Next Article in Journal
Observation of Metal–Insulator Transition (MIT) in Vanadium Oxides V2O3 and VO2 in XRD, DSC and DC Experiments
Next Article in Special Issue
Suspended 2D Materials: A Short Review
Previous Article in Journal
Non-Collinear Phase in Rare-Earth Iron Garnet Films near the Compensation Temperature
Previous Article in Special Issue
Two-Dimensional VSi2X2N2 (X = P, As, Sb, Bi) Janus Monolayers: Spin-Polarized Electronic Structure and Perpendicular Magnetic Anisotropy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Possibility of Layered Non-Van Der Waals Boron Group Oxides: A First-Principles Perspective

1
Key Laboratory of Flexible Electronics & Institute of Advanced Materials, Jiangsu National Synergetic Innovation Center for Advanced Materials, Nanjing Tech University, 30 South Puzhu Road, Nanjing 211816, China
2
College of Physics Science and Technology, Yangzhou University, Yangzhou 225002, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(9), 1298; https://doi.org/10.3390/cryst13091298
Submission received: 24 July 2023 / Revised: 19 August 2023 / Accepted: 22 August 2023 / Published: 23 August 2023

Abstract

:
Two-dimensional (2D) metal oxides have broad prospective applications in the fields of catalysis, electronic devices, sensors, and detectors. However, non-van der Waals 2D metal oxides have rarely been studied because they are hard to peel off or synthesize. In this work, taking alumina (Al2O3) as a typical representative of 2D boron group oxides, the structural stability and electrical properties of 2D Al2O3 are investigated through first-principles calculations. The thinnest Al2O3 structure is a bilayer, and the band gap of Al2O3 is found to decrease with decreasing layer thickness because of the giant surface reconstruction. The band gap of bilayer X2O3 (X = Al, Ga, and In) decreases with increasing atomic radius. Our findings provide theoretical support for the preparation of non-van der Waals 2D boron group oxide semiconductors.

1. Introduction

Since the discovery of graphene in 2004 [1], two-dimensional (2D) materials have been receiving increasing attention because of their excellent thermal, mechanical, optical, and electrical properties. Besides graphene, other 2D materials, such as black phosphorus, covalent organic frameworks, hexagonal boron nitride, layered double hydroxides, metals, metal oxides, metal–organic frameworks, transition metal carbides/nitrides (MXenes), and transition metal dichalcogenides [2,3,4], have been discovered and investigated extensively. Because metal oxides are abundant in nature, 2D metal oxides are expected to integrate with other 2D materials for electronic, spintronic, and optoelectronic applications [5,6,7,8,9,10,11]. Although some van der Waals (vdW) layered metal oxides, such as SnO and PbO [12,13], have been predicted, the corresponding experimental realization of vdW layered metal oxides is sparse [14]. Therefore, many attempts have recently been made to prepare non-vdW layered metal oxides, such as WO3 and MoO3 [15,16,17,18,19]. Non-vdW metal oxides are difficult to exfoliate into a few layers due to their strong chemical bonds in three dimensions, and the bottom-up synthesis method also poses significant difficulties and challenges [20]. As a result, less has been reported on 2D non-vdW metal oxides, and their corresponding properties have not been fully explored.
Binary non-vdW boron group oxides, such as Al2O3, Ga2O3, and In2O3, are wide-band-gap semiconductors and have been applied in electronic devices and photodetectors [21,22]. Because of its ultra-wide band gap and good thermal conductivity, Al2O3 is often used in ceramic applications [23,24,25]. A bottom-up method to synthesize graphene-like γ-Al2O3 nanosheets has been reported [26]. Using graphene oxide (GO) as template, a homogeneous aluminum sulfate layer was firstly deposited on the GO sheets. Then, the prepared GO–,Al composite sheets were calcined to remove GO and convert basic aluminum sulfate into γ-Al2O3 nanosheets [26]. The γ-Al2O3 nanosheets exhibit faster adsorption kinetics and larger adsorption capacity, which suggests broad applications in catalysis and environmental science. However, the metastable γ-Al2O3 will transform into stable α-Al2O3 under higher temperature. Theoretically, various different monolayer aluminum oxide have been predicted [27,28,29], for example, hexagonal Al2O3 [29] and rectangle AlO2 [27]. Therefore, exploring the stability of 2D α-Al2O3 is highly essential. Ga2O3 is a transparent semiconductor with wide band gap and high critical electric field, and therefore holds the promising applications in power electronics and solar blind photodetectors [30,31,32,33,34]. In2O3 doped by Sn has been widely used for electronic and optoelectronic devices, such as solar cell [35], optical window [36], film transistor [37], and Schottky contact [38]. Because of its ultra-flat surface, Al2O3 (0001) substrate is considered a promising substrate for 2D materials, such as graphene and transition metal dichalcogenides obtained by chemical vapor deposition [39,40,41,42]. Recently, Ga2O3 is also applied as a substrate to grow transition metal dichalcogenides nanoribbons. Because of their similar structures, 2D Al2O3, Ga2O3, and In2O3 are promising when coupled with 2D electronic devices. It has been reported that 2D Fe2O3 was successfully synthesized via the liquid-phase exfoliation of hematite ore in an organic solvent N,N-dimethylformamide (DMF) [43]. Because bulk Fe2O3 has a similar structure to boron group oxides, the synthesis of other non-van der Waals 2D metal oxide materials should thus also be achievable.
In this work, taking 2D Al2O3 as a representative, we design a series of stable 2D boron group oxides using first-principles calculations and investigate their crystal structures, stability, and electronic structures. Two-dimensional Al2O3 is stable down to the bilayer limit and the band gap of few-layer Al2O3 increases with increasing layer thickness. This is the opposite of the band gap thickness trend of vdW layered materials, such as black phosphorus and transition metal dichalcogenides. In addition, we investigate the stability and electronic properties of bilayer Ga2O3 and In2O3, assuming both bilayers are thermally stable. Our work predicts promising stability for 2D non-vdW boron group oxides and provides theoretical support for the future experimental preparation of 2D boron group oxides.

2. Methods

All first-principles calculations in this work, such as structural optimization, phonon dispersion, and electronic properties of Al2O3 with different number of layers, were performed using the QUANTUM ESPRESSO package based on density functional theory (DFT) [44,45]. The Perdew–Burke–Ernzerhof (PBE) functional of the generalized gradient approximation (GGA) was adopted for electron exchange and correlation interactions [46]. Ultrasoft pseudopotentials were selected to describe the interaction between electrons and ions. A vacuum spacing of 20 Å was used to minimize the interaction between neighboring slabs. The cut-off energy of 680 eV and the Brillouin zone sampling of a 10 × 10 × 1 Monkhorst–Pack k-point grid [47] were applied for all layered structures. The convergence threshold for self-consistency was set as 10−11 eV. To ensure the accuracy of the calculation, the structural optimization was continued until the residual forces converged to less than 1.4 × 10−3 eV/Å and the total energy to less than 1.4 × 10−4 eV. The phonon dispersions and Raman spectra were calculated using Phonon code within density-functional perturbation theory [48]. We used norm-conserving pseudopotentials (NCPPs) within the local density approximation (LDA) with a plane-wave cutoff energy of 816 eV to describe the interaction between electrons and ions. A 16 × 16 × 1 Monkhorst−Pack k-mesh and a 8 × 8 × 1 Monkhorst−Pack q-mesh were applied for the phonon dispersion calculations. For the Raman spectra calculations, a 10 × 10 × 1 Monkhorst−Pack k-mesh was used. The threshold for self-consistency was set at 1.36 × 10−13 eV.

3. Results and Discussions

Pristine bulk Al2O3 is trigonal with the space group R 3 ¯ c (Figure 1a). O2− ions are arranged in a hexagonal close packing and Al3+ ions are filled into the octahedral voids. The calculated band gap of bulk Al2O3 is 5.83 eV, which is close to the previously calculated value (6.05 eV) [49]. To model the Al2O3 (0001) surface, we consider the two common surface structures shown in Table 1. The thermodynamic stability of different surface structures can be evaluated based on the surface formation energy, which is defined by the following equation: Ef = Esurface − Ebulk, where Esurface is the energy of an optimized surface structure and Ebulk is the energy of optimized bulk Al2O3. The calculated surface formation energies of type I and II surfaces are listed in Table 1. The calculated surface formation energy of the type I surface is about 3.78 eV, which is a quarter of that of the type II surface. Therefore, the type I surface structure is more stable than the type II surface. In the following, all calculations are performed based on the type I surface structure.
The energetically stable structures of the monolayer, bilayer, trilayer, four-layer, and five-layer Al2O3 are shown in Figure 2a–e. There are five atoms (two Al atoms and three O atoms) in the unit cell of monolayer Al2O3. By analogy, the bilayer, trilayer, four-layer, and five-layer Al2O3 have 10, 15, 20, and 25 atoms, respectively. During the structural optimization, the Al atoms on the upper and lower surfaces gradually move inward, inducing a flat surface structure. The reconstruction of the surface may further enhance the stability of 2D Al2O3. Figure 2f shows the in-plane lattice constants of few-layer and bulk Al2O3. The in-plane lattice constant for monolayer Al2O3 is 5.84 Å, which is 21% larger than that of bulk Al2O3 (4.81 Å). The large change in the lattice constants suggests that the monolayer Al2O3 may be unstable. In addition, the lattices of the bilayer, trilayer, four-layer, and five-layer Al2O3 are similar to that of bulk Al2O3, indicating that bilayer Al2O3 is the thinnest Al2O3 structure that can be fabricated experimentally.
The dynamic stability of few-layer Al2O3 can be examined from the phonon dispersions. The absence of negative frequencies in Figure 3a,b indicates that the monolayer and bilayer Al2O3 are dynamically stable and can exist as free-standing 2D structures. Although monolayer Al2O3 is dynamically stable, the giant lattice difference between monolayer and bulk Al2O3 indicates the difficulty of synthesis monolayer Al2O3 by top-down strategy. Considering the small lattice constant change and high dynamic stability of bilayer Al2O3, it can be inferred that the synthesis of bilayer Al2O3 is quite possible by top-down or bottom-up method. Because of the large number of atoms in the cell of trilayer, four-layer, and five-layer Al2O3, the phonon calculation for these structures are extremely heavy. Here, the phonon dispersions of trilayer, four-layer, and five-layer Al2O3 are not present. The in-plane lattice constants of trilayer, four-layer, and five-layer Al2O3 are close to that of bilayer Al2O3, and we claim that the trilayer, four-layer, and five-layer Al2O3 are likely synthesized experimentally.
Figure 4a–e show the electronic band structures of few-layer Al2O3 calculated via the GGA scheme. The band gaps of the monolayer and five-layer Al2O3 are direct, with the conduction band minimum (CBM) and the valence band maximum (VBM) located at the Γ point. The band gaps of bilayer, trilayer, and four-layer Al2O3 are indirect, with the CBM located at the Γ point and the VBM located at a general point along the Γ−M line. In addition, with the increase in the number of layers, the band structure of Al2O3 gradually shows the characteristics of a flat band in the VBM, which implies a large intensity of the density of states. Moreover, such a flat band in the VBM also indicates the large mass of hole carrier, which limits the mobility of holes.
Figure 4f shows that the band gap of Al2O3 increases with an increase in the number of layers. For comparison, the band structure of bulk Al2O3 is presented in Figure 1b and shows a direct band gap of 5.83 eV. The values of the band gap for monolayer and bilayer Al2O3 are similar (~3.5 eV), and those for the trilayer, four-layer, and five-layer Al2O3 are also similar (~4.7 eV). It is well known that the band gap is always underestimated and lattice constants are overestimated when using GGA calculation due to the lack of Fock exchange. The hybrid functional will improve band gap and lattice constants because of including the exact exchange correction, but will induce a rise in computational costs [50,51]. In order to obtain the more reliable band gaps, we used the HSE06 hybrid functionals to calculate the band structures of monolayer and bilayer Al2O3, as shown in Figure 5. The band gaps are 5.08 and 4.95 eV for monolayer and bilayer Al2O3, respectively. The band gap calculated through using HSE06 hybrid functional is about 1.5 eV larger than that through GGA. However, the band gap of bilayer Al2O3 is also slightly smaller than that of monolayer Al2O3 through using HSE06 hybrid functional, which is consistent with the results through GGA calculation. In addition, the band gap of monolayer Al2O3 is direct, and that of bilayer Al2O3 is indirect when using HSE06 functional, which is also consistent with the result obtained via GGA calculation. Therefore, the band gap calculated using the HSE06 hybrid functional is 1.5 eV larger than that through GGA, but the band gap nature and band gap variation on layer thickness are independent on the functional used. The band gaps of few-layer Al2O3 are wider than 3.0 eV, corresponding to the ultraviolet spectral range. Therefore, the few-layer Al2O3 is suitable for ultraviolet optoelectronic applications. The quantum confinement effect for most semiconductors means that the band gap increases as the thickness decreases. However, for Al2O3, the band gap narrows as the thickness decreases, indicating that the band gap variation with thickness cannot be explained by the quantum confinement effect. From the partial density of the states of on-surface and interior Al atoms for bilayer Al2O3 (Figure 6), we can find that band edge states, which determine the band gap, mainly originate from on-surface Al atoms. Due to the surface states, the band gap of few-layer Al2O3 is smaller than that of bulk Al2O3 (Figure 4f). Because the surface reconstruction of few-layer Al2O3 varies with thickness (Figure 2a–e), the band gap changes with thickness accordingly. Therefore, we attribute the band gap variation with the thickness to the different giant surface reconstruction of the few-layer Al2O3.
Raman spectrum is an effective method with which to identify different materials and structures. In order to facilitate future experimental efforts on few-layer Al2O3, the Raman spectra of monolayer and bilayer Al2O3 were calculated (see Figure 7). Monolayer Al2O3 belongs to the space group P 6 ¯ 2m with 1 f.u. in the primitive cell (five atoms). The zone-center optical phonons can be classified as the following irreducible representations: Γopt = E″ + 3E′ + A″1 + A′1 + A″2 + A′2, where E″, E′, and A′1 are Raman active. For monolayer Al2O3, the peaks are centered at 387, 754, and 1043 cm−1, corresponding to the A′1, E′(1), and E′(2) modes, respectively. The corresponding vibration for A′1, E′(1), and E′(2) modes are listed in Table 2. In addition, in Figure 7a, the intensity of A′1 at 387 cm−1 is significantly higher than the other two peaks. Therefore, the A′1 mode can be used as a Raman signature of monolayer Al2O3. Bilayer Al2O3 belongs to the space group P 3 ¯ 1m with 2 f.u. (10 atoms) in the primitive cell. Therefore, bilayer Al2O3 has 30 vibrational modes. Among these modes, five Eg and three A1g modes are Raman active. For bilayer Al2O3, there are five dominant modes at 367, 521, 554, 797, and 881 cm−1 (Figure 7b) corresponding to the A1g(1), A1g(2), Eg(1), A1g(3), and Eg(2) modes with apparent intensity, respectively. The corresponding vibrations for A1g(1), A1g(2), Eg(1), A1g(3), and Eg(2) modes are shown in Table 3. Furthermore, among these peaks, A1g(3) at 797 cm−1 has the highest intensity and can be considered as the Raman signature of bilayer Al2O3.
Based on the above results, which provide evidence of the stability of bilayer Al2O3, we can assume that the bilayer Ga2O3 and In2O3 are also stable because Al, Ga, and In are in the same group. The absence of negative frequencies in phonon dispersions (Figure 8a,c) demonstrates the stability of bilayer Ga2O3 and In2O3. Bilayer Ga2O3 and In2O3 are indirect semiconductors (Figure 8b,d) with band gaps of 2.01 and 1.17 eV, respectively, which are smaller than that of bilayer Al2O3. The CBM of either bilayer Ga2O3 or In2O3 is located at the Γ point and the VBM at a general point along the K−Γ line. To obtain a variation of band structure on layer thickness, Figure 9 and Figure 10 show the band structures of few-layer Ga2O3 and In2O3, which show a similar band gap variation to that of few-layer Al2O3.
It is also necessary to provide the Raman signature for bilayer Ga2O3 and In2O3. As shown in Figure 11, the Raman spectra of bilayer Ga2O3 and In2O3 were calculated, which were similar to that of bilayer Al2O3. Bilayer Ga2O3 and In2O3 belong to the space group P 3 ¯ 1m with 2 f.u. (10 atoms) in the primitive cell. For bilayer Ga2O3, the peaks are centered at 196, 320, 410, 664, and 689 cm−1 corresponding to the A1g(1), Eg(1), A1g(2), Eg(2), and A1g(3) modes, respectively (Figure 11a). The corresponding vibration for A1g(1), Eg(1), A1g(2), Eg(2), and A1g(3) modes are listed in Table 4. The peak centered at 689 cm−1 (A′1g(3)) with strong intensity should be considered as the Raman signature of bilayer Ga2O3. As for bilayer In2O3, Figure 11b shows that there are four major peaks and two minor peaks in the Raman spectra. Six peaks centered at 133, 269, 333, 357, 580, and 607 cm−1 correspond to the A1g(1), Eg(1), A1g(2), Eg(2), Eg(3), and A1g(3) modes, respectively. All the Raman active vibrational modes for bilayer In2O3 can be found in Table 5. Similarly, the Raman signature of bilayer In2O3 should be the A1g(3) peak, which has a pronounced intensity.

4. Conclusions

In summary, using first-principles calculations based on DFT, we have investigated the structural stability and electronic properties of 2D boron group oxides. The thinnest stable Al2O3 structure is a bilayer with an indirect band gap of 3.41 eV. Interestingly, we find that the band gap of Al2O3 decreases with decreasing layer thickness, which is opposite to the trend seen in vdW 2D materials, such as MoS2. We attribute this to the giant surface reconstruction of Al2O3. In addition, the phonon dispersion curves and band structures show that both bilayer Ga2O3 and In2O3 are also stable, with band structures similar to that of bilayer Al2O3. From bilayer Al2O3 to bilayer Ga2O3 and In2O3, the band gap decreases with the increasing atomic radius. It is expected to synthesize few layer boron group oxides by top-down strategy, such as liquid exfoliation method. This work provides theoretical support for the discovery of non-vdW 2D boron group oxides. Because of their large band gap, non-vdW 2D boron group oxides hold potential applications in ultraviolet optoelectronics and high-power electronics.

Author Contributions

Y.Z. carried out the calculations. Y.C. and D.C. supervised the work. Y.Z., J.Z. and Y.C. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China grant number 62274087.

Data Availability Statement

Data generated and supporting the findings of this article are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, X.; Zeng, Z.; Zhang, H. Metal dichalcogenide nanosheets: Preparation, properties and applications. Chem. Soc. Rev. 2013, 42, 1934–1946. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-like two-dimensional materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, J.; Guo, Y.; Weng, X.; Long, F.; Xin, Y.; Lu, Y.; Ye, Z.; Ruan, S.; Zeng, Y.-J. A tailorable polarity-flipping response in self-powered, flexible Sb2Se3/ZnO bilayer photodetectors. J. Mater. Chem. C 2021, 9, 4978–4988. [Google Scholar] [CrossRef]
  6. Dral, A.P.; ten Elshof, J.E. 2D metal oxide nanoflakes for sensing applications: Review and perspective. Sens. Actuators B Chem. 2018, 272, 369–392. [Google Scholar] [CrossRef]
  7. Huang, S.; Long, Y.; Ruan, S.; Zeng, Y.-J. Enhanced photocatalytic CO2 reduction in defect-engineered Z-scheme WO3-x/g-C3N4 heterostructures. ACS Omega 2019, 4, 15593–15599. [Google Scholar] [CrossRef]
  8. Jiao, S.; Fu, X.; Zhang, L.; Zeng, Y.-J.; Huang, H. Point-defect-optimized electron distribution for enhanced electrocatalysis: Towards the perfection of the imperfections. Nano Today 2020, 31, 100833. [Google Scholar] [CrossRef]
  9. Cakici, M.; Reddy, K.R.; Alonso-Marroquin, F. Advanced electrochemical energy storage supercapacitors based on the flexible carbon fiber fabric-coated with uniform coral-like MnO2 structured electrodes. Chem. Eng. J. 2017, 309, 151–158. [Google Scholar] [CrossRef]
  10. Al-Hardan, N.H.; Abdullah, M.J.; Ahmed, N.M.; Yam, F.K.; Aziz, A.A. UV photodetector behavior of 2D ZnO plates prepared by electrochemical deposition. Superlattices Microstruct. 2012, 51, 765–771. [Google Scholar] [CrossRef]
  11. Yin, Z.; Xu, K.; Jiang, S.; Luo, D.; Chen, R.; Xu, C.; Shum, P.; Liu, Y.J. Recent progress on two-dimensional layered materials for surface enhanced Raman spectroscopy and their applications. Mater. Today Phys. 2021, 18, 100378. [Google Scholar] [CrossRef]
  12. Singh, A.K.; Hennig, R.G. Computational prediction of two-dimensional group-IV mono-chalcogenides. Appl. Phys. Lett. 2014, 105, 042103. [Google Scholar] [CrossRef]
  13. Wang, Y.; Zhang, Q.; Shen, Q.; Cheng, Y.; Schwingenschlogl, U.; Huang, W. Lead monoxide: A two-dimensional ferromagnetic semiconductor induced by hole-doping. J. Mater. Chem. C 2017, 5, 4520–4525. [Google Scholar] [CrossRef]
  14. Daeneke, T.; Atkin, P.; Orrell-Trigg, R.; Zavabeti, A.; Ahmed, T.; Walia, S.; Liu, M.; Tachibana, Y.; Javaid, M.; Greentree, A.D.; et al. Wafer-scale synthesis of semiconducting SnO monolayers from interfacial oxide layers of metallic liquid tin. ACS Nano 2017, 11, 10974–10983. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, B.; Zhou, P.; Xu, Q.; Zhou, X.; Xu, D.; Zhu, J. Engineering disorder into exotic electronic 2D TiO2 nanosheets for enhanced photocatalytic performance. RSC Adv. 2016, 6, 6133–6137. [Google Scholar] [CrossRef]
  16. Tao, J.; Luttrell, T.; Batzill, M. A two-dimensional phase of TiO2 with a reduced bandgap. Nat. Chem. 2011, 3, 296–300. [Google Scholar] [CrossRef]
  17. Addou, R.; Dahal, A.; Batzill, M. Growth of a two-dimensional dielectric monolayer on quasi-freestanding graphene. Nat. Nanotechnol. 2013, 8, 41–45. [Google Scholar] [CrossRef]
  18. Matsuzaki, K.; Hosono, H.; Susaki, T. Layer-by-layer epitaxial growth of polar MgO(111) thin films. Phys. Rev. B 2010, 82, 033408. [Google Scholar] [CrossRef]
  19. Sun, Z.; Liao, T.; Dou, Y.; Hwang, S.M.; Park, M.-S.; Jiang, L.; Kim, J.H.; Dou, S.X. Generalized self-assembly of scalable two-dimensional transition metal oxide nanosheets. Nat. Commun. 2014, 5, 3813. [Google Scholar] [CrossRef]
  20. Yang, M.; Ye, Z.; Iqbal, M.A.; Liang, H.; Zeng, Y.-J. Progress on two-dimensional binary oxide materials. Nanoscale 2022, 14, 9576–9608. [Google Scholar] [CrossRef]
  21. Lee, H.-Y.; Liu, J.-T.; Lee, C.-T. Modulated Al2O3-Alloyed Ga2O3 Materials and Deep Ultraviolet Photodetectors. IEEE Photon. Technol. Lett. 2018, 30, 549–552. [Google Scholar] [CrossRef]
  22. Liu, J.W.; Liao, M.Y.; Imura, M.; Banal, R.G.; Koide, Y. Deposition of TiO2/Al2O3 bilayer on hydrogenated diamond for electronic devices: Capacitors, field-effect transistors, and logic inverters. J. Appl. Phys. 2017, 121, 224502. [Google Scholar] [CrossRef]
  23. Chang, C.-W.; Kuo, C.-P. Evaluation of surface roughness in laser-assisted machining of aluminum oxide ceramics with Taguchi method. Int. J. Mach. Tools Manuf. 2007, 47, 141–147. [Google Scholar] [CrossRef]
  24. Komine, F.; Tomic, M.; Gerds, T.; Strub, J.R. Influence of different adhesive resin cements on the fracture strength of aluminum oxide ceramic posterior crowns. J. Prosthet. Dent. 2004, 92, 359–364. [Google Scholar] [CrossRef] [PubMed]
  25. Vekinis, G.; Ashby, M.F.; Beaumont, P.W.R. R-curbe behaviour of Al2O3 ceramics. Acta Metall. Mater. 1990, 38, 1151–1162. [Google Scholar] [CrossRef]
  26. Huang, Z.; Zhou, A.; Wu, J.; Chen, Y.; Lan, X.; Bai, H.; Li, L. Bottom-up preparation of ultrathin 2D aluminum oxide nanosheets by duplicating graphene oxide. Adv. Mater. 2016, 28, 1703–1708. [Google Scholar] [CrossRef]
  27. Ozyurt, A.K.; Molavali, D.; Sahin, H. Stable single layer structures of aluminum oxide: Vibrational and electronic characterization of magnetic phases. Comput. Mater. Sci. 2022, 214, 111745. [Google Scholar] [CrossRef]
  28. Hasan, S.R.; Abbas, Z.; Khan, M.S. The Al2O3-monolayer sensitivity towards NH3 and PH3 molecule: A DFT Study. J. Water Environ. Nanotechnol. 2023, 8, 31–40. [Google Scholar]
  29. Song, T.T.; Yang, M.; Chai, J.W.; Callsen, M.; Zhou, J.; Yang, T.; Zhang, Z.; Pan, J.S.; Chi, D.Z.; Feng, Y.P.; et al. The stability of aluminium oxide monolayer and its interface with two-dimensional materials. Sci. Rep. 2016, 6, 29221. [Google Scholar] [CrossRef]
  30. Chen, X.; Ren, F.; Gu, S.; Ye, J. Review of gallium-oxide-based solar-blind ultraviolet photodetectors. Photonics Res. 2019, 7, 381–415. [Google Scholar] [CrossRef]
  31. Dong, H.; Long, S.; Sun, H.; Zhao, X.; He, Q.; Qin, Y.; Jian, G.; Zhou, X.; Yu, Y.; Guo, W.; et al. Fast switching beta-Ga2O3 power MOSFET with a trench-gate Structure. IEEE Electron Device Lett. 2019, 40, 1385–1388. [Google Scholar] [CrossRef]
  32. Green, A.J.; Chabak, K.D.; Baldini, M.; Moser, N.; Gilbert, R.; Fitch, R.C., Jr.; Wagner, G.; Galazka, Z.; McCandless, J.; Crespo, A.; et al. Beta-Ga2O3 MOSFETs for radio frequency operation. IEEE Electron Device Lett. 2017, 38, 790–793. [Google Scholar] [CrossRef]
  33. Mukhopadhyay, P.; Schoenfeld, W.V. High responsivity tin gallium oxide Schottky ultraviolet photodetectors. J. Vac. Sci. Technol. A 2020, 38, 013403. [Google Scholar] [CrossRef]
  34. Zhang, D.; Du, Z.; Ma, M.; Zheng, W.; Liu, S.; Huang, F. Enhanced performance of solar-blind ultraviolet photodetector based on Mg-doped amorphous gallium oxide film. Vacuum 2019, 159, 204–208. [Google Scholar] [CrossRef]
  35. Manifacier, J.C.; Szepessy, L. Efficient sprayed In2O3:Sn n-type silicon heterojunction solar cell. Appl. Phys. Lett. 1977, 31, 459–462. [Google Scholar] [CrossRef]
  36. Hamberg, I.; Granqvist, C.G. Evaporated Sn-doped In2O3 films: Basic optical properties and applications to energy-efficient windows. J. Appl. Phys. 1986, 60, R123–R159. [Google Scholar] [CrossRef]
  37. Lee, J.S.; Kwack, Y.J.; Choi, W.S. Inkjet-printed In2O3 thin-film transistor below 200 degrees C. ACS Appl. Mater. Interfaces 2013, 5, 11578–11583. [Google Scholar] [CrossRef]
  38. Von Wenckstern, H.; Splith, D.; Schmidt, F.; Grundmann, M.; Bierwagen, O.; Speck, J.S. Schottky contacts to In2O3. APL Mater. 2014, 2, 046104. [Google Scholar] [CrossRef]
  39. Lee, D.; Seo, J. Graphene growth on sapphire via palladium silicidation. Appl. Surf. Sci. 2019, 492, 23–26. [Google Scholar] [CrossRef]
  40. Liu, H.; Chi, D. Dispersive growth and laser-induced rippling of large-area singlelayer MoS2 nanosheets by CVD on c-plane sapphire substrate. Sci. Rep. 2015, 5, 11756. [Google Scholar] [CrossRef]
  41. Thakur, D.; Sharma, M.; Balakrishnan, V.; Vaish, R. Reusable piezocatalytic water disinfection activity of CVD-grown few-layer WS2 on sapphire substrate. Environ. Sci. Nano 2022, 9, 805–814. [Google Scholar] [CrossRef]
  42. Yang, W.; Mu, Y.; Chen, X.; Jin, N.; Song, J.; Chen, J.; Dong, L.; Liu, C.; Xuan, W.; Zhou, C.; et al. CVD growth of large-area monolayer WS2 film on sapphire through tuning substrate environment and its application for high-sensitive strain sensor. Nanoscale Res. Lett. 2023, 18, 13. [Google Scholar] [CrossRef] [PubMed]
  43. Balan, A.P.; Radhakrishnan, S.; Woellner, C.F.; Sinha, S.K.; Deng, L.; de los Reyes, C.; Rao, B.M.; Paulose, M.; Neupane, R.; Apte, A.; et al. Exfoliation of a non-van der Waals material from iron ore hematite. Nat. Nanotechnol. 2018, 13, 602–609. [Google Scholar] [CrossRef] [PubMed]
  44. Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Nardelli, M.B.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced capabilities for materials modelling with QUANTUM ESPRESSO. J. Phys. Condens. Matter 2017, 29, 465901. [Google Scholar] [CrossRef]
  45. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef]
  46. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  47. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  48. Baroni, S.; de Gironcoli, S.; Dal Corso, A.; Giannozzi, P. Phonons and related crystal properties from density-functional perturbation theory. Rev. Mod. Phys. 2001, 73, 515–562. [Google Scholar] [CrossRef]
  49. Santos, R.C.R.; Longhinotti, E.; Freire, V.N.; Reimberg, R.B.; Caetano, E.W.S. Elucidating the high-k insulator alpha-Al2O3 direct/indirect energy band gap type through density functional theory computations. Chem. Phys. Lett. 2015, 637, 172–176. [Google Scholar] [CrossRef]
  50. Pham, A.; Assadi, M.H.N.; Yu, A.B.; Li, S. Critical role of Fock exchange in characterizing dopant geometry and magnetic interaction in magnetic semiconductors. Phys. Rev. B 2014, 89, 155110. [Google Scholar] [CrossRef]
  51. Sousa, S.F.; Fernandes, P.A.; Ramos, M.J. General performance of density functionals. J. Phys. Chem. A 2007, 111, 10439–10452. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Side view of the structure and (b) calculated electronic band structure for bulk Al2O3. The Fermi level is set at valence band maximum (VBM).
Figure 1. (a) Side view of the structure and (b) calculated electronic band structure for bulk Al2O3. The Fermi level is set at valence band maximum (VBM).
Crystals 13 01298 g001
Figure 2. A side view of the optimized structures for the (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Al2O3. The gray and red balls represent Al and O atoms, respectively. (f) The optimized in-plane lattices of the few-layer and bulk Al2O3.
Figure 2. A side view of the optimized structures for the (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Al2O3. The gray and red balls represent Al and O atoms, respectively. (f) The optimized in-plane lattices of the few-layer and bulk Al2O3.
Crystals 13 01298 g002
Figure 3. Phonon dispersions of (a) monolayer and (b) bilayer Al2O3.
Figure 3. Phonon dispersions of (a) monolayer and (b) bilayer Al2O3.
Crystals 13 01298 g003
Figure 4. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Al2O3, respectively. The Fermi level is set at the valence band maximum (VBM). (f) The bandgap of few-layer and bulk Al2O3.
Figure 4. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Al2O3, respectively. The Fermi level is set at the valence band maximum (VBM). (f) The bandgap of few-layer and bulk Al2O3.
Crystals 13 01298 g004
Figure 5. Band structures of (a) monolayer and (b) bilayer Al2O3 using HSE06 hybrid functional.
Figure 5. Band structures of (a) monolayer and (b) bilayer Al2O3 using HSE06 hybrid functional.
Crystals 13 01298 g005
Figure 6. The partial density of states (PDOS) of Al atoms for bilayer Al2O3.
Figure 6. The partial density of states (PDOS) of Al atoms for bilayer Al2O3.
Crystals 13 01298 g006
Figure 7. The calculated Raman spectra of (a) monolayer and (b) bilayer Al2O3.
Figure 7. The calculated Raman spectra of (a) monolayer and (b) bilayer Al2O3.
Crystals 13 01298 g007
Figure 8. Phonon dispersions of bilayer (a) Ga2O3 and (c) In2O3. Calculated electronic band structures of bilayer (b) Ga2O3 and (d) In2O3. The Fermi level is set at the valence band maximum (VBM).
Figure 8. Phonon dispersions of bilayer (a) Ga2O3 and (c) In2O3. Calculated electronic band structures of bilayer (b) Ga2O3 and (d) In2O3. The Fermi level is set at the valence band maximum (VBM).
Crystals 13 01298 g008
Figure 9. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Ga2O3, respectively. (f) The bandgap of few-layer Ga2O3.
Figure 9. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer Ga2O3, respectively. (f) The bandgap of few-layer Ga2O3.
Crystals 13 01298 g009
Figure 10. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer In2O3, respectively. (f) The bandgap of few-layer In2O3.
Figure 10. Band structures of (a) monolayer, (b) bilayer, (c) trilayer, (d) four-layer, and (e) five-layer In2O3, respectively. (f) The bandgap of few-layer In2O3.
Crystals 13 01298 g010
Figure 11. The calculated Raman spectra of bilayer (a) Ga2O3 and (b) In2O3.
Figure 11. The calculated Raman spectra of bilayer (a) Ga2O3 and (b) In2O3.
Crystals 13 01298 g011
Table 1. The side views and Ef (eV) for type I and II Al2O3(0001) surface before and after optimization.
Table 1. The side views and Ef (eV) for type I and II Al2O3(0001) surface before and after optimization.
Before OptimizationAfter OptimizationEf (eV)
Type ⅠCrystals 13 01298 i001Crystals 13 01298 i0023.78
Type ⅡCrystals 13 01298 i003Crystals 13 01298 i00415.12
Table 2. Vibrational modes for monolayer Al2O3.
Table 2. Vibrational modes for monolayer Al2O3.
ModeFrequency (cm−1)Top View
A′1387Crystals 13 01298 i005
E′(1)754Crystals 13 01298 i006
Crystals 13 01298 i007
E′(2)1043Crystals 13 01298 i008
Crystals 13 01298 i009
Table 3. Vibrational modes for bilayer Al2O3.
Table 3. Vibrational modes for bilayer Al2O3.
ModeFrequency (cm−1)Side View
A1g(1)367Crystals 13 01298 i010
A1g(2)521Crystals 13 01298 i011
Eg(1)554Crystals 13 01298 i012
Crystals 13 01298 i013
A1g(3)797Crystals 13 01298 i014
Eg(2)881Crystals 13 01298 i015
Crystals 13 01298 i016
Table 4. Vibrational modes for bilayer Ga2O3.
Table 4. Vibrational modes for bilayer Ga2O3.
ModeFrequency (cm−1)Side View
A1g(1)196Crystals 13 01298 i017
Eg(1)320Crystals 13 01298 i018
Crystals 13 01298 i019
A1g(2)410Crystals 13 01298 i020
Eg(2)664Crystals 13 01298 i021
Crystals 13 01298 i022
A1g(3)689Crystals 13 01298 i023
Table 5. Vibrational modes for bilayer In2O3.
Table 5. Vibrational modes for bilayer In2O3.
ModeFrequency (cm−1)Side View
A1g(1)133Crystals 13 01298 i024
Eg(1)269Crystals 13 01298 i025
Crystals 13 01298 i026
A1g(2)333Crystals 13 01298 i027
Eg(2)357Crystals 13 01298 i028
Crystals 13 01298 i029
Eg(3)580Crystals 13 01298 i030
Crystals 13 01298 i031
A1g(3)607Crystals 13 01298 i032
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

Zhou, Y.; Zhu, J.; Cai, D.; Cheng, Y. The Possibility of Layered Non-Van Der Waals Boron Group Oxides: A First-Principles Perspective. Crystals 2023, 13, 1298. https://doi.org/10.3390/cryst13091298

AMA Style

Zhou Y, Zhu J, Cai D, Cheng Y. The Possibility of Layered Non-Van Der Waals Boron Group Oxides: A First-Principles Perspective. Crystals. 2023; 13(9):1298. https://doi.org/10.3390/cryst13091298

Chicago/Turabian Style

Zhou, Yu, Jun Zhu, Dongyu Cai, and Yingchun Cheng. 2023. "The Possibility of Layered Non-Van Der Waals Boron Group Oxides: A First-Principles Perspective" Crystals 13, no. 9: 1298. https://doi.org/10.3390/cryst13091298

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