Next Article in Journal
A Novel Method for the Preparation of Fibrous CeO2–ZrO2–Y2O3 Compacts for Thermochemical Cycles
Previous Article in Journal
An Environmentally Friendly Method for Removing Hg(II), Pb(II), Cd(II) and Sn(II) Heavy Metals from Wastewater Using Novel Metal–Carbon-Based Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Screening of II-IV-V2 Materials for Photovoltaic Applications Based on Density Functional Theory Calculations

Department of Physics, Kyungpook National University, Daegu 41566, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 28 June 2021 / Revised: 24 July 2021 / Accepted: 27 July 2021 / Published: 29 July 2021

Abstract

:
The relative stability of polymorphs and their electronic structure was investigated for II-IV-V 2 materials by using first-principles density functional theory calculations. Our calculation results show that, for Zn-, Cd-, and Be-containing compounds, nitrides favor the 2H polymorph with AB stacking sequence; however, phosphides, arsenides, and antimonides are more stable in the 3C polymorph with the ABC stacking sequence. The electronic band gap of materials was calculated by using hybrid density functional theory methods, and then materials with an ideal band gap for photovoltaic applications were chosen. The experimental synthesis of the screened materials is reported, except for CdSiSb2, which was found to be unstable in our calculation. The absorption coefficient of the screened materials, especially ZnGeAs2, was high enough to make thin-film solar cells. The higher stacking fault energy in ZnGeAs2 than the others is consistent with the larger formation energy difference between the 2H and 3C polymorphs.

1. Introduction

The progress of photovoltaic technology has been largely due to the finding and optimization of new materials [1,2,3,4]. Among many approaches, cation mutation has remained a notable design principle for finding new materials [5]. Starting from diamond Si, we obtain zinc-blende CdTe (GaAs) that satisfies the octet rule by replacing every two Si atoms with Cd and Te (Ga and As). If two Cd atoms are replaced with Cu and In while Se substitutes for Te, then chalcopyrite CuInSe2 is obtained as illustrated in Figure 1. If we further apply the operation, we can obtain kesterite Cu2ZnSnSe4. All of these materials were investigated extensively for photovoltaic applications both theoretically and experimentally [4]. II-IV-V 2 is another category of materials obtained by applying the design principle to III-V, while ZnSnN2 is the most well-known material for photovoltaic application in this category [6].
Following the cation mutation principle, at least several tens of material can be made, theoretically, as many atoms have the same oxidation number. Such materials can be examined and screened by first-principles calculation as in previous studies [7,8,9,10]. There are many physical properties of constituent materials that determine the solar conversion efficiency [11,12,13,14]. Among them, the band gap is used as the most important metric to screen materials, as the solar conversion efficiency is fundamentally determined by the band gap [4,15].
Shaposhnikov et al. performed a comprehensive computational study of A II B IV C 2 V semiconductors [16], and concluded that ZnSiAs2 and ZnSnAs2 are ideal for photovoltaic applications in terms of the high absorption coefficient and the dipole matrix, not the electronic band gap. To combinatorially generate the materials, Be, Mg, Zn, Cd, Si, Ge, Sn, P, and As were used but antimonides were not considered. More recently, Pandey et al. investigated other II-IV-V 2 materials computationally and suggested several candidates for photovoltaic applications [17]. In their study, the heat of the formation of materials was calculated, using mBEEF meta-GGA functional [18], and then the electronic band gap was evaluated, using a semilocal GLLB-SC functional [19]. However, a specific exchange–correlation functional does not perfectly predict the ground structure of all materials, and therefore it is needed to double-check the prediction using other functionals [20,21,22]. Furthermore, Be was not included in the later study, even though Be also predominantly favors a 2+ oxidation state. It is also worth noting that BeO is stable in the wurtzite phase (2H), while BeS, BeSe, and BeTe are stable in the zinc-blende structure (3C) [23].
In this article, we revisit the stability and the electronic structure of II-IV-V 2 materials to screen materials for photovoltaic applications. The relative formation energy of the 2H and the 3C polymorphs was determined by using three exchange-correlation functionals based on the majority voting. The electronic structure of the materials was calculated using a more stable structure at the hybrid density functional theory. The materials were selected based on the band gap, and the absorption coefficient of the screened materials was obtained accordingly. We also obtained the stacking fault energy for some screened materials for photovoltaic applications because the planar defects have critical roles in terms of transport properties [24,25].

2. Calculation Method

We performed first-principles DFT calculations using the projector-augmented wave (PAW) method [26], and implemented in the Vienna ab initio simulation package (VASP) [27]. We used three exchange-correlation functionals, such as LDA [28], SCAN+rVV10 [29], and TPSS [30], to double-check the relative stability between polymorphs. The energy cutoff for the plane waves was set to 500 eV. A 6 × 6 × 6 k-point grid was used for the primitive cells. The volume and the internal coordinates of atoms were fully relaxed until the residual forces were below 0.01 eV/Å. The electronic structure was investigated by using a hybrid density functional suggested by Heyd, Scuseria, and Ernzerhof (HSE06) [31]. For the HSE06 calculation, the energy cutoff was determined by setting the PREC tag to normal. A body-centered tetragonal unit cell of the 3C polymorph and an orthorhombic unit cell of the 2H polymorph were used throughout the study. The body-centered tetragonal and the orthorhombic cell unit cell contain 2 and 4 formula units, respectively.

3. Results and Discussion

ZnSnN 2 satisfies the octet rule because Zn, Sn, and N have +2, +4, −3 oxidation states, respectively. If each atom is replaced with an atom with the same oxidation state, the resulting material satisfies the octet rule as well. For instance, Zn can be replaced with Be or Cd, keeping the octet rule. Mg is another element that prefers +2 oxidation states, but we did not consider it because we restricted our focus to materials that are stable in the 2H and 3C polymorphs. MgO, MgS, and MgSe favor the rock-salt structure [23], and there is a report of a rock-salt type synthesis of MgSnN2 [32]. Similarly, Si and Ge can substitute for Sn, while N sites can be occupied by P, As, and Sb atoms. The number of combinations satisfying the octet rule is 36.
This category of materials that consist of the tetrahedron is mostly stable in either of the 2H and the 3C polymorphs. To obtain which stacking sequence the materials favor, we compared the relative formation energy of II-IV-V 2 materials in the 2H and the 3C polymorphs. The relative formation energy of the 2H with respect to the 3C was calculated by using three different exchange-correlation functionals, as summarized in Figure 2.
Remarkably, every functional predicts that the same structure has lower formation energy. The only exception is ZnSnSb2, which favored the 2H polymorph a little bit in our LDA calculation. However, in the other two calculations, the ZnSnSb2 in the 3C polymorph has lower energy than that in the 2H polymorph. Following the majority voting rule, we concluded that ZnSnSb2 more favors the 3C structure than the 2H structure. Every nitride was calculated to be stable in the 2H polymorph; the phosphides, arsenides, and antimonides favored the 3C polymorph.
We performed hybrid DFT calculations in the 2H structure for nitride and the 3C for the others to obtain the electronic band gap, as summarized in Figure 3. We employed structures optimized by using SCAN+rVV10 functional for self-consistent field hybrid calculation. The optimized lattice constants were in good agreement with experimental values as summarized in Table 1. Some of the general chemical trends are as follows: If Zn is replaced with Cd, the band gap decreases on average, but on the other hand, the substitution of Be increases the band gap. The electronic band gap increases as Sn atoms are substituted with Ge and Si. The smaller the anion (Sb→As→P→N), the larger the band gap obtained.
The ideal materials for photovoltaic applications should have a direct band gap between 1 eV and 1.5 eV [15]. We chose materials with a direct band gap between 0.9 eV and 1.6 eV, considering a potential error in the calculation. The following five materials satisfied the criteria: CdSnP 2 , ZnSnN 2 , ZnGeAs 2 , CdSiAs 2 , and CdSiSb 2 . The experimental synthesis of CdSiAs 2 , CdSnP 2 , ZnSnN 2 , and ZnGeAs 2 were reported in the literature [33,34,35,36,37,38,39,40]. Here, we note that CdSiAs 2 , CdSnP 2 , and ZnGeAs 2 received less attention, compared to ZnSnN 2 . Here, we also note that Pandey et al. suggested ZnSiSb2, ZnSnN2, and ZnSnP2 as small band gap materials for tandem applications. However, in our calculation, ZnSiSb2 has an indirect band gap and ZnSnP2 has a band gap of about 1.75 eV, and is probably suitable for a larger band gap material in photovoltaic applications. As shown in Table 1, the calculated band gaps are in good agreement with the experimental studies.
On the other hand, we could not find any experimental study of CdSiSb 2 . To check whether CdSiSb 2 can be formed in the chalcopyrite structure, we obtained the heat of formation as follows:
Δ H f = E t o t ( CdSiSb 2 ) μ Cd μ Si 2 μ Sb .
The chemical potential of Cd, Si, and Sb was obtained from hexagonal close-packed Cd, diamond Si, and trigonal Sb, respectively. In our SCAN+rVV10 calculation, the heat of the formation was calculated to be positive ( Δ H f = 0.35 eV/f.u.), indicating that this material is not chemically stable. We double-checked the stability by using the LDA exchange–correlation functional [22]; the calculation result was not changed ( Δ H f = 0.35 eV/f.u.). This explains why there has been no experimental report of this material.
We employed the electronic band gap to screen materials; however, the absorption coefficient might be a more direct physical quantity to design solar cells. To obtain the absorption coefficient ( α ), we obtained the frequency-dependent dielectric function by performing hybrid DFT calculations. To reduce the computational cost, we employed a sparser k-point grid for the Fock potential by setting the reduction factor of 2 [41]. A 10 × 10 × 10 k-point grid was used for ZnSnN2. For the other three materials, a 14 × 14 × 14 grid was employed.
Figure 4 shows the absorption coefficients of the four selected materials. CdSnP2, CdSiAs2, and ZnGeAs2 show comparable absorption coefficients to the ordered ZnSnN2 at wavelengths longer than 600 nm, and for the shorter wavelengths, the absorption coefficient is even higher. Among the four selected materials, ZnGeAs2 has the highest absorption coefficient. This explains the motivation behind previous research studies to make solar cells based on these materials. Here, we point out that this class of materials is relatively new, compared to other mature technologies [6]. Further optimization should be made to improve efficiency, as solar conversion efficiency is a complex function of many parameters [12,13]. Here, we also point out that ZnSnN2 is usually disordered and that the material properties are largely different from the ordered one considered in this study [42,43,44].
The energy difference between the 2H and the 3C polymorphs is generally small as summarized in Figure 2. This small energy difference might cause the simultaneous formation of both polymorphs in the lattice. The stacking fault is a special case in that the metastable phase (the 2H or the 3C) is sandwiched in the host material. For instance, in the zinc-blende materials, the stacking fault can be regarded as a buried wurtzite phase. This planar defect can be formed in the host material by removing or inserting a layer (e.g., ⋯ABCABABC⋯ in the 3C). Both computational and experimental studies show that such planar defects are harmful in terms of carrier transport and defect accumulation [45,46,47].
To examine whether the stacking disorders likely happens in the selected materials, we obtained the stacking fault energy by using the SCAN+rVV10 exchange–correlation functional. We did not calculate the defect in ZnSnN2 because we already investigated it in our previous study. According to our recent study, the stacking fault formation energy in ZnSnN2 is about 0.28 eV/nm 2 [48]. In CdSnP 2 , CdSiAs 2 , and ZnGeAs 2 , the stacking fault energy was 0.06 eV/nm 2 , 0.15 eV/nm 2 , and 0.47 eV/nm 2 , respectively. The higher stacking fault energy in ZnGeAs 2 is well explained by the higher relative formation energy of the 2H polymorph than the other two materials, as shown in Figure 2. The stacking fault can be formed in CdSnP 2 and CdSiAs 2 significantly, but much less formation is expected in ZnGeAs 2 .

4. Summary

We revisited II-IV-V 2 materials for photovoltaic applications based on DFT calculations. Nitrides favor AB stacking sequence, but phosphides, arsenides, and antimonides prefer ABC stacking sequence. Five materials were calculated to have an ideal band gap for a single-junction solar cell, but CdSiSb2 was not thermodynamically stable. The absorption coefficient of CdSnP 2 , ZnSnN 2 , ZnGeAs 2 , and CdSiAs 2 was high enough to make thin-film solar cells. The stacking fault energy was low enough to be formed in CdSnP 2 and CdSiAs 2 , but much higher in ZnGeAs 2 .

Author Contributions

Conceptualization, J.-S.P.; Data curation, B.-H.J., M.J., Y.S. and K.P.; Formal analysis, B.-H.J., M.J., Y.S. and K.P.; Project administration, J.-S.P.; Supervision, J.-S.P.; Writing—Original draft, J.-S.P.; Writing—Review & editing, B.-H.J., M.J., Y.S., K.P. and J.-S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Research Foundation of Korea (NRF) grants funded by the Korea government (MSIT) (Nos. 2019M3D1A2104108 and 2020R1F1A1053606).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

This work was supported by the National Supercomputing Center with supercomputing resources, including technical support (KSC-2020-CRE-0002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Okada, Y.; Ekins-Daukes, N.; Kita, T.; Tamaki, R.; Yoshida, M.; Pusch, A.; Hess, O.; Phillips, C.; Farrell, D.; Yoshida, K.; et al. Intermediate band solar cells: Recent progress and future directions. Appl. Phys. Rev. 2015, 2, 021302. [Google Scholar] [CrossRef] [Green Version]
  2. Haverkort, J.E.; Garnett, E.C.; Bakkers, E.P. Fundamentals of the nanowire solar cell: Optimization of the open circuit voltage. Appl. Phys. Rev. 2018, 5, 031106. [Google Scholar] [CrossRef] [Green Version]
  3. Li, S.; Li, C.Z.; Shi, M.; Chen, H. New phase for organic solar cell research: Emergence of Y-series electron acceptors and their perspectives. ACS Energy Lett. 2020, 5, 1554–1567. [Google Scholar] [CrossRef]
  4. Nayak, P.K.; Mahesh, S.; Snaith, H.J.; Cahen, D. Photovoltaic solar cell technologies: Analysing the state of the art. Nat. Rev. Mater. 2019, 4, 269–285. [Google Scholar] [CrossRef]
  5. Walsh, A.; Chen, S.; Wei, S.H.; Gong, X.G. Kesterite thin-film solar cells: Advances in materials modelling of Cu2ZnSnS4. Adv. Energy Mater. 2012, 2, 400–409. [Google Scholar] [CrossRef]
  6. Khan, I.S.; Heinselman, K.N.; Zakutayev, A. Review of ZnSnN2 semiconductor material. J. Phys. Energy 2020, 2, 032007. [Google Scholar] [CrossRef]
  7. Emery, A.A.; Wolverton, C. High-throughput dft calculations of formation energy, stability and oxygen vacancy formation energy of ABO3 perovskites. Sci. Data 2017, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  8. Jain, A.; Voznyy, O.; Sargent, E.H. High-throughput screening of lead-free perovskite-like materials for optoelectronic applications. J. Phys. Chem. C 2017, 121, 7183–7187. [Google Scholar] [CrossRef]
  9. Huo, Z.; Wei, S.H.; Yin, W.J. High-throughput screening of chalcogenide single perovskites by first-principles calculations for photovoltaics. J. Phys. D Appl. Phys. 2018, 51, 474003. [Google Scholar] [CrossRef]
  10. Oganov, A.R.; Pickard, C.J.; Zhu, Q.; Needs, R.J. Structure prediction drives materials discovery. Nat. Rev. Mater. 2019, 4, 331–348. [Google Scholar] [CrossRef]
  11. Niki, S.; Contreras, M.; Repins, I.; Powalla, M.; Kushiya, K.; Ishizuka, S.; Matsubara, K. CIGS absorbers and processes. Prog. Photovolt. Res. Appl. 2010, 18, 453–466. [Google Scholar] [CrossRef]
  12. Crovetto, A.; Hansen, O. What is the band alignment of Cu2ZnSn(S,Se)4 solar cells? Sol. Energy Mater. Sol. Cells 2017, 169, 177–194. [Google Scholar] [CrossRef] [Green Version]
  13. Kanevce, A.; Reese, M.O.; Barnes, T.; Jensen, S.; Metzger, W. The roles of carrier concentration and interface, bulk, and grain-boundary recombination for 25% efficient CdTe solar cells. J. Appl. Phys. 2017, 121, 214506. [Google Scholar] [CrossRef]
  14. Kurtz, S.; Repins, I.; Metzger, W.K.; Verlinden, P.J.; Huang, S.; Bowden, S.; Tappan, I.; Emery, K.; Kazmerski, L.L.; Levi, D. Historical analysis of champion photovoltaic module efficiencies. IEEE J. Photovolt. 2018, 8, 363–372. [Google Scholar] [CrossRef]
  15. Shockley, W.; Queisser, H.J. Detailed balance limit of efficiency of p-n junction solar cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  16. Shaposhnikov, V.; Krivosheeva, A.; Borisenko, V.; Lazzari, J.L.; d’Avitaya, F.A. Ab initio modeling of the structural, electronic, and optical properties of AIIBIVC2V semiconductors. Phys. Rev. B 2012, 85, 205201. [Google Scholar] [CrossRef]
  17. Pandey, M.; Kuhar, K.; Jacobsen, K.W. II–IV–V2 and III–III–V2 Polytypes as Light Absorbers for Single Junction and Tandem Photovoltaic Devices. J. Phys. Chem. C 2017, 121, 17780–17786. [Google Scholar] [CrossRef]
  18. Wellendorff, J.; Lundgaard, K.T.; Jacobsen, K.W.; Bligaard, T. mBEEF: An accurate semi-local Bayesian error estimation density functional. J. Chem. Phys. 2014, 140, 144107. [Google Scholar] [CrossRef] [Green Version]
  19. Gritsenko, O.; van Leeuwen, R.; van Lenthe, E.; Baerends, E.J. Self-consistent approximation to the Kohn-Sham exchange potential. Phys. Rev. A 1995, 51, 1944. [Google Scholar] [CrossRef] [Green Version]
  20. Zhang, Y.; Kitchaev, D.A.; Yang, J.; Chen, T.; Dacek, S.T.; Sarmiento-Pérez, R.A.; Marques, M.A.; Peng, H.; Ceder, G.; Perdew, J.P.; et al. Efficient first-principles prediction of solid stability: Towards chemical accuracy. Npj Comput. Mater. 2018, 4, 1–6. [Google Scholar] [CrossRef]
  21. Yang, J.H.; Kitchaev, D.A.; Ceder, G. Rationalizing accurate structure prediction in the meta-GGA SCAN functional. Phys. Rev. B 2019, 100, 035132. [Google Scholar] [CrossRef] [Green Version]
  22. Park, J.S. Comparison study of exchange-correlation functionals on prediction of ground states and structural properties. Curr. Appl. Phys. 2021, 22, 61–64. [Google Scholar] [CrossRef]
  23. Wyckoff, R.W.G. Crystal Structures; Interscience Publishers: New York, NY, USA, 1965; Volume 1. [Google Scholar]
  24. Park, J.S.; Walsh, A. Modeling Grain Boundaries in Polycrystalline Halide Perovskite Solar Cells. Annu. Rev. Condens. Matter Phys. 2020, 12, 95–109. [Google Scholar] [CrossRef]
  25. Park, J.S.; Li, Z.; Wilson, J.N.; Yin, W.J.; Walsh, A. Hexagonal Stacking Faults Act as Hole-Blocking Layers in Lead Halide Perovskites. ACS Energy Lett. 2020, 5, 2231–2233. [Google Scholar] [CrossRef]
  26. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. 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. [Google Scholar] [CrossRef] [PubMed]
  28. Ceperley, D.M.; Alder, B.J. Ground state of the electron gas by a stochastic method. Phys. Rev. Lett. 1980, 45, 566. [Google Scholar] [CrossRef] [Green Version]
  29. Peng, H.; Yang, Z.H.; Perdew, J.P.; Sun, J. Versatile van der Waals density functional based on a meta-generalized gradient approximation. Phys. Rev. X 2016, 6, 041005. [Google Scholar] [CrossRef] [Green Version]
  30. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the density functional ladder: Nonempirical meta–generalized gradient approximation designed for molecules and solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [Green Version]
  31. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  32. Kawamura, F.; Imura, M.; Murata, H.; Yamada, N.; Taniguchi, T. Synthesis of a Novel Rocksalt-Type Ternary Nitride Semiconductor MgSnN2 Using the Metathesis Reaction Under High Pressure. Eur. J. Inorg. Chem. 2020, 2020, 446–451. [Google Scholar] [CrossRef]
  33. Reddy, N.; Kistaiah, P.; Murthy, K. High-temperature thermal expansion study of cadmium stannidiphosphide by an X-ray method. J. Phys. D Appl. Phys. 1982, 15, 2247. [Google Scholar] [CrossRef]
  34. Avirović, M.; Lux-Steiner, M.; Elrod, U.; Hönigschmid, J.; Bucher, E. Single crystal growth of CdSiAs2 by chemical vapor transport; its structure and electrical properties. J. Cryst. Growth 1984, 67, 185–194. [Google Scholar] [CrossRef]
  35. Chelluri, B.; Chang, T.; Ourmazd, A.; Dayem, A.; Zyskind, J.; Srivastava, A. Molecular beam epitaxial growth of II–V semiconductor Zn3As2 and II–IV–V chalcopyrite ZnGeAs2. J. Cryst. Growth 1987, 81, 530–535. [Google Scholar] [CrossRef]
  36. Kimmel, M.; Lux-Steiner, M.; Vögt, M.; Bucher, E. n-ZnO/p-CdSiAs2 heterojunction solar cells. J. Appl. Phys. 1988, 64, 1560–1561. [Google Scholar] [CrossRef]
  37. Feldberg, N.; Aldous, J.; Linhart, W.; Phillips, L.; Durose, K.; Stampe, P.; Kennedy, R.; Scanlon, D.; Vardar, G.; Field, R., III; et al. Growth, disorder, and physical properties of ZnSnN2. Appl. Phys. Lett. 2013, 103, 042109. [Google Scholar] [CrossRef]
  38. Peshek, T.J.; Zhang, L.; Singh, R.K.; Tang, Z.; Vahidi, M.; To, B.; Coutts, T.J.; Gessert, T.A.; Newman, N.; Van Schilfgaarde, M. Criteria for improving the properties of ZnGeAs2 solar cells. Prog. Photovolt. Res. Appl. 2013, 21, 906–917. [Google Scholar] [CrossRef]
  39. Martinez, A.D.; Fioretti, A.N.; Toberer, E.S.; Tamboli, A.C. Synthesis, structure, and optoelectronic properties of II–IV–V2 materials. J. Mater. Chem. A 2017, 5, 11418–11435. [Google Scholar] [CrossRef]
  40. Nakatsuka, S.; Inoue, R.; Nose, Y. Fabrication of CdSnP2 Thin Films by Phosphidation for Photovoltaic Application. ACS Appl. Energy Mater. 2018, 1, 1635–1640. [Google Scholar] [CrossRef] [Green Version]
  41. Park, J.S. Examination of high-throughput hybrid calculations using coarser reciprocal space meshes. Curr. Appl. Phys. 2020, 20, 379–383. [Google Scholar] [CrossRef]
  42. Veal, T.D.; Feldberg, N.; Quackenbush, N.F.; Linhart, W.M.; Scanlon, D.O.; Piper, L.F.; Durbin, S.M. Band gap dependence on cation disorder in ZnSnN2 solar absorber. Adv. Energy Mater. 2015, 5, 1501462. [Google Scholar] [CrossRef] [Green Version]
  43. Quayle, P.C.; Blanton, E.W.; Punya, A.; Junno, G.T.; He, K.; Han, L.; Zhao, H.; Shan, J.; Lambrecht, W.R.; Kash, K. Charge-neutral disorder and polytypes in heterovalent wurtzite-based ternary semiconductors: The importance of the octet rule. Phys. Rev. B 2015, 91, 205207. [Google Scholar] [CrossRef] [Green Version]
  44. Lany, S.; Fioretti, A.N.; Zawadzki, P.P.; Schelhas, L.T.; Toberer, E.S.; Zakutayev, A.; Tamboli, A.C. Monte Carlo simulations of disorder in ZnSnN2 and the effects on the electronic structure. Phys. Rev. Mater. 2017, 1, 035401. [Google Scholar] [CrossRef]
  45. Sun, C.; Lu, N.; Wang, J.; Lee, J.; Peng, X.; Klie, R.F.; Kim, M.J. Creating a single twin boundary between two CdTe (111) wafers with controlled rotation angle by wafer bonding. Appl. Phys. Lett. 2013, 103, 252104. [Google Scholar] [CrossRef]
  46. Park, J.S.; Kim, S.; Walsh, A. Opposing effects of stacking faults and antisite domain boundaries on the conduction band edge in kesterite quaternary semiconductors. Phys. Rev. Mater. 2018, 2, 014602. [Google Scholar] [CrossRef] [Green Version]
  47. Li, W.; Rothmann, M.U.; Zhu, Y.; Chen, W.; Yang, C.; Yuan, Y.; Choo, Y.Y.; Wen, X.; Cheng, Y.B.; Bach, U.; et al. The critical role of composition-dependent intragrain planar defects in the performance of MA1−xFAxPbI3 perovskite solar cells. Nat. Energy 2021, 6, 624–632. [Google Scholar] [CrossRef]
  48. Jeong, B.H.; Park, J.S. Stability and electronic structure of stacking faults and polytypes in ZnSnN2, ZnGeN2, and ZnSiN2. J. Korean Phys. Soc. 2021. [Google Scholar] [CrossRef]
  49. Pikhtin, A.; Razbegaev, V.; Goryunova, N.; Leonov, E.; Orlov, V.; Karavaev, G.; Poplavnoi, A.; Chaldyshev, V. Energy band structure and optical properties of CdSnP2. Phys. Status Solidi A 1971, 4, 311–318. [Google Scholar] [CrossRef]
  50. Shay, J.; Buehler, E. Electroreflectance Spectra of CdSiAs2 and CdGeAs2. Phys. Rev. B 1971, 3, 2598. [Google Scholar] [CrossRef]
  51. Shah, S.; Greene, J. Growth and physical properties of ZnGeAs2. Mater. Lett. 1983, 2, 115–118. [Google Scholar] [CrossRef]
  52. Peshek, T.; Tang, Z.; Zhang, L.; Singh, R.; To, B.; Gessert, T.; Coutts, T.; Newman, N.; Van Schilfgaarde, M. ZnGeAs2 thin films properties: A potentially useful semiconductor for photovoltaic applications. In Proceedings of the 2009 34th IEEE Photovoltaic Specialists Conference (PVSC), Philadelphia, PA, USA, 7–12 June 2009; pp. 001367–001369. [Google Scholar]
  53. Senabulya, N.; Feldberg, N.; Makin, R.A.; Yang, Y.; Shi, G.; Jones, C.M.; Kioupakis, E.; Mathis, J.; Clarke, R.; Durbin, S.M. Stabilization of orthorhombic phase in single-crystal ZnSnN2 films. AIP Adv. 2016, 6, 075019. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Atomic structure of (a) zinc-blende CdTe, (b) chalcopyrite CuInSe2, (c) wurtzite GaN, and (d) ZnSnN2.
Figure 1. Atomic structure of (a) zinc-blende CdTe, (b) chalcopyrite CuInSe2, (c) wurtzite GaN, and (d) ZnSnN2.
Crystals 11 00883 g001
Figure 2. (ac) The relative formation energy of material calculated by using LDA, SCAN+rVV10, and TPSS. The positive value means that the 3C polymorph has lower formation energy than the 2H polymorph.
Figure 2. (ac) The relative formation energy of material calculated by using LDA, SCAN+rVV10, and TPSS. The positive value means that the 3C polymorph has lower formation energy than the 2H polymorph.
Crystals 11 00883 g002
Figure 3. The calculated electronic band gap. Indirect band gap materials are denoted by I.
Figure 3. The calculated electronic band gap. Indirect band gap materials are denoted by I.
Crystals 11 00883 g003
Figure 4. The calculated absorption coefficient of selected materials. The vertical lines represent the band gap energies.
Figure 4. The calculated absorption coefficient of selected materials. The vertical lines represent the band gap energies.
Crystals 11 00883 g004
Table 1. Physical properties of the screened materials for single-junction solar cells. The lattice constants a, b, and c are those of the structure optimized by using SCAN+rVV10. The values in parentheses are experimental values. E g and E SFE stand for the electronic band gap and the stacking fault energy, respectively. The electronic band gap of ZnSnN2 varies depending on the cation ordering [6].
Table 1. Physical properties of the screened materials for single-junction solar cells. The lattice constants a, b, and c are those of the structure optimized by using SCAN+rVV10. The values in parentheses are experimental values. E g and E SFE stand for the electronic band gap and the stacking fault energy, respectively. The electronic band gap of ZnSnN2 varies depending on the cation ordering [6].
Materiala (Å)b (Å)c (Å) E g (eV) E SFE (eV/nm 2 )Ref.
CdSnP 2 5.8965.89611.5481.230.06
(5.9015)(5.9015)(11.5144)(1.165) [33,49]
CdSiAs 2 5.8785.87810.8961.560.15
(5.8848)(5.8848)(10.8820)(1.55) [34,50]
ZnGeAs 2 5.6345.63411.1331.160.47
(5.672)(5.672)(11.153)(1.15) [51,52]
ZnSnN 2 6.6995.8225.4641.350.28[48]
(6.8852)(5.9557)(5.5778) [53]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jeong, B.-H.; Jeong, M.; Song, Y.; Park, K.; Park, J.-S. Screening of II-IV-V2 Materials for Photovoltaic Applications Based on Density Functional Theory Calculations. Crystals 2021, 11, 883. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11080883

AMA Style

Jeong B-H, Jeong M, Song Y, Park K, Park J-S. Screening of II-IV-V2 Materials for Photovoltaic Applications Based on Density Functional Theory Calculations. Crystals. 2021; 11(8):883. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11080883

Chicago/Turabian Style

Jeong, Byeong-Hyeon, Minwoo Jeong, Youbin Song, Kanghyeon Park, and Ji-Sang Park. 2021. "Screening of II-IV-V2 Materials for Photovoltaic Applications Based on Density Functional Theory Calculations" Crystals 11, no. 8: 883. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11080883

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