maeda, K.; Domen, K.; Moewes, A. Structural and Band Gap Investigation of GaN:ZnO Heterojunction Solid Solution Photocatalyst Probed by Soft X-ray Spectroscopy. J. Phys. Chem. C 2012, 116, 7694–7700. [Google Scholar] [CrossRef]
  80. Cui, Z.; Li, E.; Ke, X.; Zhao, T.; Yang, Y.; Ding, Y.; Liu, T.; Qu, Y.; Xu, S. Adsorption of alkali-metal atoms on GaN nanowires photocathode. Appl. Surf. Sci. 2017, 423, 829–835. [Google Scholar] [CrossRef]
  81. Liao, J.; Sa, B.; Zhou, J.; Ahuja, R.; Sun, Z. Design of High-Efficiency Visible-Light Photocatalysts for Water Splitting: MoS2/AlN(GaN) Heterostructures. J. Phys. Chem. C 2014, 118, 17594–17599. [Google Scholar] [CrossRef]
  82. Lou, P.; Lee, J.Y. GeC/GaN vdW Heterojunctions: A Promising Photocatalyst for Overall Water Splitting and Solar Energy Conversion. ACS Appl. Mater. Interfaces 2020, 12, 14289–14297. [Google Scholar] [CrossRef]
  83. Capelle, K. A bird’s-eye view of density-functional theory. Braz. J. Phys. 2006, 36, 1318–1343. [Google Scholar] [CrossRef]
  84. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. 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]
  86. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  87. Heyd, J.; Peralta, J.; Scuseria, G.E.; Martin, R.L. Energy band gaps and lattice parameters evaluated with the Heyd-Scuseria-Ernzerhof screened hybrid functional. J. Chem. Phys. 2005, 123, 174101. [Google Scholar] [CrossRef]
  88. Ren, K.; Wang, S.; Luo, Y.; Chou, J.-P.; Yu, J.; Tang, W.; Sun, M. High-efficiency photocatalyst for water splitting: A Janus MoSSe/XN (X = Ga, Al) van der Waals heterostructure. J. Phys. D Appl. Phys. 2020, 53, 185504. [Google Scholar] [CrossRef]
  89. Singh, A.; Mathew, K.; Zhuang, H.L.; Hennig, R.G. Computational Screening of 2D Materials for Photocatalysis. J. Phys. Chem. Lett. 2015, 6, 1087–1098. [Google Scholar] [CrossRef]
  90. Voinova, V.V.; Selivanov, N.A.; Plyushchenko, I.V.; Vokuev, M.F.; Bykov, A.Y.; Klyukin, I.N.; Novikov, A.S.; Zhdanov, A.P.; Grigoriev, M.S.; Rodin, I.A.; et al. Fused 1,2-Diboraoxazoles Based on closo-Decaborate Anion–Novel Members of Diboroheterocycle Class. Molecules 2021, 26, 248. [Google Scholar] [CrossRef]
  91. Adonin, S.A.; Bondarenko, M.A.; Novikov, A.S.; Sokolov, M.N. Halogen Bonding in Isostructural Co(II) Complexes with 2-Halopyridines. Crystals 2020, 10, 289. [Google Scholar] [CrossRef] [Green Version]
  92. Soldatova, N.S.; Suslonov, V.V.; Kissler, T.Y.; Ivanov, D.M.; Novikov, A.S.; Yusubov, M.S.; Postnikov, P.S.; Kukushkin, V.Y. Halogen Bonding Provides Heterooctameric Supramolecular Aggregation of Diaryliodonium Thiocyanate. Crystals 2020, 10, 230. [Google Scholar] [CrossRef] [Green Version]
  93. Nenajdenko, V.; Shikhaliyev, N.; Maharramov, A.; Bagirova, K.; Suleymanova, G.; Novikov, A.; Khrustalev, V.; Tskhovrebov, A. Halogenated Diazabutadiene Dyes: Synthesis, Structures, Supramolecular Features, and Theoretical Studies. Molecules 2020, 25, 5013. [Google Scholar] [CrossRef] [PubMed]
  94. Mikherdov, A.; Novikov, A.; Kinzhalov, M.; Zolotarev, A.; Boyarskiy, V. Intra-/Intermolecular Bifurcated Chalcogen Bonding in Crystal Structure of Thiazole/Thiadiazole Derived Binuclear (Diaminocarbene)PdII Complexes. Crystals 2018, 8, 112. [Google Scholar] [CrossRef] [Green Version]
  95. Kryukova, M.A.; Sapegin, A.V.; Novikov, A.S.; Krasavin, M.; Ivanov, D.M. New Crystal Forms for Biologically Active Compounds. Part 1: Noncovalent Interactions in Adducts of Nevirapine with XB Donors. Crystals 2019, 9, 71. [Google Scholar] [CrossRef] [Green Version]
  96. Kryukova, M.A.; Sapegin, A.V.; Novikov, A.S.; Krasavin, M.; Ivanov, D.M. New Crystal Forms for Biologically Active Compounds. Part 2: Anastrozole as N-Substituted 1,2,4-Triazole in Halogen Bonding and Lp-π Interactions with 1,4-Diiodotetrafluorobenzene. Crystals 2020, 10, 371. [Google Scholar] [CrossRef]
  97. Ostras’, A.S.; Ivanov, D.M.; Novikov, A.S.; Tolstoy, P.M. Phosphine Oxides as Spectroscopic Halogen Bond Descriptors: IR and NMR Correlations with Interatomic Distances and Complexation Energy. Molecules 2020, 25, 1406. [Google Scholar] [CrossRef] [Green Version]
  98. Bolotin, D.S.; Il’In, M.V.; Suslonov, V.V.; Novikov, A.S. Symmetrical Noncovalent Interactions Br···Br Observed in Crystal Structure of Exotic Primary Peroxide. Symmetry 2020, 12, 637. [Google Scholar] [CrossRef] [Green Version]
  99. Klyukin, I.; Vlasova, Y.; Novikov, A.; Zhdanov, A.; Zhizhin, K.; Kuznetsov, N. Theoretical Study of closo-Borate Anions [BnHn]2− (n = 5–12): Bonding, Atomic Charges, and Reactivity Analysis. Symmetry 2021, 13, 464. [Google Scholar] [CrossRef]
  100. Chen, X.; Tian, F.; Persson, C.; Duan, W.; Chen, N.-X. Interlayer interactions in graphites. Sci. Rep. 2013, 3, 3046. [Google Scholar] [CrossRef]
  101. Björkman, T.; Guļāns, A.; Krasheninnikov, A.; Nieminen, R.M. van der Waals Bonding in Layered Compounds from Advanced Density-Functional First-Principles Calculations. Phys. Rev. Lett. 2012, 108, 235502. [Google Scholar] [CrossRef] [Green Version]
  102. Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511–519. [Google Scholar] [CrossRef] [Green Version]
  103. Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 2009, 21, 084204. [Google Scholar] [CrossRef] [PubMed]
  104. Bernardi, M.; Palummo, M.; Grossman, J.C. Extraordinary Sunlight Absorption and One Nanometer Thick Photovoltaics Using Two-Dimensional Monolayer Materials. Nano Lett. 2013, 13, 3664–3670. [Google Scholar] [CrossRef] [PubMed]
  105. Bruno, M.; Palummo, M.; Marini, A.; Del Sole, R.; Ossicini, S. From Si Nanowires to Porous Silicon: The Role of Excitonic Effects. Phys. Rev. Lett. 2007, 98, 036807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The atomic structure and the band structure of the (a) Hf2CO2, (b) AlN and (c) GaN monolayers. The grey, red, black, pink, green and blue marks are Hf, O, C, Al, Ga and N atoms, respectively, and the Fermi energy level is 0, shown by gray dashes.
Figure 1. The atomic structure and the band structure of the (a) Hf2CO2, (b) AlN and (c) GaN monolayers. The grey, red, black, pink, green and blue marks are Hf, O, C, Al, Ga and N atoms, respectively, and the Fermi energy level is 0, shown by gray dashes.
Nanomaterials 11 02236 g001
Figure 2. Top and side views of the (a) AA, (b) AB, (c) AC, (d) AD, (e) AE and (f) AF stacking configurations of the Hf2CO2/MN heterostructure.
Figure 2. Top and side views of the (a) AA, (b) AB, (c) AC, (d) AD, (e) AE and (f) AF stacking configurations of the Hf2CO2/MN heterostructure.
Nanomaterials 11 02236 g002
Figure 3. The calculated AIMD snapshots of the (a) Hf2CO2/AlN and (c) Hf2CO2/GaN vdW heterostructures at 300 K by 5 ps, and the monitoring of the energy and the temperature during the AIMD simulation for the (b) Hf2CO2/AlN and (d) Hf2CO2/GaN vdW heterostructures, respectively.
Figure 3. The calculated AIMD snapshots of the (a) Hf2CO2/AlN and (c) Hf2CO2/GaN vdW heterostructures at 300 K by 5 ps, and the monitoring of the energy and the temperature during the AIMD simulation for the (b) Hf2CO2/AlN and (d) Hf2CO2/GaN vdW heterostructures, respectively.
Nanomaterials 11 02236 g003
Figure 4. The calculated projected band structure of the (a) Hf2CO2/AlN and (c) Hf2CO2/GaN vdW heterostructures; the projected density of states of the (b) Hf2CO2/AlN and (d) Hf2CO2/GaN vdW heterostructures. The energy of the Fermi is 0.
Figure 4. The calculated projected band structure of the (a) Hf2CO2/AlN and (c) Hf2CO2/GaN vdW heterostructures; the projected density of states of the (b) Hf2CO2/AlN and (d) Hf2CO2/GaN vdW heterostructures. The energy of the Fermi is 0.
Nanomaterials 11 02236 g004
Figure 5. Schematic of the carrier transport in the interface of the type-I band structure for the Hf2CO2/MN vdW heterostructure. (a) Feasible and (b) restricted charge transfer path.
Figure 5. Schematic of the carrier transport in the interface of the type-I band structure for the Hf2CO2/MN vdW heterostructure. (a) Feasible and (b) restricted charge transfer path.
Nanomaterials 11 02236 g005
Figure 6. The potential drop for the (a) Hf2CO2/AlN and (b) Hf2CO2/GaN vdW heterostructures between the interface. The yellow demonstration shows the gaining of the electrons, while the cyan one means the loss of electrons. 0.0001 |e| is used for the isosurface level.
Figure 6. The potential drop for the (a) Hf2CO2/AlN and (b) Hf2CO2/GaN vdW heterostructures between the interface. The yellow demonstration shows the gaining of the electrons, while the cyan one means the loss of electrons. 0.0001 |e| is used for the isosurface level.
Nanomaterials 11 02236 g006
Figure 7. The HSE06 functional obtained optical absorption spectrum for the Hf2CO2, AlN, GaN monolayers and Hf2CO2/AlN, Hf2CO2/GaN vdW heterostructures.
Figure 7. The HSE06 functional obtained optical absorption spectrum for the Hf2CO2, AlN, GaN monolayers and Hf2CO2/AlN, Hf2CO2/GaN vdW heterostructures.
Nanomaterials 11 02236 g007
Table 1. The optimized lattice parameter (a, Å), bond length (B, Å), binding energy (Ebinding, meV/Å−2), interface height (H, Å) and bandgap (Eg, eV) obtained by HSE06 method for the Hf2CO2, AlN, GaN monolayers and Hf2CO2/AlN, Hf2CO2/GaN heterostructures.
Table 1. The optimized lattice parameter (a, Å), bond length (B, Å), binding energy (Ebinding, meV/Å−2), interface height (H, Å) and bandgap (Eg, eV) obtained by HSE06 method for the Hf2CO2, AlN, GaN monolayers and Hf2CO2/AlN, Hf2CO2/GaN heterostructures.
aBHf–CBHf–OBM–NEbindingHEg
Hf2CO23.3632.3692.132 1.820
AlN3.127 1.805 4.042
GaN3.283 1.895 3.203
Hf2CO2/AlN3.3282.3542.1211.922−56.981.9241.826
Hf2CO2/GaN3.3292.3552.1011.922−52.442.2351.734
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ren, K.; Zheng, R.; Xu, P.; Cheng, D.; Huo, W.; Yu, J.; Zhang, Z.; Sun, Q. Electronic and Optical Properties of Atomic-Scale Heterostructure Based on MXene and MN (M = Al, Ga): A DFT Investigation. Nanomaterials 2021, 11, 2236. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092236

AMA Style

Ren K, Zheng R, Xu P, Cheng D, Huo W, Yu J, Zhang Z, Sun Q. Electronic and Optical Properties of Atomic-Scale Heterostructure Based on MXene and MN (M = Al, Ga): A DFT Investigation. Nanomaterials. 2021; 11(9):2236. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092236

Chicago/Turabian Style

Ren, Kai, Ruxin Zheng, Peng Xu, Dong Cheng, Wenyi Huo, Jin Yu, Zhuoran Zhang, and Qingyun Sun. 2021. "Electronic and Optical Properties of Atomic-Scale Heterostructure Based on MXene and MN (M = Al, Ga): A DFT Investigation" Nanomaterials 11, no. 9: 2236. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092236

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