Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Charge transition levels and stability of Ni- and Ir-doped β-Ga2O3: a comprehensive hybrid functional study

Quoc Duy Hoa, K. Dien Vobc, Nguyen Thanh Tiend, Huynh Anh Huye, Duc-Quang Hoangfg and Duy Khanh Nguyen*hi
aDepartment of Mathematics and Physics, Universitetet i Stavanger, Stavanger, Norway
bEngineering Research Group, Dong Nai Technology University, Bien Hoa City, Vietnam
cFaculty of Engineering, Dong Nai Technology University, Bien Hoa City, Vietnam
dCollege of Natural Sciences, Can Tho University, Can Tho City, Vietnam
eFaculty of Physics, School of Education, Can Tho University, Can Tho City, Vietnam
fSemiconductor Materials Division, Department of Physics, Chemistry and Biology, Linköping University, SE-581 83, Linköping, Sweden
gFaculty of Applied Sciences, HCMC University of Technology & Education, 01 Vo Van Ngan, Thu Duc, Ho Chi Minh City, Vietnam
hLaboratory for Computational Physics, Institute for Computational Science and Artificial Intelligence, Van Lang University, Ho Chi Minh City, Vietnam. E-mail: khanh.nguyenduy@vlu.edu.vn
iFaculty of Mechanical – Electrical and Computer Engineering, School of Technology, Van Lang University, Ho Chi Minh City, Vietnam

Received 25th December 2024 , Accepted 6th February 2025

First published on 21st February 2025


Abstract

In this study, the optimized hybrid functional HSE(0.26,0.0) is employed to investigate the incorporation of nickel (Ni) and iridium (Ir) dopants in β-Ga2O3. The formation energies and charge transition levels of Ni and Ir at gallium sites are calculated. The results show that Ni prefers substitution at the octahedral (Ga2) site, with a formation energy approximately 1 eV lower than at the tetrahedral (Ga1) site. Ni at the Ga2 site (NiGa2) exhibits both donor and acceptor behaviors, with charge transition levels at (+1/0) 1.0 eV and (0/−1) 2.24 eV above the VBM, respectively. Ir similarly favors the octahedral site, displaying donor behaviors with charge transition levels at (+2/+1) 1.04 eV and (+1/0) 3.15 eV above the VBM. Our computational findings for the charge transition levels of Ni and Ir ions are in good agreement with recent experimental measurements, and they explain the correlation between Ni3+ and Ir4+ ion concentrations observed in electron paramagnetic resonance studies. Additionally, the calculated vertical transitions at 2.56 eV and 4.25 eV for NiGa2, and at 2.91 eV and 4.62 eV for IrGa2, below the conduction band minimum, are in good agreement with optical absorption results, confirming the presence of Ni and Ir substitutions at the Ga2 site in β-Ga2O3. These computational results provide a detailed understanding of the behavior of Ni- and Ir-doped β-Ga2O3, highlighting the potential applications of Ni- and Ir-doped β-Ga2O3 for optoelectronic devices.


1. Introduction

The monoclinic phase of gallium oxide (β-Ga2O3), which is the most thermodynamically stable phase of Ga2O3, has a large band gap of approximately 5 eV.1,2 In recent years, β-Ga2O3 has received a lot of attention and is considered a new competitor to 4H–SiC and GaN materials. In addition to its high thermal and chemical stability, the wide to ultra-wide band gap of β-Ga2O3, resulting in a high breakdown field of around 8 MV cm−1, makes it a potential candidate for applications in high power electronics.3,4 The availability of high-quality single crystals and thin films of β-Ga2O3 holds significant promise for applications in gas sensors,5,6 electroluminescent devices,7,8 and photocatalysis.9–11

Unintentional dopants such as Si and H are often present in β-Ga2O3 crystals, leading to a moderate n-type doping concentration of around mid-1017 cm−3.12–14 Additionally, Sn, Ge, and Nb impurities are also used to create n-type β-Ga2O3 crystals.15–18 However, fabricating p-type crystals of β-Ga2O3 has proven to be very challenging, which limits its potential applications. Recently, considerable attention has been directed towards identifying suitable p-type dopants for β-Ga2O3. Despite extensive efforts, p-type single crystals of β-Ga2O3 have not yet been successfully produced. Various metal acceptors, such as Zn, Mg, and Ni, have been investigated as potential dopants in β-Ga2O3, but the results suggested deep acceptor behaviors of the dopants. Among these dopants, Ni-doped β-Ga2O3 samples show promise for producing semi-insulating substrates for lateral power devices. Gustafson et al.19 have grown a Ni-doped β-Ga2O3 crystal by the Czochralski method using an iridium crucible and radio frequency heating. As a result, the presence of Ir impurity is unavoidable. In fact, by using electron paramagnetic resonance and optical absorption, the presence of Ni and Ir substitution at gallium sites was reported by Gustafson et al.19 The optical absorption peaks at 303 nm and 442 nm were assigned to Ni at the octahedral site. When irradiated at 275 nm, the concentration of Ni3+ increases and the Ir4+ concentration decreases. After increasing the temperature above 375 °C, the concentration of Ni3+ and Ir4+ is restored. In addition, the Ni acceptor level is reported at 1.4 eV above the VBM. In the report of Seyidov et al.20 the acceptor level of Ni was reported at 1.9 eV and a deep donor level at 1.1 eV above the VBM, which was not reported by Gustafson et al.19 In the case of unintentional doping with Ir, IrGa was reported to be an n-type donor, the EPR study indicated nonmagnetic (S = 0) Ir3+, while Ir4+ was reported as S = 1/2 in ref. 19 and 21. In ref. 22, Ir exhibited both double and single donor behaviors, whereas in the study by Zachinskis and co-workers,23 only the single donor state of Ir at the Ga site was observed.

While standard density functional theory (DFT) and its modified versions, such as DFT + U and DFT + U + V, have played a significant role in understanding the electrical and optical properties of semiconductors.24–27 They are not enough when applied to defects in wide band-gap materials.28,29 In recent years, hybrid functionals have emerged as a promising alternative for studying wide band-gap semiconductors, particularly screened hybrids30,31 with the HSE06 functional developed by Heyd, Scuseria, and Ernzerhof being especially notable32,33 More recently, the optimized hybrid functional calculation method has demonstrated its advantages in studying defects in various wide-bandgap semiconductors, including GaN, CuGaS2, and cubic-ZnS.34–36 For β-Ga2O3, the optimized hybrid functional with parameters (α = 0.26, μ = 0.00) has already shown its advantages in reproducing and explaining experimental results both with intrinsic defects and,13,37,38 with the substitution of metals at the Ga site.39,40 The procedure for obtaining the optimized parameters that reproduce the gap-optimized, Koopmans' theorem-compliant hybrid functional for β-Ga2O3 has been discussed in the section III of ref. 37. Although, numerous studies have investigated Ni and Ir in β-Ga2O3, a comprehensive understanding of these two impurities considering various aspects such as formation energy, spin states, and the Fermi-level remains incomplete. In this research, the optimized hybrid functional will be employed to investigate Ni- and Ir-doped β-Ga2O3. In this paper, HSE(0.26,0.00) will be used to investigate Ni- and Ir-doped β-Ga2O3. By comparing our calculated results with previous experimental and simulation studies, this research aims to provide a complete understanding of Ni and Ir impurities in β-Ga2O3. This includes examining formation energies, the most favorable spin states of defects, and correlating Ni and Ir concentrations with EPR spectra, optical absorption, and predicting luminescence emission peaks associated with Ni and Ir substitution in β-Ga2O3.

2. Computational details

In this report, the gap-optimized, Koopmans' theorem-compliant hybrid functional HSE, with 26% Hartree–Fock exchange (α = 0.26) and a conventional hybrid functional (μ = 0.00), where exchange interactions are treated uniformly, was employed. Calculations were carried out using the Vienna Ab initio Simulation Package (VASP 5.4.4) along with the projector augmented wave (PAW) method.41–43 The semi-core d electrons of gallium were treated as part of the valence shell. A plane-wave basis set cutoff was set at 420 eV, with a kinetic energy cutoff for augmentation charges at 840 eV. In this study, Ni and Ir are doped into the 3D bulk structure of β-Ga2O3. A 160-atom supercell, constructed as a 1 × 4 × 2 expansion of the base-centered monoclinic unit cell, was used for defect calculations. The supercell has dimensions of approximately 12.2 Å × 9.1 Å × 11.6 Å, and the lattice parameters adopted for this work are a = 12.25 Å, b = 3.04 Å, c = 5.82 Å, and β = 103.8°.37 The self-consistent electronic energy was converged to 10−4 eV, with ion relaxation continuing until forces on each ion were reduced to below 0.02 eV Å−1. Calculations employed the Γ-approximation for Brillouin zone sampling, proven to be convergent within 0.1 eV for this supercell size in the previous work.44 To mitigate artificial interactions between repeated images of charged defects, the charge correction method from ref. 45 and a high-frequency dielectric constant of ε = 3.55 (ref. 2) have been used in our calculations.

3. Results and discussion

In β-Ga2O3, there are two nonequivalent gallium sites: one with tetrahedral coordination and one with octahedral coordination. The formation energies of nickel and iridium incorporation on the 2 inequivalent gallium sites in different charge states are shown in Fig. 1, as a function of the Fermi-level position between the calculated valence band maximum (VBM) and the conduction band minimum (CBM), using the following expression:
 
image file: d4ra09002k-t1.tif(1)
where Etot[Xq] and Etot[Ga2O3] represent the total energies of the structures containing the substitution X (Ni and Ir in the current research) at charge state q and the pristine β-Ga2O3 supercell, respectively. ni indicates the number of atoms added to (ni > 0) or removed from (ni < 0) the pristine supercell to form the defect. Eqcorr is the correction for the finite-size supercell approach as described in the computational details. μi represents the chemical potential of the atoms in the reservoir corresponding to the crystal growth conditions. The EF is the Fermi-level position between the VBM and CBM. The chemical potential of μNi is bounded by the formation of NiGa2O4 in the O-rich limits (ΔH[NiGa2O4] = −13.79 eV, the calculated heat of formation of NiGa2O4) and Ga3Ni2 in the Ga-rich limit (ΔH[Ga3Ni2] = −2.24 eV). While in the case of iridium the chemical potential μIr is set by Ga9Ir2 in the Ga-rich condition (ΔH[Ga9Ir2] = −4.95 eV) and IrO2 in the O-rich condition (Δ[IrO2] = −2.92 eV).

image file: d4ra09002k-f1.tif
Fig. 1 Formation energy (eV) of NiGa and IrGa under as a function of the Fermi-level position.

In Fig. 1, the red lines represent the charge transition levels of Ni incorporating on Ga sites (NiGa). The results indicate that Ni substitution at the octahedral site (NiGa2) is energetically preferred over the tetrahedral site (NiGa1) across all charge states. This result is in good agreement with previous studies.19,20 NiGa show deep acceptor behaviors with low formation energies for both sites under n-type conditions. The calculated acceptor levels (0/−1) are 2.02 eV and 2.30 eV above the VBM for NiGa2 and NiGa1, respectively. Our calculated results are higher in energy compared to the theoretical charge transition levels in the work of Seyidov et al.20 For the donor states (0/+1), the transition levels are found at 1.26 eV and 1.48 eV above the VBM for NiGa2 and NiGa1, respectively. The differences in the charge transitions levels in our calculations and those in ref. 20 are likely from the choice of using the parameters in hybrid functional and dielectric constant used in charge correction.

In this study, it was also found that substitutional Ni at both Ga sites is preferred in the spin-triplet states in the acceptor state, and the spin-singlet state is more favorable in the neutral charge state. This finding aligns well with the research of Seyidov and co-workers.20 In the donor state, NiGa2 has the lowest formation energy in the singlet spin state, while NiGa1 is stable in the spin-triplet state compared to the non-magnetic singlet spin state which is 0.65 eV higher in formation energy. The spin-triplet state of NiGa1 in the donor charge state is likely caused by the difference in ionic radii between tetrahedral Ga3+ (47 pm) and Ni2+ (69 pm). The larger substitutional Ni ion at the tetrahedral Ga site causes a distortion that likely triggers the Jahn–Teller effect which removes the degeneracy and lowers the system's energy, thus achieving a more stable configuration in the triplet spin state. The singlet spin state of NiGa2 in the +1 charge state aligns with the findings of ref. 20, but the donor state of NiGa1 site was reported to be stable in spin-quartet state in that report.20

Fig. 1 also displays the formation energy of Ir incorporating at Ga sites (IrGa), by the blue lines. Similar to Ni substitution, Ir substitution at the octahedral Ga site is energetically preferred compared to the tetrahedral Ga site due to the ionic radius of Ir (68 pm). The octahedral Ga ion (62 pm) provides more space to accommodate the larger ionic radius of Ir compared to the tetrahedral sites. Ir substitution at the Ga2 site (IrGa2) minimizes lattice strain and leads to more stable dopant incorporation. IrGa2 shows donor behavior, with charge transition levels at 0.94 eV (+2/+1) and 2.9 eV (+1/0). For substitution at the tetrahedral site, the donor states have charge transition levels at 1.35 eV and 3.03 eV for (+2/+1) and (+1/0), respectively. It was also found that IrGa1 has an acceptor state with a charge transition (0/−1) occurring in the band gap at 4.61 eV above the VBM, the compensated acceptor level of IrGa1 was also reported by Ritter et al.22 Our calculated results for IrGa charge transition levels and formation energies are in good agreement at the (+2/+1) transition, but show deeper energy at (+1/0) compared to other studies.22,23

It should be noted that while Ritter and coworkers22 reported all charge states, the authors did not specify the spin state of the substitution. On the other hand, Zachinskis et al.23 demonstrated that the high spin state S = 4/2 has the lowest formation energy for IrGa2 in the neutral charge state, and S = 3/2 is energetically favorable for IrGa1 in the singly donor state. However, no information on the double donor state was provided. In our calculation, for the neutral charge state, IrGa2 adopts the singlet spin state, while IrGa1 has the lowest formation energy in the spin-triplet state, which is 0.25 eV lower than that of the singlet spin state. It is due to the distortion caused by the larger Ir dopant at the Ga1 site and the Jahn–Teller effect, as discussed above with NiGa1 in the donor charge state. In the singly donor states, IrGa at both substitution sites is most stable in the S = 1/2 spin state. The most favorable state for IrGa2 in the doublet spin state is in good agreement with the observation of Ir4+ in EPR measurements.19,21

Experimental studies have shown that Ir is often present in β-Ga2O3 as an unintentional dopant, especially in Ni-doped β-Ga2O3 crystals grown by the Czochralski method.22,46,47 EPR measurements by Gustafson et al.19 have identified the signals of Ni and Ir substitution at Ga sites and their concentration correlation. Specifically, when a Ni-doped sample is irradiated with 275 nm light, the concentration of Ni3+ ions increase, while the concentration of Ir4+ ions decrease. The concentration of Ni2+ can be restored by heating above 375 °C.19 The above EPR results can be explained as follows: when the sample is doped with an acceptor dopant like Ni, the correlation between unintentional n-type dopants (such as Si and H) and Ni pins the Fermi-level around the middle of the band gap, similar Fermi-level range was also reported in ref. 20. If one considers the Fermi-level as in the shaded orange row of Fig. 2, NiGa2 and IrGa2 are energetically favorable in the −1 (Ni2+) and +1 (Ir4+) charge states, respectively. When the sample is irradiated with 275 nm light, electrons are excited to the conduction band, leading to NiGa2 and IrGa2 are identified at 0 (Ni3+) and +2 (Ir5+) charge states. Thus, the EPR measurement shows an increase of Ni3+ ion and a decrease in Ir4+ ion concentration after irradiation. When the sample is heated, electrons recombine with holes, that witness the restoration of Ni2+ and Ir4+ concentrations.


image file: d4ra09002k-f2.tif
Fig. 2 The schematic band diagram illustrates the defect-level energies for NiGa and IrGa defects. The shaded orange row indicates the estimated Fermi-energy range.

Fig. 3 illustrates the localization of electrons at the spin state S = 1/2 of NiGa20 and IrGa21+, corresponding to Ni3+ and Ir4+, as addressed in the experimental studies.19 The wave functions of electrons in this spin state are not only localized at the substitutional site but also at several oxygen sites around the substitutional sites. This spread of wave function localization explains the broad EPR lines of Ni3+ and Ir4+ ions in the experimental observation.19,21 In the analysis of Ni-doped β-Ga2O3 crystals, Gustafson et al.19 also found broad optical absorption bands at 442 nm (2.81 eV) and 303 nm (4.09 eV), which were attributed to Ni substitutions. Experimental data of optical absorption peaks are often interpreted as vertical electron transitions from the defect to the CBM.29,37 In our calculations, the vertical charge transitions of Ni at (+1/0) and (0/−1) are 2.56 eV and 4.25 eV below the CBM, respectively. This is in good agreement with the experimental results in ref. 19. In the case of IrGa2, the vertical transition levels are calculated at 2.92 eV and 4.54 eV for the neutral and +1 charge states, respectively. The vertical transition of the neutral charge state is in line with the study by Ritter et al.,46 which reports an absorption at 2.8 eV. Further details on the comparison of the optical absorption and the calculated vertical transition energies for Ni and Ir substitutions are provided in Table 1.


image file: d4ra09002k-f3.tif
Fig. 3 Diagram of (a) Ni and (b) Ir substitution at Ga2 site. The green and red balls represent Ga, O, respectively. The blue lobes show the localization of electron (a) NiGa20 and (b) IrGa21+ in the spin-singlet state.
Table 1 Charge transition levels with respect to the calculated VBM and CBM (eV), in comparison with values in recent studies in both experimental and HSE calculations
Defects Transition levels Experimental transition Recent HSE calculations This work HSE(0.00,0.26) Optical absorption Recent HSE calculations This work HSE(0.00,0.26)
NiGa2 +1/0 X VBM +1.09 (ref. 20) VBM +1.26 CBM −4.09 (ref. 19) CBM −3.8 (ref. 20) CBM −4.25
CBM −2.81 (ref. 19)
0/−1 VBM +2.0 (ref. 20) VBM +1.88 (ref. 20) VBM +2.02 CBM −3.16 (ref. 20) CBM −2.56
VBM +1.4 (ref. 19)
IrGa2 +2/+1 CBM −2.2 (ref. 46) VBM +1.01 (ref. 22) VBM +0.94 CBM −2.9, −3.5, −4.4 (ref. 48) X CBM −4.54
+1/0 CBM −2.25 (ref. 22) (VBM +2.6) CBM −2.1 (VBM +2.9) CBM −2.8 (ref. 22) CBM −2.99 (ref. 23) CBM −2.92
CBM −2.58 (ref. 23) (VBM +2.15)


We also predicted the emission energy, which corresponds to the emitted wavelength from the recombination of an electron with a hole. The recombination was calculated considering a limited-size supercell, resulting in an accuracy of 100 meV for the photoluminescence energy. The difference between an electron in a shallow donor state and a conduction band electron bound by the trapped hole has not been considered. Therefore, an electron is considered to be at the CBM. The equilibrium geometry of the bound exciton was relaxed under the constraint of the orbital occupations, and the spin multiplicity was kept constant during the entire process. As explained above, the correlation between unintentional dopants and Ni acceptors places the Fermi-level in the middle of the gap. Therefore, under irradiation, the electron of donor NiGa2 is excited to the CBM and then recombines with the hole. In the case of IrGa2 this process occurs in the +1 charge state. The calculated recombination energy between a trapped hole at Ni2+ and an electron at the CBM was found to be 2.23 eV, while for Ir4+, it is 2.55 eV. The luminescence of IrGa2 has yet to be confirmed, but it is likely due to the inadvertent Ir impurity.

4. Conclusion

This study shows a detailed theoretical investigation of Ni and Ir dopants in β-Ga2O3 using the optimized hybrid functional HSE(0.26,0.00), which closely agree with experimental results. The calculations results reveal that both Ni and Ir exhibit a strong preference for substitution at the octahedral gallium site. Ni dopant demonstrates dual donor and acceptor behaviors at charge transition levels (+1/0) 1.0 eV and (0/−1) 2.24 eV above the VBM, Ir impurity shows donor behaviors with charge transition levels (+2/+1) at 1.04 eV and (+1/0) at 3.15 eV above the VBM. These results align well with recent experimental measurements, including the correlation between Ni and Ir ion concentrations observed in electron paramagnetic resonance studies. Moreover, the calculated vertical transition levels for NiGa2 (2.56 eV and 4.25 eV below the CBM) and IrGa2 (2.91 eV and 4.62 eV below the CBM) correspond closely with experimental optical absorption peaks. This comprehensive analysis advances the understanding of Ni- and Ir-doped β-Ga2O3 and underscores their potential applications in optoelectronic devices, particularly in high-resistivity and semi-insulating substrates.

Data availability

The data that supports the findings of this study are available within the article.

Author contributions

Quoc-Duy Ho: conceptualization, investigation, data curation, methodology, formal analysis, visualization, writing – original draft. K. Dien Vo, Nguyen Thanh Tien, Huynh Anh Huy, and Duc-Quang Hoang: investigation, data curation, formal analysis, visualization, writing – editing. Duy Khanh Nguyen: supervision, investigation, data curation, validation, formal analysis, manuscript review and editing.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

Duy Khanh Nguyen acknowledges the supports of Van Lang University (VLU), Ho Chi Minh City, Vietnam.

References

  1. M. Orita, H. Ohta, M. Hirano and H. Hosono, Appl. Phys. Lett., 2000, 77, 4166 CrossRef CAS.
  2. J. Furthmüller and F. Bechstedt, Phys. Rev. B, 2016, 93, 115204 CrossRef.
  3. S. J. Pearton, J. Yang, P. H. I. V. Cary, F. Ren, J. Kim, M. J. Tadjer and M. A. Mastro, Appl. Phys. Rev., 2018, 5, 011301 Search PubMed.
  4. J. Zhang, J. Shi, D.-C. Qi, L. Chen and K. H. L. Zhang, APL Mater., 2020, 8, 020906 CrossRef CAS.
  5. J. Zhu, Z. Xu, S. Ha, D. Li, K. Zhang, H. Zhang and J. Feng, Materials, 2022, 15, 7339 CrossRef CAS PubMed.
  6. A. Afzal, J. Mater., 2019, 5, 542 Search PubMed.
  7. M. M. Afandi, S. Jeong and J. Kim, Opt. Mater., 2023, 146, 114613 CrossRef CAS.
  8. T. Minami, T. Shirai, T. Nakatani and T. Miyata, Jpn. J. Appl. Phys., 2000, 39, L524 CrossRef CAS.
  9. Q. Li, et al., Catal. Commun., 2019, 120, 23 CrossRef CAS.
  10. Y. Hou, L. Wu, X. Wang, Z. Ding, Z. Li and X. Fu, J. Catal., 2007, 250, 12 CrossRef CAS.
  11. E. Filippo, M. Tepore, F. Baldassarre, T. Siciliano, G. Micocci, G. Quarta, L. Calcagnile and A. Tepore, Appl. Surf. Sci., 2015, 338, 69 CrossRef CAS.
  12. J. B. Varley, J. R. Weber, A. Janotti and C. G. Van de Walle, Appl. Phys. Lett., 2010, 97, 142106 CrossRef.
  13. N. T. Son, Q. D. Ho, K. Goto, H. Abe, T. Ohshima, B. Monemar, Y. Kumagai, T. Frauenheim and P. Deák, Appl. Phys. Lett., 2020, 117, 032101 CrossRef CAS.
  14. P. D. C. King, I. McKenzie and T. D. Veal, Appl. Phys. Lett., 2010, 96, 062110 CrossRef.
  15. S. C. Siah, et al., Appl. Phys. Lett., 2015, 107, 252103 CrossRef.
  16. A. Mauze, Y. Zhang, T. Itoh, E. Ahmadi and J. S. Speck, Appl. Phys. Lett., 2020, 117, 222102 CrossRef CAS.
  17. F. Alema, G. Seryogin, A. Osinsky and A. Osinsky, APL Mater., 2021, 9, 091102 CrossRef CAS.
  18. W. Zhou, C. Xia, Q. Sai and H. Zhang, Appl. Phys. Lett., 2017, 111, 242103 CrossRef.
  19. T. D. Gustafson, N. C. Giles, B. C. Holloway, J. Jesenovec, B. L. Dutton, J. S. McCloy, M. D. McCluskey and L. E. Halliburton, J. Appl. Phys., 2022, 132, 185705 CrossRef CAS.
  20. P. Seyidov, J. B. Varley, J.-X. Shen, Z. Galazka, T.-S. Chou, A. Popp, M. Albrecht, K. Irmscher and A. Fiedler, J. Appl. Phys., 2023, 134, 205701 CrossRef CAS.
  21. C. A. Lenyk, et al., J. Appl. Phys., 2019, 125, 045703 CrossRef.
  22. J. R. Ritter, J. Huso, P. T. Dickens, J. B. Varley, K. G. Lynn and M. D. McCluskey, Appl. Phys. Lett., 2018, 113, 052101 CrossRef.
  23. A. Zachinskis, J. Grechenkov, E. Butanovs, A. Platonenko, S. Piskunov, A. I. Popov, J. Purans and D. Bocharov, Sci. Rep., 2023, 13, 8522 CrossRef CAS PubMed.
  24. Y. Z. Abdullahi, R. Caglayan, A. Mogulkoc, Y. Mogulkoc and F. Ersan, JPhys Mater., 2023, 6, 025003 CrossRef CAS.
  25. A. Akyildiz, I. Ilgaz Aysan, Y. Z. Abdullahi, B. Akgenc Hanedar, Z. Demir Vatansever and F. Ersan, Phys. Chem. Chem. Phys., 2024, 26, 26566 RSC.
  26. Y. Z. Abdullahi, Comput. Condens. Matter, 2021, 29, e00614 CrossRef.
  27. Y. Z. Abdullahi and F. Ersan, RSC Adv., 2023, 13, 3290 RSC.
  28. W. R. L. Lambrecht, Phys. Status Solidi B, 2011, 248, 1547 CrossRef CAS.
  29. C. Freysoldt, B. Grabowski, T. Hickel, J. Neugebauer, G. Kresse, A. Janotti and C. G. Van de Walle, Rev. Mod. Phys., 2014, 86, 253 CrossRef.
  30. S. J. Clark and J. Robertson, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 085208 CrossRef.
  31. X. Zheng, A. J. Cohen, P. Mori-Sánchez, X. Hu and W. Yang, Phys. Rev. Lett., 2011, 107, 026403 CrossRef PubMed.
  32. J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2003, 118, 8207 CrossRef CAS.
  33. A. V. Krukau, O. A. Vydrov, A. F. Izmaylov and G. E. Scuseria, J. Chem. Phys., 2006, 125, 224106 CrossRef PubMed.
  34. P. Deák, M. Lorke, B. Aradi and T. Frauenheim, Phys. Rev. B, 2019, 99, 085206 CrossRef.
  35. M. Han, Z. Zeng, T. Frauenheim and P. Deák, Phys. Rev. B, 2017, 96, 165204 CrossRef.
  36. Q. Duy Ho and M. Castillo, Comput. Mater. Sci., 2023, 216, 111827 CrossRef CAS.
  37. P. Deák, Q. Duy Ho, F. Seemann, B. Aradi, M. Lorke and T. Frauenheim, Phys. Rev. B, 2017, 95, 075208 CrossRef.
  38. Q. D. Ho, T. Frauenheim and P. Deák, Phys. Rev. B, 2018, 97, 115163 CrossRef CAS.
  39. Q. D. Ho, D. K. Nguyen and H. A. Huy, Comput. Condens. Matter, 2022, 32, e00727 CrossRef.
  40. Q. D. Ho, T. Frauenheim and P. Deák, J. Appl. Phys., 2018, 124, 145702 CrossRef.
  41. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  42. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS PubMed.
  43. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 49, 14251 CrossRef CAS PubMed.
  44. P. Deák, Q. D. Ho, F. Seemann, B. Aradi, M. Lorke and T. Frauenheim, Phys. Rev. B, 2017, 95, 075208 CrossRef.
  45. C. Freysoldt, J. Neugebauer and C. G. Van de Walle, Phys. Rev. Lett., 2009, 102, 016402 CrossRef PubMed.
  46. J. R. Ritter, K. G. Lynn and M. D. McCluskey, J. Appl. Phys., 2019, 126, 225705 CrossRef.
  47. Z. Galazka, et al., ECS J. Solid State Sci. Technol., 2017, 6, Q3007 CrossRef CAS.
  48. J. Jesenovec, J. Varley, S. E. Karcher and J. S. McCloy, J. Appl. Phys., 2021, 129, 225702 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.