Citation: Radi A. Jishi. Modified Becke-Johnson exchange potential: improved modeling of lead halides for solar cell applications[J]. AIMS Materials Science, 2016, 3(1): 149-159. doi: 10.3934/matersci.2016.1.149
[1] | Etsana Kiros Ashebir, Berhe Tadese Abay, Taame Abraha Berhe . Sustainable A2BⅠBⅢX6 based lead free perovskite solar cells: The challenges and research roadmap for power conversion efficiency improvement. AIMS Materials Science, 2024, 11(4): 712-759. doi: 10.3934/matersci.2024036 |
[2] | Silvia Colodrero . Conjugated polymers as functional hole selective layers in efficient metal halide perovskite solar cells. AIMS Materials Science, 2017, 4(4): 956-969. doi: 10.3934/matersci.2017.4.956 |
[3] | Z. Aboub, B. Daoudi, A. Boukraa . Theoretical study of Ni doping SrTiO3 using a density functional theory. AIMS Materials Science, 2020, 7(6): 902-910. doi: 10.3934/matersci.2020.6.902 |
[4] | M. F. Achoi, S. Kato, N. Kishi, T. Soga . Improved photovoltaic properties of ((CH3NH3)1-xCsx)3Bi2I9: (x = 0-1.0) hybrid perovskite solar cells via a hot immersion method. AIMS Materials Science, 2024, 11(4): 605-619. doi: 10.3934/matersci.2024031 |
[5] | Andrea Ehrmann, Tomasz Blachowicz . Recent coating materials for textile-based solar cells. AIMS Materials Science, 2019, 6(2): 234-251. doi: 10.3934/matersci.2019.2.234 |
[6] | Takuto Eguchi, Shinya Kato, Naoki Kishi, Tetsuo Soga . Effect of thickness on photovoltaic properties of amorphous carbon/fullerene junction. AIMS Materials Science, 2022, 9(3): 446-454. doi: 10.3934/matersci.2022026 |
[7] | Avner Neubauer, Shira Yochelis, Gur Mittelman, Ido Eisenberg, Yossi Paltiel . Simple down conversion nano-crystal coatings for enhancing Silicon-solar cells efficiency. AIMS Materials Science, 2016, 3(3): 1256-1265. doi: 10.3934/matersci.2016.3.1256 |
[8] | Elena Díaz, Rafael Gutiérrez, Christopher Gaul, Gianaurelio Cuniberti, Francisco Domínguez-Adame . Coherent spin dynamics in a helical arrangement of molecular dipoles. AIMS Materials Science, 2017, 4(5): 1052-1061. doi: 10.3934/matersci.2017.5.1052 |
[9] | Said Karim Shah, Jahangeer Khan, Irfan Ullah, Yaqoob Khan . Optimization of active-layer thickness, top electrode and annealing temperature for polymeric solar cells. AIMS Materials Science, 2017, 4(3): 789-799. doi: 10.3934/matersci.2017.3.789 |
[10] | Katrin Gossen, Marius Dotter, Bennet Brockhagen, Jan Lukas Storck, Andrea Ehrmann . Long-term investigation of unsealed DSSCs with glycerol-based electrolytes of different compositions. AIMS Materials Science, 2022, 9(2): 283-296. doi: 10.3934/matersci.2022017 |
Organolead halide perovskites are attracting great interest, mainly because of their photovoltaic applications. These compounds have the general formula ABX3, where A is an organic cation, B is lead, and X is a halide ion. An example is methylammonium lead iodide (CH3NH3PbI3, abbreviated as MAPbI3), where the MA ion is coordinated to 12 I ions, while Pb is octahedrally coordinated to six I ions, with every two adjacent octahedra sharing a corner.
Following publication of the first report [1] on the use of these compounds as light harvesters in photovoltaic cells, many experimental studies have tried to improve upon material preparation methods and enhance their solar-to-electric power conversion efficiency. [2,3,4,5,6,7,8,9,10,11,12,13,14] Theoretical studies have also been undertaken to explore the electronic properties of these compounds and to develop accurate models for computing the energy bands. [15,16,17,18,19,20,21,22,23,24,25]
MAPbI3 has a band gap of 1.5-1.6 eV. [12,26] The possibility of tuning the band gap by various substitutions has been considered. One approach is to replace I with Br or Cl, yielding MAPbBr3 or MAPbCl3, with band gaps given by 2.22 eV and 3.17 eV, respectively.[12,27,28] Another approach is to replace the methylammonium cation with another cation of a different size. When CH3NH3 is replaced with FA = NH2CHNH2 (formamidinium), FAPbI3 with a band gap of 1.43-1.48 eV is obtained. [29,30,31,32,33] The lower band gap in FAPbI3, compared to the one in MAPbI3, leads to an increase in the material’s optical absorption range. Unfortunately, FAPbI3 is unstable under ambient conditions. However, it has been shown recently that the incorporation of MAPbBr3 into FAPbI3 stabilizes the perovskite phase of FAPbI3 and enhances the power conversion efficiency of the solar cell to 18 percent.[34]
In addition to the extensive work demonstrating the applications of organometallic halides in solar cells, it has been shown that these materials have optoelectronic applications in light-emitting diodes.[35,36] For such applications, it is desirable to have materials where the optical band gap can be tuned over a wide range of the visible spectrum. Recent studies [27] have shown that, in MAPbBr3, it is easy to substitute Cl for Br, yielding air-stable MAPbBr3-xClx with x=0-3. The simple cubic lattice structure of MAPbBr3 is maintained as x increases, and the lattice constant decreases linearly with increasing x; the reduction is 5% at x = 3. On the other hand, the band gap increases with increasing x, leading to a tunable band gap in the 400 to 550 nm wavelength range.
The crystal structure of the organometallic halides depends on the radii of the constituent ions
through the Goldschmidt tolerance factor t, which, for the compound ABX3, is given by
t=rA+rX√2(rB+rX) | (1) |
In calculating t, one is faced with the problem of what value to use for the ionic radius of the organic cation A; different values result from different methods of calculation. Amat et al.[37] obtained an ionic radius of 2.70 Å for CH3NH3+ by calculating the volume inside a contour of 0.001 electrons/bohr3 density. Kieslich et al.[38] considered a hard sphere model in which the cation rotates freely about its center of mass. The ionic radius of CH3NH3+ is then taken to be the distance from the cation’s center of mass to nitrogen, plus 1.46 Å (the ionic radius of nitrogen). This method yields a value of 2.17 Å for the ionic radius of the methylammonium ion.
We can set some bounds on the value of the ionic radius of CH3NH3+ by noting that both MAPbBr3 and MAPbCl3 adopt an ideal cubic perovskite structure. Taking the ionic radii of Pb, Br, and Cl to be 1.19 Å, 1.96 Å, and 1.81 Å, respectively, we find that for the tolerance factor t to lie between 0.9 and 1.0, the ionic radius of CH3NH3+ should lie between 2.04 Å and 2.50 Å, These values should not be considered strict limits; they are reasonable estimates. We should note that, in organolead halide compounds, the methylammonium ion does not rotate freely about its center of mass. There is disorder, manifested by the existence of 12 equivalent positions for C and N. In all these positions, the midpoint of the C-N bond is always at, or extremely close to, the center of the cubic unit cell.[39] Thus, we may estimate the ionic radius of CH3NH3+ as one half the C-N bond length plus the ionic radius of nitrogen. Optimizing the structure of the methylammonium cation by using the 6-31G** basis set of gaussian orbitals and the B3LYP exchange potential [40] as implemented within Gaussian 09, [41] the ionic radius of the cation is found to be 2.23 Å. The resulting Goldschmidt tolerance factors are 0.952, 0.941, and 0.924 for MAPbCl3, MAPbBr3, and MAPbI3, respectively. These values are consistent with the fact that these compounds adopt a perovskite structure.
The electronic properties of organometallic halides depend on various factors which can be controlled experimentally. These factors include lattice constants, which can be varied by applying external pressure or internal chemical pressure; the type of halide ion, controlled by chemical substitution; and the type of organic ion. To obtain meaningful results, it is important to use a computational model which accurately describes the known electronic properties of these compounds and which can predict the effect of these variables upon them. Density functional theory (DFT) in the Kohn-Sham formulation [42] is the most widely used method. Here, the exchange potential is approximated by a functional of the electronic density, with the most common approximations being the local density approximation (LDA) [42] and the generalized gradient approximation (GGA). [43] Although the ground state is well described by LDA and GGA, these approximations fail to account for excited-state properties. In many semiconductors the values of the band gaps are severely underestimated. Improved values for the band gaps are obtained by using the GW method.[44] The usefulness of this method, however, is hampered by its high computational cost.
A different exchange potential, introduced by Becke and Johnson, [45] was recently modified by Tran and Blaha.[46] The modified Becke-Johnson (mBJ) potential is given by
VmBJ(r)=cVBRx(r)+(3c−2)1π√512 [2t(r)/ρ(r)]1/2 | (2) |
c=A+B√g | (3) |
We have recently shown that, upon using the modified Becke-Johnson exchange potential with A = 0.4 and B = 1.0 bohr1/2, the calculated band gaps of MAPbI3, MAPbBr3, RbPbI3, and CsPbX3 ( X = Cl, Br, I) are in excellent agreement with experimental values. [24] In this work, we show that applying this method to the lead halide compounds PbCl2, PbBr2, and MAPbBr3-xClx for x = 1, 2, and 3 produces accurate values for the band gaps. We then use this method to demonstrate that a small reduction in the lattice constants of MAPbBr3 and MAPbI3 produces a considerable downshift in the band gaps. Reduction in the lattice constants can be achieved by replacing CH3NH3+ with slightly smaller cations, such as N2H5+ and N2H3+.
Total energy calculations are carried out using the all-electron, full potential, linearized augmented plane wave (FP-LAPW) method as implemented in the WIEN2k code.[48] In this method, space is divided into two regions. One region consists of the interior of non-overlapping muffin-tin spheres centered at the atom sites. The rest of the space (the interstitial) constitutes the other region. In all the calculations reported in this work, the radii of the muffin-tin spheres are 2.1a0 for Pb, Cl, Br, and I, where a0 is the Bohr radius. On the other hand, the radii for C, N, and H are chosen such that the muffintin spheres on adjacent atoms almost touch. The electronic wave function is expanded in terms of a set of basis functions which take different forms in the two regions mentioned above. Inside the muffintin spheres, the basis functions are atomic-like functions which are expanded in terms of spherical harmonics up to lmax = 10. In the interstitial region, they are plane waves with a maximum wave vector Kmax. Each plane wave is augmented by one atomic-like function in each muffin-tin sphere. Usually, Kmax is chosen such that RmtKmax = 6-9, where Rmt is the radius of the smallest muffin-tin sphere in the unit cell. However, due to the very smallness of the muffin-tin radius of hydrogen (RH = 0.65-0.70 a0), we set RHKmax = 3.0. For orthorhombic and tetragonal systems, a 4x4x4 Monkhorst-Pack grid [49] was used for sampling the Brillouin zone, while an 8x8x8 grid was chosen for cubic systems.
The charge density is Fourier-expanded up to a maximum wave vector Gmax = 20a0-1. In GGA calculations, PBEsol functional [50] is used for the exchange potential. This method is well-suited for optimizing atomic positions and lattice constants. In the self-consistent calculations, the total energy and charge were converged to within 0.1 mRy and 0.001 e, respectively. In calculations which employ the modified Becke-Johnson exchange potential, we choose A = 0.4 and B = 1.0 bohr1/2, where A and B are the parameters which appear in Eq.(3).
We begin by carrying out DFT calculations on PbCl2, PbBr2, and MAPbBr3-xClx, for x = 0, 1, 2, and 3. At room temperature, PbCl2 and PbBr2 adopt an orthorhombic crystal structure [51] with space group Pbnm, while MAPbBr3-xClx crystals have a cubic unit cell [12,27]. The experimental values of the lattice constants of these crystals are given in Table 1.
Compound | Structure | Lattice constants (Å) |
PbCl2 | Orthorhombic | a=9.03, b=7.608, c=4.525 |
PbBr2 | Orthorhombic | a=9.466, b=8.068, c=4.767 |
MAPbBr3 | Cubic | a=5.933 |
MAPbBr2Cl | Cubic | a=5.88 |
MAPbBrCl2 | Cubic | a=5.78 |
MAPbCl3 | Cubic | a=5.71 |
In DFT calculations, it is important to take into account spin-orbit coupling, mainly because of the presence of Pb. Our results are summarized in Table 2. In all the compounds under consideration, we find that GGA+SOC severely underestimates the values of the band gaps. On the other hand, our present method (mBJ+SOC, with A = 0.4 and B = 1.0 bohr1/2 in Eq. (3)) yields values for the band gaps which are in excellent agreement with experiment. These results, along with previous calculations [24] on other lead halide compounds, give us confidence in the ability of DFT combined with the mBJ exchange potential to accurately predict the band gaps in all lead halide compounds.
In Fig.1 we present the calculated density of states in PbBr2. The figure shows that the lower-lying conduction bands are derived from Pb orbitals, while both Pb and Br orbitals contribute to the highest valence band. The situation is similar in PbCl2 and MAPbBr3-xClx; the low conduction bands are derived mainly from Pb orbitals while the valence band is composed of Pb and halide orbitals.
We now apply this method to study the variation of the band gap in MAPbBr3 and MAPbI3 with the reduction of the lattice constants. Our results indicate that the band gap in these materials is very sensitive to lattice constant variation. At the experimental values of the lattice constants, our calculations give a band gap of Eg = 1.54 eV in MAPbI3 and Eg = 2.23 eV in MAPbBr3. For a 1% reduction in the lattice constants, we obtain Eg = 1.31 eV in MAPbI3 and Eg = 1:95 eV in MAPbBr3, while for a 2% reduction, we obtain Eg = 1.17 eV in MAPbI3 and Eg = 1.75 eV in MAPbBr3. In arriving at these results, we have assumed that, with reduced lattice constants, MAPbI3 maintains its body-centered tetragonal structure, and MAPbBr3 its simple cubic structure.
A reduction in the lattice constants of organometallic halides can be achieved by applying an external pressure or by replacing CH3NH3+ with a cation of a slightly smaller ionic radius. We consider two cations: N2H5+ (diazanium) and N2H3+ (diazenium). In each case, the ionic radius is taken to be equal to one-half the N-N bond length plus the ionic radius of nitrogen, as discussed in the introduction. Using coupled cluster theory with a perturbative treatment of the triple excitations, Matus et al.[56] calculated the N-N bond lengths in N2H5+ and N2H3+ to be 1.46 Å and 1.24 Å, respectively. Using these values, we obtain 2.19 Å and 2.08 Å for the ionic radii of N2H5+ and N2H3+, respectively. These values are only slightly smaller than the corresponding value for CH3NH3+, estimated by the same method to be 2.23 Å. The Goldschmidt tolerance factors for N2H5PbBr3 and N2H5PbI3 are 0.932 and 0.916, respectively, while for N2H3PbBr3 and N2H3PbI3 they are 0.907 and 0.893, respectively. Thus, the replacement of CH3NH3 with N2H5 or N2H3 leads to only a small change in the tolerance factor. Hence, we assume that N2H5PbBr3 and N2H3PbBr3 will have a cubic unit cell, similar to MAPbBr3, and that N2H5PbI3 and N2H3PbI3 will maintain a body-centered tetragonal structure, as seen in MAPbI3.
Upon carrying out structure optimization, we find that the lattice constants for N2H5PbBr3 and N2H3PbBr3 are a = 5.86 Å and 5.806 Å, respectively. On the other hand, we find a = 8.81 Å and c = 12.59 Å for N2H5PbI3, while a = 8.76 Å and c = 12.52 Å for N2H3PbI3. The calculated band gaps of these compounds, using the modified Becke-Johnson exchange potential with A = 0.4 and B = 1.0 bohr1/2 in Eq.(3) and taking into account spin-orbit coupling, are presented in Table 3. These results show that the replacement of CH3NH3 with N2H5 or N2H3 causes a considerable redshift in the band gap values.
Compound | Band gap (eV) |
N2H5PbBr3 | 1.94 |
N2H3PbBr3 | 1.77 |
N2H5PbI3 | 1.41 |
N2H3PbI3 | 1.13 |
The calculated density of states in N2H5PbI3 is presented in Fig.2. It is noted that, similar to CH3NH3PbI3, the low-lying conduction bands are derived mainly from Pb p orbitals, whereas the highest valence band is composed of both Pb s and I p states. Bands in the energy range -4 eV to -2 eV are derived mostly from iodine p orbitals. The character of the valence and conduction bands is also made clear in Fig.3, where the energy bands along some high-symmetry directions in the first Brillouin zone are plotted. The size of the circles is proportional to the contribution of the chosen atomic orbital to the eigenstates at each k-point. The fact that s (l = 0) and p (l = 1) orbitals on the same atom (Pb) make large contributions to the wave functions at the valence band maximum and conduction band minimum is responsible for the large optical absorption coefficients that occur in these compounds, and hence their usefulness in solar cell applications.
In N2H5PbI3, the valence band maximum (VBM) and conduction band minimum (CBM) occur at the Γ-point, the Brillouin zone center. In ideal cubic perovskites, VBM and CBM occur at the zone’s corner point R(1/2, 1/2, 1/2). Here, N2H5PbI3 is assumed to have a body-centered tetragonal structure with two formula units per primitive cell. The conventional tetragonal unit cell, with four formula units, is a slight distortion of the √2×√2×2 supercell of the ideal cubic unit cell, and point R is zone-folded into point Γ.
Spin-orbit coupling (SOC) has a profound effect on the band structure in organolead halide compounds. In Fig.3, we see that at the Γ point, the lowest conduction band has energy 1.41 eV, while the next two higher bands have energy close to 3 eV.. In the absence of SOC, those three bands would be almost degenerate at the Γ point, and all of them would occur at about 2.5 eV. In a cubic perovskite structure, such as the one found in MAPbBr3, the conduction band minimum at point R is six-fold degenerate (including spin degeneracy); SOC partially lifts the degeneracy, giving rise to a doublet ( j = 1/2) with a lower energy and a quartet ( j = 3/2) with a higher energy. In a body-centered tetragonal structure, CBM occurs at point Γ, degeneracy is now only approximate (it was exact in the cubic structure), and SOC again splits the almost six-fold degenerate level into one lower doublet and two higher doublets.
We have presented calculations on various lead halide compounds using density functional theory with modified Becke-Johnson exchange potential. For the compounds PbCl2, PbBr2, and CH3NH3PbBr3-xClx, for x = 0, 1, 2, and 3, we showed that the calculated band gaps are in excellent agreement with experimental values. We then used this computational method to predict the electronic structure of similar compounds, namely, those that result from the replacement of the methylammonium cation in MAPbBr3 and MAPbI3 with the slightly smaller cations N2H5+ and N2H3+. A significant downshift in the band gap values is predicted to occur as a result of these replacements. In particular, we predict that N2H5PbI3 and N2H3PbI3 have band gaps given by 1.41 eV and 1.13 eV, respectively. Therefore. these compounds, if synthesized, would be excellent light harvesters in solar cells. It should be noted, however, that the instability of the diazenium cation (N2H3+) may make it difficult to use it as a replacement for the methyl ammonium cation.
The author gratefully acknowledges support by National Science Foundation under grant No. HRD- 0932421 and NSF PREM Program: Cal State L.A. & Penn State Partnership for Materials Research and Education, award DMR-1523588.
The author reports no conflict of interest in this research.
[1] |
Kojima A, Teshima K, Shirai Y, et al. (2009) Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J Am Chem Soc 131: 6050–6051. doi: 10.1021/ja809598r
![]() |
[2] |
Etgar L, Gau P, Xue Z, et al. (2012) Mesoscopic CH3NH3PbI3/TiO2 heterojunction solar cells. J Am Chem Soc 134: 17396–17399. doi: 10.1021/ja307789s
![]() |
[3] |
Ball J, Lee M, Hey A, et al. (2013) Low-temperature processed meso-superstructured to thin-film perovskite solar cells. Energy Env Sci 6: 1739–1743. doi: 10.1039/c3ee40810h
![]() |
[4] |
Heo H, Im S, Noh J, et al. (2013) Efficient inorganic-organic hybrid heterojunction solar cells containing perovskite compound and polymeric hole conductors. Nat Photonics 7: 486–491. doi: 10.1038/nphoton.2013.80
![]() |
[5] | Kim H-S, Lee J-W, Yantara N, et al. (2013) High efficiency solid-state sensitized solar cell based on submicrometer rutile TiO2 nanorod and CH3NH3PbI3 perovskite sensitizer. Nano Lett 13: 2412–2417. |
[6] | Bi D, Yang L, Boschloo G, et al. (2013) Effect of different hole transport materials on recombination in CH3NH3PbI3 perovskite-sensitized mesoscopic solar cells. J Phys Chem Lett 4: 1532–1536. |
[7] |
Cai B, Xing Y, Yang Z, et al. (2013) High performance hybrid solar cells sensitized by organolead halide perovskites. Energy Env Sci 6: 1480–1485. doi: 10.1039/c3ee40343b
![]() |
[8] |
Eperon G, Burlakov V, Docampo P, et al. (2014) Morphological control for high performance, solution-processed planar heterojunction perovskite solar cells. Adv Funct Mater 24: 151–157. doi: 10.1002/adfm.201302090
![]() |
[9] | Laban W, Etgar L. (2014) Depleted hole conductor-free lead halide iodide heterojunction solar cells. Energy Env Sci 6: 3249–3253. |
[10] |
Stranks S, Eperon G, Grancini G, et al. (2013) Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 342: 341–344. doi: 10.1126/science.1243982
![]() |
[11] |
Lee M, Teuscher J, Miyasaka T, et al. (2012) Efficient hybrid solar cells based on mesosuperstructured organometal halide perovskites. Science 338: 643–647. doi: 10.1126/science.1228604
![]() |
[12] | Noh J, Im S, Heo J, et al. (2013) Chemical management for colorful, efficient, and stable inorganicorganic hybrid nanostructured solar cells. Nano Lett 13: 1764–1769. |
[13] |
Burschka J, Pellet N, Moon S, et al. (2013) Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 499: 316–319. doi: 10.1038/nature12340
![]() |
[14] |
Liu M, Johnston M, Snaith H (2013) Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature 501: 395–398. doi: 10.1038/nature12509
![]() |
[15] |
Mosconi E, Amat A, Nazeeruddin M, et al. (2013) First-principles modeling of mixed halide organometal perovskites for photovoltaic applications. J Phys Chem C 117: 13902–13913. doi: 10.1021/jp4048659
![]() |
[16] |
Wang Y, Gould T, Dobson J, et al. (2014) Density functional theory analysis of structural and electronic properties of orthorhombic perovskite CH3NH3PbI3. Phys Chem Chem Phys 16: 1424–1429. doi: 10.1039/C3CP54479F
![]() |
[17] | Umari P, Mosconi E, De Angelis, F (2014) Relativistic GW calculations on CH3NH3PbI3 and CH3NH3SnI3 perovskites for solar cell applications. Sci Rep 4: Article number: 4467. |
[18] |
Even J, Pedesseau L, Jancu J, et al. (2013) Importance of spin–orbit coupling in hybrid organic/inorganic perovskites for photovoltaic applications. J Phys Chem Lett 4: 2999–3005. doi: 10.1021/jz401532q
![]() |
[19] |
Even J, Pedesseau L, Dupertuis M, et al. (2012) Electronic model for self-assembled hybrid organic/perovskite semiconductors: reverse band edge electronic states ordering and spin-orbit coupling. Phys Rev B 86: 205301. doi: 10.1103/PhysRevB.86.205301
![]() |
[20] |
Even J, Pedesseau L, Katan C (2014) Comments on “density functional theory analysis of structural and electronic properties of orthorhombic perovskite CH3NH3PbI3.” Phys Chem Chem Phys 16:8697-8698 doi: 10.1039/C3CP55006K
![]() |
[21] |
Feng J, Xiao B (2014) Correction to “crystal structures, optical properties, and effective mass tensors of CH3NH3PbI3 (X=I and Br) phases predicted from HSE06.” J Phys Chem Lett 5: 1719-1720. doi: 10.1021/jz500831m
![]() |
[22] | Brivio F, Butler K, Walsh A (2014) Relativistic quasiparticle self-consistent electronic structure of hybrid halide perovskite photovoltaic absorbers. Phys Rev B 89: 155024 |
[23] | Filippetti A, Mattoni A (2014) Hybrid perovskites for photovoltaics: insights from first principles. Phys Rev B 89: 125203 |
[24] |
Jishi R, Ta O, Sharif A (2014) Modeling of lead halide compounds for photovoltaic applications. J Phys Chem C 118: 28344–28349. doi: 10.1021/jp5050145
![]() |
[25] |
Motta C, El-Mellouhi F, Kais S, et al. (2015) Revealing the role of organic cations in hybrid halide perovskite CH3NH3PbI3. Nat Commun 6: 7026. doi: 10.1038/ncomms8026
![]() |
[26] | Baikie T, Fang Y, Kadro J, et al. (2013) Synthesis and crystal chemistry of the hybrid perovskite (CH3NH3)PbI3 for solid-state sensitised solar cell applications. J Mater Chem A 1: 5628–5641. |
[27] |
Comin R, Walters G, Thibau E, et al. (2015) Structural, optical, and electronic studies of widebandgap lead halide perovskites. J Mater Chem C 3: 8839–8843. doi: 10.1039/C5TC01718A
![]() |
[28] | Buin A, Comin R, Xu J, et al. (2015) Halide-dependent electronic structure of organolead perovskite materials. Chem Mater 27: 4405–4412. |
[29] |
Pang S, Hu H, Zhang J, et al. (2014) NH2CH=NH2PbI3: An alternative organolead iodide perovskite sensitizer for mesoscopic solar cells. Chem Mater 26: 1485–1491. doi: 10.1021/cm404006p
![]() |
[30] | Stoumpos C, Malliakas C, Kanatzidis M (2013) Semiconducting tin and lead iodide perovskites with organic cations: phase transitions, high mobilities, and near-infrared photoluminescent properties. Inorg Chem 52: 9019–9038. |
[31] | Stoumpos C, Kanatzidis G (2015) The renaissance of halide perovskites and their evolution as emerging semiconductors. Acc Chem Res 48: 2791–2802. |
[32] |
Eperon G, Stranks S, Menelaou C, et al. (2014) Formamidinium lead trihalide: a broadly tunable perovskite for efficient planar heterojunction solar cells. Energy Environ Sci 7: 982–988. doi: 10.1039/c3ee43822h
![]() |
[33] |
Koh T, Fu K, Fang Y, et al. (2014) Formamidinium-containing metal halide: an alternative material for near-IR absorption perovskite solar cells. J Phys Chem C 118: 16458–16462. doi: 10.1021/jp411112k
![]() |
[34] |
Jeon N, Noh J, Yang W, et al. (2015) Compositional engineering of perovskite materials for highperformance solar cells. Nature 517: 476–480. doi: 10.1038/nature14133
![]() |
[35] |
Tan Z, Moghaddam R, Lai M, et al. (2014) Bright light-emitting diodes based on organometal halide perovskite. Nature Nanotech 9: 687–692. doi: 10.1038/nnano.2014.149
![]() |
[36] |
Kim Y.-H, Cho H, Heo J, et al. (2015) Multicolored organic/inorganic hybrid perovskite lightemitting diodes. Adv Mater 27: 1248–1254. doi: 10.1002/adma.201403751
![]() |
[37] |
Amat A, Mosconi E, Ronca E, et al. (2014) Cation-induced band-gap tuning in organohalide perovskites: interplay of spin-orbit coupling and octahedra tilting. Nano Lett 14: 3608–3616. doi: 10.1021/nl5012992
![]() |
[38] |
Kieslich G, Sun S, Cheetham A (2015) An extended tolerance factor approach for organic-inorganic perovskites. Chem Sci 6: 3430–3433. doi: 10.1039/C5SC00961H
![]() |
[39] | Mashiyama H, Kurihara Y, Azetsu T (1998) Disordered cubic perovskite structure of CH3NH3PbX3(X = Cl, Br, I). J Korean Phys Soc 32: S156-S158. |
[40] | Becke A (1993) Density-functional thermochemistry. III. The role of exact exchange. J Chem Phys 98: 5648–5652. |
[41] | Frisch M, Trucks G, Schlegel H, et al. (2009) Gaussian 09, Gaussian, Inc: Willingford, CT. |
[42] | Kohn W, Sham L (1965) Self-consistent equations including exchange and correlation effects. Phys Rev 140: A1133–A1138. |
[43] | Perdew J, Burke K, Ernzerhof M (1996) Generalized gradient approximation made simple. Phys Rev Lett 77: 3865–3868. |
[44] |
Bechstedt F, Fuchs F, Kresse G (2009) Ab-initio theory of semiconductor band structures: new developments and progress. Phys Status Solidi B 246: 1877–1892. doi: 10.1002/pssb.200945074
![]() |
[45] | Becke A, Johnson E (2006) A simple effective potential for exchange. J Chem Phys 124: 221101. |
[46] | Tran F, Blaha P (2009) Accurate band gaps of semiconductors and insulators with a semilocal exchange-correlation potential. Phys Rev Lett 102: 226401. |
[47] |
Becke A, Roussel M (1989) Exchange holes in inhomogeneous systems: a coordinate-space model. Phys Rev A 39: 3761–3767. doi: 10.1103/PhysRevA.39.3761
![]() |
[48] | Blaha P, Schwarz K, Madsen G, et al. (2001) WIEN2K: an augmented plane wave + local orbitals program for calculating crystal properties. |
[49] |
Monkhorst H, Pack J (1976) Special points for Brillouin-zone integrations. Phys Rev B 13: 5188–5192. doi: 10.1103/PhysRevB.13.5188
![]() |
[50] |
Perdew J, Ruzsinzky A, Csonka G, et al. (2008) Restoring the density-gradient expansion for exchange in solids. Phys Rev Lett 100: 136406. doi: 10.1103/PhysRevLett.100.136406
![]() |
[51] | Wyckoff R (1963) Crystal structures, 2nd ed. (Wiley, New York) Vol. 1. |
[52] |
Plekhanov V (2004) Lead halides: electronic properties and applications. Prog Mater Sci 49: 787–886. doi: 10.1016/S0079-6425(03)00049-5
![]() |
[53] |
Zaldo C, Sol´ e J, Di ´ eguez E, et al. (1985) Optical spectroscopy of PbCl2 particles embedded in NaCl host matrix. J Chem Phys 83: 6197–6200. doi: 10.1063/1.449859
![]() |
[54] |
Plekhanov V (1973) Optical constants of lead halides. Phys Stat Sol B 57: K55–K59. doi: 10.1002/pssb.2220570157
![]() |
[55] |
Iwanaga M, Watanabe M, Hayashi T (2000) Charge separation of excitons and the radiative recombination process in PbBr2 crystals. Phys Rev B 62: 10766–10773. doi: 10.1103/PhysRevB.62.10766
![]() |
[56] |
Matus M, Arduengo A, Dixon D (2006) The heats of formation of diazene, hydrazine, N2H3+, N2H5+, N2H, and N2H3 and the methyl derivatives CH3NNH, CH3NNCH3, and CH3HNNHCH3. J Phys Chem A 110: 10116–10121. doi: 10.1021/jp061854u
![]() |
1. | Arpita Varadwaj, Pradeep R. Varadwaj, Koichi Yamashita, Halogen in materials design: Fluoroammonium lead triiodide (FNH3PbI3) perovskite as a newly discovered dynamical bandgap semiconductor in 3D, 2018, 118, 00207608, e25621, 10.1002/qua.25621 | |
2. | T. Malsawmtluanga, Benjamin Vanlalruata R. K. Thapa, Investigation of half-metallicity of GeKMg and SnKMg by Using mBJ potential method, 2016, 765, 1742-6588, 012018, 10.1088/1742-6596/765/1/012018 | |
3. | Pradeep R. Varadwaj, Arpita Varadwaj, Helder M. Marques, Koichi Yamashita, Halogen in materials design: Chloroammonium lead triiodide perovskite (ClNH 3 PbI 3 ) a dynamical bandgap semiconductor in 3D for photovoltaics , 2018, 39, 0192-8651, 1902, 10.1002/jcc.25366 | |
4. | Patrik Ščajev, Džiugas Litvinas, Vaiva Soriu̅tė, Gediminas Kreiza, Sandra Stanionytė, Saulius Juršėnas, Crystal Structure Ideality Impact on Bimolecular, Auger, and Diffusion Coefficients in Mixed-Cation CsxMA1–xPbBr3 and CsxFA1–xPbBr3 Perovskites, 2019, 123, 1932-7447, 23838, 10.1021/acs.jpcc.9b05824 | |
5. | M Shakil, Arfan Akram, I Zeba, Riaz Ahmad, S S A Gillani, M Asghar Gadhi, Effect of mixed halide contents on structural, electronic, optical and elastic properties of CsSnI3−xBrx for solar cell applications: first-principles study, 2020, 7, 2053-1591, 025513, 10.1088/2053-1591/ab727d | |
6. | Lung-Chien Chen, Zong-Liang Tseng, Jun-Kai Huang, Cheng-Chiang Chen, Sheng Chang, Fullerene-Based Electron Transport Layers for Semi-Transparent MAPbBr3 Perovskite Films in Planar Perovskite Solar Cells, 2016, 6, 2079-6412, 53, 10.3390/coatings6040053 | |
7. | Patrik Ščajev, Džiugas Litvinas, Gediminas Kreiza, Sandra Stanionytė, Tadas Malinauskas, Roland Tomašiūnas, Saulius Juršėnas, Highly efficient nanocrystalline CsxMA1−xPbBrx perovskite layers for white light generation, 2019, 30, 0957-4484, 345702, 10.1088/1361-6528/ab1a69 | |
8. | Ala’a O. El-Ballouli, Osman M. Bakr, Omar F. Mohammed, Compositional, Processing, and Interfacial Engineering of Nanocrystal- and Quantum-Dot-Based Perovskite Solar Cells, 2019, 31, 0897-4756, 6387, 10.1021/acs.chemmater.9b01268 | |
9. | Fabien Tran, Jan Doumont, Leila Kalantari, Ahmad W. Huran, Miguel A. L. Marques, Peter Blaha, Semilocal exchange-correlation potentials for solid-state calculations: Current status and future directions, 2019, 126, 0021-8979, 110902, 10.1063/1.5118863 | |
10. | Yasir Saeed, Bin Amin, Haleema Khalil, Fida Rehman, Hazrat Ali, M. Imtiaz Khan, Asif Mahmood, M. Shafiq, Cs2NaGaBr6: a new lead-free and direct band gap halide double perovskite, 2020, 10, 2046-2069, 17444, 10.1039/D0RA01764G | |
11. | Markus Becker, Thorsten Klüner, Michael Wark, Formation of hybrid ABX3perovskite compounds for solar cell application: first-principles calculations of effective ionic radii and determination of tolerance factors, 2017, 46, 1477-9226, 3500, 10.1039/C6DT04796C | |
12. | Priyanka Samanta, Yitang Wang, Shadi Fuladi, Jinjing Zou, Ye Li, Le Shen, Christopher Weber, Fatemeh Khalili-Araghi, Molecular determination of claudin-15 organization and channel selectivity, 2018, 150, 0022-1295, 949, 10.1085/jgp.201711868 | |
13. | Razieh Beiranvand, Vahid Mehrabi, Structural, electronic and optical properties of bulk and monolayer iron diselenide: A density functional study, 2021, 96, 0031-8949, 065803, 10.1088/1402-4896/abec01 | |
14. | Tudor Luca Mitran, Rachel Elizabeth Brophy, Marina Cuzminschi, Nicolae Filipoiu, Movaffaq Kateb, Ioana Pintilie, Andrei Manolescu, George Alexandru Nemnes, 2023, 9780323885225, 153, 10.1016/B978-0-323-88522-5.00012-0 | |
15. | Zhirong Liu, Zhiyong Liu, Qiang Sun, Tao Zhang, Haixuan Yu, Xuning Zhang, Letian Dai, Guanglan Liao, Yan Shen, Xiao-Li Zhang, Jun Zhu, Mingkui Wang, A stable self-powered ultraviolet photodetector using CH3NH3PbCl3 with weak-light detection capacity under working conditions, 2022, 10, 2050-7526, 7147, 10.1039/D2TC00637E | |
16. | M. Zia ur Rehman, Shaimaa A.M. Abdelmohsen, Eman A. Mahmoud, M. Usman Saeed, M. Idress, M. Shafiq, B. Amin, Y. Saeed, First principles study of structural, electronic, elastic and optical properties of Cs2LiTlBr6 and Cs2NaTlBr6, 2022, 151, 13698001, 106993, 10.1016/j.mssp.2022.106993 | |
17. |
M. Usman Saeed, Tayyba Usman, Sardar Mohsin Ali, Shamiala Pervaiz, Hosam O. Elansary, Ihab Mohamed Moussa, Mohamed A. El-Sheikh, Aziz-Ur-Rahim Bacha, Y. Saeed,
Exploring electronic, optical, elastic, and photocatalytic properties in new double perovskites Cs2TlSbCl6 and Cs2TlBiCl6 materials: A GGA+SOC and hybrid functional study,
2024,
0884-2914,
10.1557/s43578-024-01449-8
|
|
18. | Laraib Sajid, M. Usman Saeed, S. H. Mashadi, S. Sheryar Abid, Shamiala Pervaiz, Zeeshan Ali, Yousef Mohammed Alanazi, Aziz-Ur-Rahim Bacha, Y. Saeed, Ab initio study of electronic, elastic, thermodynamic, photocatalytic properties of double antiperovskite, Cs6AgBiX2 (X = Cl, Br, I), 2024, 14, 2046-2069, 35348, 10.1039/D4RA05661B | |
19. | Mohammad Tanvir Ahmed, Shariful Islam, Farid Ahmed, A‐Site Cation Replacement of Hydrazinium Lead Iodide Perovskites by Borane Ammonium Ions: A DFT Calculation, 2024, 13, 2191-1363, 10.1002/open.202300207 | |
20. | Banat Gul, Muhammad Salman Khan, Muhammad Aasim, Ahmad A. lfseisi, Gulzar Khan, Hijaz Ahmad, First-Principles Investigation of Novel Alkali-Based Lead-Free Halide Perovskites for Advanced Optoelectronic Applications, 2023, 8, 2470-1343, 32784, 10.1021/acsomega.3c03756 | |
21. | Banat Gul, Muhammad Salman Khan, Muhammad Aasim, Gulzar Khan, Hijaz Ahmad, Phatiphat Thounthong, First-principles study of the optoelectronic and thermoelectric properties of lead-free ASnI3 (A = K, Rb, and Cs) novel halide perovskites, 2023, 669, 09214526, 415316, 10.1016/j.physb.2023.415316 |
Compound | Structure | Lattice constants (Å) |
PbCl2 | Orthorhombic | a=9.03, b=7.608, c=4.525 |
PbBr2 | Orthorhombic | a=9.466, b=8.068, c=4.767 |
MAPbBr3 | Cubic | a=5.933 |
MAPbBr2Cl | Cubic | a=5.88 |
MAPbBrCl2 | Cubic | a=5.78 |
MAPbCl3 | Cubic | a=5.71 |
Compound | Band gap (eV) |
N2H5PbBr3 | 1.94 |
N2H3PbBr3 | 1.77 |
N2H5PbI3 | 1.41 |
N2H3PbI3 | 1.13 |
Compound | Structure | Lattice constants (Å) |
PbCl2 | Orthorhombic | a=9.03, b=7.608, c=4.525 |
PbBr2 | Orthorhombic | a=9.466, b=8.068, c=4.767 |
MAPbBr3 | Cubic | a=5.933 |
MAPbBr2Cl | Cubic | a=5.88 |
MAPbBrCl2 | Cubic | a=5.78 |
MAPbCl3 | Cubic | a=5.71 |
Compound | GGA+SOC | mBJ+SOC | Experiment |
PbCl2 | 3.18 | 5.13 | 5.38, [52] 4.86 [53] |
PbBr2 | 2.46 | 4.19 | 4.23, [54] 4.1 [55] |
MAPbBr3 | 0.45 | 2.23 | 2.28 [12] |
MAPbBr2Cl | 0.46 | 2.42 | 2.65 [27] |
MAPbBrCl2 | 0.57 | 2.76 | 2.90 [27] |
MAPbCl3 | 1.0 | 3.22 | 3.17 [27] |
Compound | Band gap (eV) |
N2H5PbBr3 | 1.94 |
N2H3PbBr3 | 1.77 |
N2H5PbI3 | 1.41 |
N2H3PbI3 | 1.13 |