On the basis of the screened 29 hybrid halide compounds from our prior study [Y. Li and K. Yang, Energy Environ. Sci. 12, 2233–2243 (2019)], here, we reported a systematic computational study of the stability diagram, defect tolerance, and optical absorption coefficients for these candidate materials using high-throughput first-principles calculations. We took two exemplary compounds, MA2SnI4 and MA3Sb2I9, as examples to show the computational process, and they are discussed in detail. This work is expected to provide a detailed guide for further experimental synthesis and characterization, with the potential to develop novel lead-free optoelectronic devices.

Organic–inorganic hybrid halide perovskites have emerged as one important class of semiconductor materials for various types of optoelectronic applications, such as solar cells and light-emitting diodes, because of their low-temperature solution-based fabrication techniques and excellent optoelectronic properties.1–6 As a proof of example, the power conversion efficiency (PCE) of the perovskite-based solar cells at lab-scale testing has increased rapidly from 3.8% to 25.2% in the past few years.7 This is mainly attributed to their exceptional optoelectronic properties such as tunable bandgaps, high absorption coefficient, long carrier diffusion length and lifetime, strong defect tolerance, and high carrier mobility.8–14 

Despite the exceptional properties, it is still of urgent demand to search for stable and non-toxic alternative hybrid materials because of low stability and the presence of toxic lead in the halide perovskite.6 Some great research efforts have been made to search for new halide perovskite materials using the high-throughput material design approach,15–24 although nearly all of them are focusing on either purely inorganic perovskites or a limited number of hybrid perovskite structures (cubic structure or its distorted derivatives). This is mainly due to great computational challenges caused by the structural complexity of the hybrid structures in the large-scale first-principles calculations. It is also worth noting that quantitative physical properties of hybrid perovskites cannot be directly derived from their inorganic bulk analogs.

In our recent work, instead of focusing on the prototype perovskite structure, we have carried out a systematic investigation of all the possible hybrid halide materials on the basis of the 24 prototype structures and five relatively small organic cations using the high-throughput computational material design approach.25 By employing a group of combinatorial material descriptors that cover the critical features of electronic band structures, several energetic parameters, structural integrity at room-temperature, and exciton binding energies, we have successfully identified 13 candidates for solar energy conversion and 23 candidates for light-emitting diodes (a total number of 29 hybrid compounds) in five different types of crystal structures out of a quantum material repository of 4507 hybrid halide materials.25 

In this work, to provide a more detailed guide for the experimental synthesis and characterization, we systemically studied the stability diagrams, defect tolerance, and absorption coefficients for the identified 29 promising hybrid halide materials. To do this, we first calculated thermodynamically stable ranges for all these compounds and considered non-equivalent lattice sites in each compound for building defect structures. A total number of about 5000 defect structures (including neutral and charged) were computed to analyze the defect transition energy levels using high-throughput first-principles calculations. The absorption coefficients were calculated from the dielectric function. For convenience, two representative compounds, MA2SnI4 and MA3Sb2I9, were selected for discussing the computational and analysis process. The calculated stability diagrams, defect formation energies and transition levels, and absorption coefficients for all the 29 compounds are presented in the supplementary material.

The automatic framework AFLOW26 based on the Vienna Ab initio Simulation Package (VASP)27 was used for high-throughput first-principles Density Functional Theory (DFT) calculations. The Projector Augmented Wave (PAW) pseudopotentials were used for describing electron–ion interactions,28 and the Generalized Gradient Approximation (GGA) parameterized by Perdew, Burke, and Ernzerhof (PBE) was used for treating electron–electron exchange-correlation functional.29 The van der Waals (vdW) functional DFT-D3 is incorporated to properly describe the long-range dispersion interactions between the organic molecules.30 A k-point grid of 0.05 Å−1 was automatically set for structural relaxation with a convergence tolerance of 0.01 meV/atom, and a denser grid of 0.04 Å−1 was used in static calculations. Other computational settings such as the cutoff energy (the highest value of the pseudopotentials) were managed by the AFLOW code that generates appropriate entries for the structural relaxation and static calculations sequentially and automatically.26 The supercell model (with more than 40 atoms) was built for each material system to study defect properties, in which the supercell size was determined based on the convergence tests for the formation energy of iodine vacancy; see Fig. A1 and Table A1 in the supplementary material. Our calculations indicate that the supercell with more than 40 atoms is large enough to produce converged defect formation energy. Similar conclusions can also be found in the previous literature on perovskite oxides.31,32 The hybrid DFT calculations within the Heyd–Scuseria–Ernzerhof (HSE) formalism with 25% Hartree–Fock (HF) exchange are employed to calculate the absorption coefficients.33,34

All the possible intrinsic point defects, including vacancies (VA, VB, and VX) and interstitial (Ai, Bi, and Xi) and antisite (AB, BA, AX, XA, BX, and XB) defects, were considered for all the 29 hybrid halide compounds AxByXz; see the list of compounds in Table I. The defect structures were built based on all the possible non-equivalent lattice sites in each compound. The Python Charge Defects Toolkit (PyCDT) based on the Interstitial Finding Tool (InFiT) was used to determine possible interstitial sites.35,36 The building process yields a total number of 2160 defect structures. After first-principles total energy calculations, we selected the defect structure with the lowest total energy for the same type of defects for the further investigation of defect transition levels. This process leads to a total number of 348 selected structures, which are the lowest-energy site representatives for 12 intrinsic point defects in each of the 29 compounds and used for charged defect calculations.

TABLE I.

List of 29 candidate hybrid halide semiconductor materials. Calculated properties: compound, Pearson symbol, lattice parameters (Å), bandgap Eg (eV) using hybrid functional calculations (* marks calculations with spin-orbit coupling), absorption coefficient α (×105 cm−1) at 450 nm, and defect tolerance level. The lattice parameters and bandgaps are adapted from Ref. 25.

Lattice parameters
CompoundPearson symbolabcEgαDefect tolerance
(MA)2GeBr4 tI14 5.54 5.54 19.02 1.99 1.10 High 
(MA)2GeI4 tI14 5.94 5.87 20.11 1.66 1.95 High 
(MA)2SnCl4 tI14 5.56 5.48 18.35 2.49 0.41 Low 
(MA)2SnBr4 tI14 5.72 5.74 19.08 1.67 1.61 High 
(MA)2SnI4 tI14 6.12 6.11 20.00 1.42 2.43 High 
(FA)2SnBr4 tI14 5.64 5.71 20.60 1.69 1.58 High 
(AD)2GeI4 tI14 6.14 6.16 18.97 2.30 1.12 Low 
(AD)2SnBr4 tI14 5.87 5.86 18.09 2.51 0.84 Low 
(AD)2SnI4 tI14 6.22 6.21 19.17 1.87 1.89 High 
(MA)3In2I9 hP14 8.23 8.31 11.00 2.18 0.22 Low 
(MA)3Sb2Br9 hP14 7.71 7.71 10.37 2.70 0.49 Low 
(MA)3Sb2I9 hP14 8.26 8.26 10.91 2.02 3.10 High 
(MA)3Bi2I9 hP14 8.35 8.35 11.00 1.82* 0.63 High 
(FA)3Ga2I9 hP14 7.94 8.25 11.83 2.17 0.08 Low 
(FA)3In2Br9 hP14 7.38 7.83 11.17 2.91 0.03 High 
(FA)3In2I9 hP14 8.04 8.30 11.67 1.90 0.30 Low 
(FA)3Bi2I9 hP14 8.06 8.36 11.60 1.80* 0.86 Low 
(AD)3Sb2Br9 hP14 8.30 8.36 9.55 2.52 0.41 Low 
(AD)3Sb2I9 hP14 8.69 8.77 10.20 1.91 2.90 High 
(AD)3Bi2I9 hP14 8.78 8.84 10.26 1.80* 0.55 High 
(FA)3Sb2I9 hP28 8.02 8.32 23.18 2.54 0.53 Low 
(AD)3In2I9 hP28 8.61 8.71 20.54 2.00 0.51 Low 
(MA)2ZrI6 tI18 8.18 8.19 12.37 2.60 0.39 High 
(AD)2HfI6 tI18 8.47 8.46 12.17 2.89* 0.08 Low 
(AD)2SnBr6 tI18 8.13 8.12 11.56 2.42 0.06 High 
(AD)2TeBr6 tI18 8.29 8.30 11.18 2.71 0.20 High 
(MA)2SnI6 cF36 11.61 11.44 12.55 0.80 0.60 High 
(MA)2TeI6 cF36 11.67 11.51 12.58 1.77 2.93 Low 
(AD)2TeI6 cF36 12.54 11.63 12.17 1.87 2.65 Low 
Lattice parameters
CompoundPearson symbolabcEgαDefect tolerance
(MA)2GeBr4 tI14 5.54 5.54 19.02 1.99 1.10 High 
(MA)2GeI4 tI14 5.94 5.87 20.11 1.66 1.95 High 
(MA)2SnCl4 tI14 5.56 5.48 18.35 2.49 0.41 Low 
(MA)2SnBr4 tI14 5.72 5.74 19.08 1.67 1.61 High 
(MA)2SnI4 tI14 6.12 6.11 20.00 1.42 2.43 High 
(FA)2SnBr4 tI14 5.64 5.71 20.60 1.69 1.58 High 
(AD)2GeI4 tI14 6.14 6.16 18.97 2.30 1.12 Low 
(AD)2SnBr4 tI14 5.87 5.86 18.09 2.51 0.84 Low 
(AD)2SnI4 tI14 6.22 6.21 19.17 1.87 1.89 High 
(MA)3In2I9 hP14 8.23 8.31 11.00 2.18 0.22 Low 
(MA)3Sb2Br9 hP14 7.71 7.71 10.37 2.70 0.49 Low 
(MA)3Sb2I9 hP14 8.26 8.26 10.91 2.02 3.10 High 
(MA)3Bi2I9 hP14 8.35 8.35 11.00 1.82* 0.63 High 
(FA)3Ga2I9 hP14 7.94 8.25 11.83 2.17 0.08 Low 
(FA)3In2Br9 hP14 7.38 7.83 11.17 2.91 0.03 High 
(FA)3In2I9 hP14 8.04 8.30 11.67 1.90 0.30 Low 
(FA)3Bi2I9 hP14 8.06 8.36 11.60 1.80* 0.86 Low 
(AD)3Sb2Br9 hP14 8.30 8.36 9.55 2.52 0.41 Low 
(AD)3Sb2I9 hP14 8.69 8.77 10.20 1.91 2.90 High 
(AD)3Bi2I9 hP14 8.78 8.84 10.26 1.80* 0.55 High 
(FA)3Sb2I9 hP28 8.02 8.32 23.18 2.54 0.53 Low 
(AD)3In2I9 hP28 8.61 8.71 20.54 2.00 0.51 Low 
(MA)2ZrI6 tI18 8.18 8.19 12.37 2.60 0.39 High 
(AD)2HfI6 tI18 8.47 8.46 12.17 2.89* 0.08 Low 
(AD)2SnBr6 tI18 8.13 8.12 11.56 2.42 0.06 High 
(AD)2TeBr6 tI18 8.29 8.30 11.18 2.71 0.20 High 
(MA)2SnI6 cF36 11.61 11.44 12.55 0.80 0.60 High 
(MA)2TeI6 cF36 11.67 11.51 12.58 1.77 2.93 Low 
(AD)2TeI6 cF36 12.54 11.63 12.17 1.87 2.65 Low 

For charged defect calculations, we considered defect charge ranges derived from all the possible oxidation states of involved element(s) and the defect type based on the selected 348 defect structures. In terms of defect type, vacancies of atom An (n is the formal charge of A in the compound AxByXz) can have charge states in the range of [−n, +n]; interstitial Am (m represents all the possible oxidation states of A) can have charge states in the range of [min(m), max(m)]; and antisite AB have charge ranges of [min(p), max(p)], where p represents all possible oxidation states of A and B.35 According to the above charge ranges, a total number of about 2400 charged defect calculations were to be carried out.

It is noted that, in the charged defects calculations under the periodic boundary conditions, the Coulombic interaction between a charged defect and its periodic images is the dominant source of error for defect energetics. The interaction is inversely proportional to the supercell periodic length. The supercells used in the high-throughput calculations are finitely sized, so the charged defect images cannot be effectively isolated. Thus, the Coulombic interaction cannot be completely eliminated either. To correct supercell finite-size effects in the charged defects calculations, we employed the image charge correction proposed by Freysoldt et al. in our calculations by calling the pymatgen package.36,37

We first computed stability diagrams of the 29 hybrid halide compounds by determining their thermodynamically stable chemical ranges with respect to their decomposers. Note that the dynamic stability, as a more realistic descriptor for screening novel materials than the thermodynamic stability, has been investigated in our prior study using ab initio molecular dynamics (AIMD) simulations.25 Here, the calculated thermodynamic stability diagrams have two major purposes: (i) to determine the chemical potential ranges for synthesizing the hybrid materials and (ii) to choose appropriate chemical potentials for subsequent defect formation energy calculations. The first representative compound, MA2SnI4, a tetragonal compound consisting of layers of corner-sharing SnI6, was taken as an example to discuss the computational process; see Fig. 1(a). Under thermodynamic equilibrium growth conditions, the formation of MA2SnI4 should satisfy the following equation:

2ΔμMA+ΔμSn+4ΔμI=ΔH(MA2SnI4),
(1)

in which ΔμX = μXEX is the chemical potential change in the X component (organic molecule MA was treated as one component) and ΔH is the formation enthalpy. The total energy of MA species (EMA) was calculated using a standalone neutral MA molecule in a big box, which represents the gaseous phase as the reference ground state for the species. In addition, to prevent the formation of secondary phases MAI, SnI2, and SnI4, the following stability limits should also be satisfied:

ΔμMA+ΔμI<ΔH(MAI),
(2)
ΔμSn+2ΔμI<ΔH(SnI2),
(3)
ΔμSn+4ΔμI<ΔH(SnI4).
(4)

By applying these thermodynamic conditions, we plotted the stability diagram for MA2SnI4 against ΔμSn and ΔμI, as shown in Fig. 1(b). The green region is the thermodynamically stable range for MA2SnI4. In this region, we have selected three representative points: A (ΔμSn = −4.41 eV, ΔμI = 0 eV), B (ΔμSn = −1.56 eV, ΔμI = −0.75 eV), and C (ΔμSn = 0 eV, ΔμI = −2.2 eV), which represent chemical potential conditions of X-rich/B-poor, X-moderate/B-moderate, and X-poor/B-rich, respectively.

FIG. 1.

Schematic illustration of (a) crystal structure, (b) stability diagram, and [(c) and (d)] defect properties of MA2SnI4. In the stability diagram, the green region depicts the thermodynamically stable range for the equilibrium growth of MA2SnI4 under different Sn and I chemical potentials; outside this region, the compound decomposes into MAI, SnI2, or SnI4. Three representative points A (ΔμSn = −4.41 eV, ΔμI = 0 eV), B (ΔμSn = −1.56 eV, ΔμI = −0.75 eV), and C (ΔμSn = 0 eV, ΔμI = −2.2 eV) are selected to calculate the defect formation energy. (c) Defect formation energies as a function of the Fermi level at chemical potentials A, B, and C. (d) Defect transition energy levels, in which gray levels indicate the metastable charge states and red (blue) lines indicate the donor (acceptor) levels. The dashed lines in (c) and (d) indicate high defect formation energies. Note that some defects present in (c) but absent in (d) mean that these defects do not create transition energy levels within the bandgap or near the band edges and that the defects present in (d) but absent in (c) mean that they have high formation energies at all the considered chemical potential conditions A, B, and C. In compound MA2SnI4, the defects creating deep transition levels include SnI, MAI, VSn, and VMA, which, however, have high formation energies at chemical potential condition B, indicating a high defect tolerance.

FIG. 1.

Schematic illustration of (a) crystal structure, (b) stability diagram, and [(c) and (d)] defect properties of MA2SnI4. In the stability diagram, the green region depicts the thermodynamically stable range for the equilibrium growth of MA2SnI4 under different Sn and I chemical potentials; outside this region, the compound decomposes into MAI, SnI2, or SnI4. Three representative points A (ΔμSn = −4.41 eV, ΔμI = 0 eV), B (ΔμSn = −1.56 eV, ΔμI = −0.75 eV), and C (ΔμSn = 0 eV, ΔμI = −2.2 eV) are selected to calculate the defect formation energy. (c) Defect formation energies as a function of the Fermi level at chemical potentials A, B, and C. (d) Defect transition energy levels, in which gray levels indicate the metastable charge states and red (blue) lines indicate the donor (acceptor) levels. The dashed lines in (c) and (d) indicate high defect formation energies. Note that some defects present in (c) but absent in (d) mean that these defects do not create transition energy levels within the bandgap or near the band edges and that the defects present in (d) but absent in (c) mean that they have high formation energies at all the considered chemical potential conditions A, B, and C. In compound MA2SnI4, the defects creating deep transition levels include SnI, MAI, VSn, and VMA, which, however, have high formation energies at chemical potential condition B, indicating a high defect tolerance.

Close modal

Similarly, the stability limits and stability diagram of the second example compound MA3Sb2I9 were calculated and plotted, as shown in Figs. 2(a) and 2(b). The three representative Δμ points for MA3Sb2I9 are A (ΔμSb = −2.73 eV, ΔμI = 0 eV), B (ΔμSb = −1.04 eV, ΔμI = −0.35 eV), and C (ΔμSb = 0 eV, ΔμI = −0.91 eV). The calculated stability diagrams for all the 29 compounds, plotted against the chemical potential change ΔμB and ΔμX, are shown in Figures B1–B5 of the supplementary material. To deduce all the possible thermodynamic stability limits for each of the 29 compounds, we have carefully considered all the possible neighboring secondary phases on the basis of the open quantum material repositories AFLOWLIB38 and Materials Project.39 The binary halide compounds BXn for B elements in the 29 AxByXz compounds were systemically recalculated in the DFT-D3 approach.

FIG. 2.

Schematic illustration of (a) crystal structure, (b) stability diagram, and [(c) and (d)] defect properties of MA3Sb2I9. Three representative points A (ΔμSb = −2.73 eV, ΔμI = 0 eV), B (ΔμSb = −1.04 eV, ΔμI = −0.35 eV), and C (ΔμSb = 0 eV, ΔμI = −0.91 eV) are selected to calculate the defect formation energy. (c) Defect formation energies as a function of the Fermi level at chemical potentials A, B, and C. (d) Defect transition energy levels. In compound MA3Sb2I9, only two defects, SbI and Sbi, can create deep transition levels, but they have high formation energies at chemical potential condition A, indicating a high defect tolerance. At chemical potential conditions B and C, SbI has a relatively low formation energy, and thus, its deep transition levels could be introduced, though.

FIG. 2.

Schematic illustration of (a) crystal structure, (b) stability diagram, and [(c) and (d)] defect properties of MA3Sb2I9. Three representative points A (ΔμSb = −2.73 eV, ΔμI = 0 eV), B (ΔμSb = −1.04 eV, ΔμI = −0.35 eV), and C (ΔμSb = 0 eV, ΔμI = −0.91 eV) are selected to calculate the defect formation energy. (c) Defect formation energies as a function of the Fermi level at chemical potentials A, B, and C. (d) Defect transition energy levels. In compound MA3Sb2I9, only two defects, SbI and Sbi, can create deep transition levels, but they have high formation energies at chemical potential condition A, indicating a high defect tolerance. At chemical potential conditions B and C, SbI has a relatively low formation energy, and thus, its deep transition levels could be introduced, though.

Close modal

Defect tolerance significantly influences the optoelectronic properties of semiconductor materials and can be evaluated from the defect transition levels and the defect formation energy. A strong defect tolerance can prevent the undesired recombination of photoexcited electrons and holes, and it has the following features: the intrinsic defects with a low formation energy will not create deep gap states, while those producing deep gap states have a relatively high formation energy.9,25 Formation energy (Ef) of a defect X at the charge state q can be calculated using the following equation:40 

Ef[Xq]=Etot[Xq]Etot[bulk]iniμi+qEF+Ecorr,
(5)

in which Etot[Xq] and Etot[bulk] are total energies of defected and pristine supercells, respectively. ∑niμi is the sum of chemical potentials to compensate in the defected supercell, EF is the Fermi energy, and Ecorr is a correction term for electrostatic interactions between supercell images. Here, we only considered defect charge states derived from involved elements’ oxidation states in the compound, which gives upper and lower limits of defects’ charge states. All the charge states between these limits are plotted in the formation energy plots as a function of Fermi energy. This consideration leads to a convenient visualization of the most probable charge states and avoidance of redundancy in the plots, which is also well justified by a prior computational study of point defects in the hybrid halide systems.9 By taking iodine as one example, one only needs to consider its oxidation state of 1− in the hybrid iodides. As a result, an interstitial iodine (Ii) can have two possible charge states of 0 and 1-, and one iodine vacancy (VI) can have two possible charge states of 0 and 1+. The reason why iodine’s positive oxidation states are not considered is that there are no elements oxidizing iodine in the hybrid metal iodide materials.

Figures 1(c) and 2(c) show the formation energies of intrinsic point defects under chemical potential conditions A, B, and C for MA2SnI4 and MA3Sb2I9, respectively. The lower Ef are shown in colored solid lines, while higher Ef are shown in black dashed lines. The slope of the line represents charge q, so horizontal lines show Ef of neutral defects (q = 0). Ef lines change as the ∑niμi term changes among A, B, and C, which is in agreement with chemical condition changes in the involved elements. For example, as the chemical condition becomes more I-rich, Ef of VI and Ii increases and decreases, respectively. Accordingly, the intersection points between different charge states of a defect correspond to the same EF. As shown in Fig. 1(c), the defects with relatively low Ef in MA2SnI4 are VSn, ISn, and MASn at A; VI and Sni at B; SnI, MAI, MAi, VI, and Sni at C.

The transition energy level (ε(q1/q2)) of a defect is the Fermi level position that corresponds to the intersection point of two different charge states (q1 and q2) of a defect in the Ef plots, where the formation energies of the two charge states are equal. In other words, ε(q1/q2) is the energy level where the defect can accept or donate electrons so that the charge states of a defect can be changed. As shown in Fig. 1(c), the lowest Ef of VSn (green line) is composed of two parts, i.e., VSn0 (left) and VSn2 (right), and the two parts intersect at Ef = 1.02 eV. At this transition energy level, VSn0 accepts two electrons and becomes VSn2. On the basis of Eq. (5), ε(q1/q2) can be calculated using the following equation:40 

ε(q1/q2)=Ef(Xq1;EF=0)Ef(Xq2;EF=0)q2q1,
(6)

where Ef(Xq; EF = 0) is the formation energy of defect Xq, when the Fermi level is at 0 [valence band (VB) maximum]. One defect with ε(q1/q2) close to the middle of the bandgap has a deep transition level that will attract electrons/holes and acts as undesired nonradiative recombination centers. In contrast, defects with ε(q1/q2) close to band edges or at the inside of the conduction band (CB) or valence band (VB) will not significantly affect the optoelectronic performance.

The calculated transition levels of MA2SnI4 and MA3Sb2I9 are shown in Figs. 1(d) and 2(d), respectively. In the ε(q1/q2) figures, the red and blue levels indicate intrinsic donors and acceptors, and the top and bottom gray regions indicate CB and VB, respectively. For MA2SnI4 (a bandgap of 1.42 eV), as shown in Fig. 1(d), the defects that induce deep transition levels are SnI (ε(1+/2+) = 1.04 eV and ε(2+/3+) = 0.52 eV), MAI (ε(0/1+) = 1.03 eV and ε(1+/2+) = 0.22 eV), VSn (ε(0/2−) = 1.02 eV), IMA (ε(0/1− = 1.06 eV), and VMA (ε(0/1−) = 0.74 eV). Among these defects, IMA has very high formation energies at all chemical conditions, as indicated by the dashed lines, and, thus, will not lead to deep gap states, indicating a high defect tolerance. SnI (yellow) has a high formation energy at chemical conditions A and B, MAI (orange) has a high formation energy at chemical conditions A and B, VSn (green) has a high formation energy at chemical conditions B and C, and VMA (lime) has a relatively high formation energy at all chemical conditions; see Fig. 1(c). Therefore, the best chemical condition to synthesize MA2SnI4 is at B (I-moderate/Sn-moderate), under which undesired Shockley–Read–Hall (SRH) nonradiative recombination centers can be largely prevented in the material.

For MA3Sb2I9 (bandgap= 2.02 eV), as shown in Fig. 2(d), the defects that create deep transition levels are SbI [ε(2+/4+) = 0.86 eV] and Sbi [ε(1+/2+) = 1.10 eV and ε(2+/3+) = 0.24 eV]. Both SbI (yellow) and Sbi (brown) have high formation energies at chemical condition A, but very low formation energies at B and C. Therefore, to prevent the deep transition levels in MA3Sb2I9, the material should be synthesized under chemical condition A (I-rich/Sb-poor).

The calculated defect formation energies and transition energy levels for all the 29 compounds are shown in Figs. C1–C29 of the supplementary material. To determine the defect tolerance for each compound, we listed the defects that have low formation energies and, meanwhile, can create deep transition levels in Tables D1–D5 of the supplementary material. The defect tolerance was determined based on the three chemical conditions A, B, and C and also summarized in Table I. In specific, for one compound, if there is at least one chemical condition that prevents all kinds of defects mentioned above, the compound is determined to have high defect tolerance; if such defects cannot be prevented at any chemical condition, the compound is determined to have low defect tolerance. From this perspective, there are 15 out of 29 compounds that show high defect tolerance.

In this section, we investigated the optical absorption coefficients using the following equation:41 

α(ω)=2ωcε1(ω)2+ε2(ω)2ε1(ω),
(7)

where ω is the photon frequency and c is the speed of light in vacuum. The real part ε1(ω) and imaginary part ε2(ω) are from the complex dielectric function ε(ω). The real part ε1(ω) follows the Kramer–Kronig relationship, and the imaginary part ε2(ω) is calculated from the appropriate momentum matrix elements between the occupied and the unoccupied wave functions within the selection rules over the Brillouin zone. On the basis of the ground-state electronic structure calculations at the HSE06 level, we obtained ε2(ω) and ε1(ω) of the dielectric function and calculated the optical absorption coefficient of the screened 29 compounds.

The calculated absorption coefficients as a function of wavelength (nm) for the two representative compounds, MA2SnI4 and MA3Sb2I9, are shown in Figs. 3(a) and 3(b), respectively. The averaged (non-polarized) absorption coefficients at 450 nm are estimated to be 2.4 × 105 cm−1 for MA2SnI4 and 3.1 × 105 cm−1 for MA3Sb2I9, even higher than 2 × 105 cm−1 for MAPbI3. Interestingly, MA2SnI4 shows a much higher absorption coefficient along the in-plane (x and y) direction than that along the out-of-plane (z) direction; see Fig. 3(a). The anisotropy of the calculated absorption coefficients resembles that of the two-dimensional Ruddlesden–Popper perovskites such as Ge-based compound in our prior work, which is attributed to the two-dimensional structural feature.14 However, it is noted that in these quantum well-like perovskites, local fields and excitonic effects have major impacts on their optical properties. To accurately describe optical properties of these materials, DFT calculations with GW approximation and the Bethe–Salpeter equation (BSE) are required,42,43 which, however, is more time-consuming. Therefore, to have a direct comparison with the well-studied MAPbI3 at the same computational level, the ground-state HSE06 calculations were adopted in this work. All the calculated absorption coefficients for the 29 compounds are shown in Figs. E1–E5 of the supplementary material.

FIG. 3.

Calculated optical absorption coefficients (α) of (a) MA2SnI4 and (b) MA3Sb2I9 with hybrid functional theory calculations at the HSE06 level.

FIG. 3.

Calculated optical absorption coefficients (α) of (a) MA2SnI4 and (b) MA3Sb2I9 with hybrid functional theory calculations at the HSE06 level.

Close modal

In summary, by using high-throughput first-principles electronic structure calculations, we have systematically studied the stability diagrams, defect tolerance, and absorption coefficients for the screened 29 lead-free candidate hybrid semiconductors for optoelectronic applications. The calculated stability diagram outlines the thermodynamically stable range for the equilibrium growth of the predicted compound with different chemical potentials. The defect tolerance was evaluated from the calculated defect formation energies and transition levels, and the absorption coefficients were calculated from dielectric functions. These computational studies provide a detailed guide to the further experimental synthesis and characterization of these hybrid compounds, with a potential to facilitate the development of novel optoelectronic devices.

See the supplementary material for the complete phase diagram, defect formation energy, transition energy level, defect tolerance, and absorption coefficient for all 29 candidate compounds studied.

This work was supported by the Academic Senate General Campus Research Grant Committee at the University of California, San Diego, the National Science Foundation under Award No. ACI-1550404, and the Global Research Outreach (GRO) Program of Samsung Advanced Institute of Technology under Award No. 20164974. This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by the National Science Foundation, Grant No. ACI-1548562. D.M.-L. and V.R.-V. acknowledge the ENLACE bi-national summer research program at the University of California, San Diego.

1.
A.
Kojima
,
K.
Teshima
,
Y.
Shirai
, and
T.
Miyasaka
, “
Organometal halide perovskites as visible-light sensitizers for photovoltaic cells
,”
J. Am. Chem. Soc.
131
,
6050
6051
(
2009
).
2.
M.
Grätzel
, “
The light and shade of perovskite solar cells
,”
Nat. Mater.
13
,
838
842
(
2014
).
3.
N. J.
Jeon
,
J. H.
Noh
,
W. S.
Yang
,
Y. C.
Kim
,
S.
Ryu
,
J.
Seo
, and
S. I.
Seok
, “
Compositional engineering of perovskite materials for high-performance solar cells
,”
Nature
517
,
476
480
(
2015
).
4.
D. P.
McMeekin
,
G.
Sadoughi
,
W.
Rehman
,
G. E.
Eperon
,
M.
Saliba
,
M. T.
Hörantner
,
A.
Haghighirad
,
N.
Sakai
,
L.
Korte
,
B.
Rech
,
M. B.
Johnston
,
L. M.
Herz
, and
H. J.
Snaith
, “
A mixed-cation lead mixed-halide perovskite absorber for tandem solar cells
,”
Science
351
,
151
155
(
2016
).
5.
W. S.
Yang
,
B.-W.
Park
,
E. H.
Jung
,
N. J.
Jeon
,
Y. C.
Kim
,
D. U.
Lee
,
S. S.
Shin
,
J.
Seo
,
E. K.
Kim
,
J. H.
Noh
, and
S. I.
Seok
, “
Iodide management in formamidinium-lead-halide–based perovskite layers for efficient solar cells
,”
Science
356
,
1376
1379
(
2017
).
6.
A. K.
Jena
,
A.
Kulkarni
, and
T.
Miyasaka
, “
Halide perovskite photovoltaics: Background, status, and future prospects
,”
Chem. Rev.
119
,
3036
3103
(
2019
).
7.
NREL
, Research cell record efficiency chart, 2019, online accessed 23 August 2019.
8.
G.
Xing
,
N.
Mathews
,
S.
Sun
,
S. S.
Lim
,
Y. M.
Lam
,
M.
Grätzel
,
S.
Mhaisalkar
, and
T. C.
Sum
, “
Long-range balanced electron- and hole-transport lengths in organic–inorganic CH3NH3PbI3
,”
Science
342
,
344
347
(
2013
).
9.
W.-J.
Yin
,
T.
Shi
, and
Y.
Yan
, “
Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber
,”
Appl. Phys. Lett.
104
,
063903
(
2014
).
10.
H.
Zhou
,
Q.
Chen
,
G.
Li
,
S.
Luo
,
T.-b.
Song
,
H.-S.
Duan
,
Z.
Hong
,
J.
You
,
Y.
Liu
, and
Y.
Yang
, “
Interface engineering of highly efficient perovskite solar cells
,”
Science
345
,
542
546
(
2014
).
11.
C.
Bernal
and
K.
Yang
, “
First-principles hybrid functional study of the organic–inorganic perovskites CH3NH3SnBr3 and CH3NH3SnI3
,”
J. Phys. Chem. C
118
,
24383
24388
(
2014
).
12.
Q.
Dong
,
Y.
Fang
,
Y.
Shao
,
P.
Mulligan
,
J.
Qiu
,
L.
Cao
, and
J.
Huang
, “
Electron-hole diffusion lengths > 175 μm in solution-grown CH3NH3PbI3 single crystals
,”
Science
347
,
967
970
(
2015
).
13.
Y.
Li
,
M.
Behtash
,
J.
Wong
, and
K.
Yang
, “
Enhancing ferroelectric dipole ordering in organic-inorganic hybrid perovskite CH3NH3PbI3: Strain and doping engineering
,”
J. Phys. Chem. C
122
,
177
184
(
2018
).
14.
L.
Wu
,
P.
Lu
,
Y.
Li
,
Y.
Sun
,
J.
Wong
, and
K.
Yang
, “
First-principles characterization of two-dimensional (CH3(CH2)3NH3)2(CH3NH3)n−1GenI3n+1 perovskite
,”
J. Mater. Chem. A
6
,
24389
24396
(
2018
).
15.
M. R.
Filip
and
F.
Giustino
, “
Computational screening of homovalent lead substitution in organic–inorganic halide perovskites
,”
J. Phys. Chem. C
120
,
166
173
(
2016
).
16.
A.
Jain
,
O.
Voznyy
, and
E. H.
Sargent
, “
High-throughput screening of lead-free perovskite-like materials for optoelectronic applications
,”
J. Phys. Chem. C
121
,
7183
7187
(
2017
).
17.
X.-G.
Zhao
,
J.-H.
Yang
,
Y.
Fu
,
D.
Yang
,
Q.
Xu
,
L.
Yu
,
S.-H.
Wei
, and
L.
Zhang
, “
Design of lead-free inorganic halide perovskites for solar cells via cation-transmutation
,”
J. Am. Chem. Soc.
139
,
2630
2638
(
2017
).
18.
D.
Yang
,
J.
Lv
,
X.
Zhao
,
Q.
Xu
,
Y.
Fu
,
Y.
Zhan
,
A.
Zunger
, and
L.
Zhang
, “
Functionality-directed screening of Pb-free hybrid organic–inorganic perovskites with desired intrinsic photovoltaic functionalities
,”
Chem. Mater.
29
,
524
538
(
2017
).
19.
X.-G.
Zhao
,
D.
Yang
,
Y.
Sun
,
T.
Li
,
L.
Zhang
,
L.
Yu
, and
A.
Zunger
, “
Cu–In halide perovskite solar absorbers
,”
J. Am. Chem. Soc.
139
,
6718
6725
(
2017
).
20.
T.
Nakajima
and
K.
Sawada
, “
Discovery of Pb-free perovskite solar cells via high-throughput simulation on the K computer
,”
J. Phys. Chem. Lett.
8
,
4826
4831
(
2017
).
21.
S.
Chakraborty
,
W.
Xie
,
N.
Mathews
,
M.
Sherburne
,
R.
Ahuja
,
M.
Asta
, and
S. G.
Mhaisalkar
, “
Rational design: A high-throughput computational screening and experimental validation methodology for lead-free and emergent hybrid perovskites
,”
ACS Energy Lett.
2
,
837
845
(
2017
).
22.
S.
Lu
,
Q.
Zhou
,
Y.
Ouyang
,
Y.
Guo
,
Q.
Li
, and
J.
Wang
, “
Accelerated discovery of stable lead-free hybrid organic-inorganic perovskites via machine learning
,”
Nat. Commun.
9
,
3405
(
2018
).
23.
R.
Jacobs
,
G.
Luo
, and
D.
Morgan
, “
Materials discovery of stable and nontoxic halide perovskite materials for high-efficiency solar cells
,”
Adv. Funct. Mater.
29
,
1804354
(
2019
).
24.
Y.
Cai
,
W.
Xie
,
Y. T.
Teng
,
P. C.
Harikesh
,
B.
Ghosh
,
P.
Huck
,
K. A.
Persson
,
N.
Mathews
,
S. G.
Mhaisalkar
,
M.
Sherburne
, and
M.
Asta
, “
High-throughput computational study of halide double perovskite inorganic compounds
,”
Chem. Mater.
31
,
5392
5401
(
2019
).
25.
Y.
Li
and
K.
Yang
, “
High-throughput computational design of organic-inorganic hybrid halide semiconductors beyond perovskites for optoelectronics
,”
Energy Environ. Sci.
12
,
2233
2243
(
2019
).
26.
S.
Curtarolo
,
W.
Setyawan
,
G. L. W.
Hart
,
M.
Jahnatek
,
R. V.
Chepulskii
,
R. H.
Taylor
,
S.
Wang
,
J.
Xue
,
K.
Yang
,
O.
Levy
,
M.
Mehl
,
H. T.
Stokes
,
D. O.
Demchenko
, and
D.
Morgan
, “
AFLOW: An automatic framework for high-throughput materials discovery
,”
Comput. Mater. Sci.
58
,
218
226
(
2012
).
27.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
,
11169
11186
(
1996
).
28.
P. E.
Blöchl
, “
Projector augmented-wave method
,”
Phys. Rev. B
50
,
17953
17979
(
1994
).
29.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
30.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
, “
A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu
,”
J. Chem. Phys.
132
,
154104
(
2010
).
31.
A. A.
Emery
,
J. E.
Saal
,
S.
Kirklin
,
V. I.
Hegde
, and
C.
Wolverton
, “
High-throughput computational screening of perovskites for thermochemical water splitting applications
,”
Chem. Mater.
28
,
5621
5634
(
2016
).
32.
M. T.
Curnan
and
J. R.
Kitchin
, “
Effects of concentration, crystal structure, magnetism, and electronic structure method on first-principles oxygen vacancy formation energy trends in perovskites
,”
J. Phys. Chem. C
118
,
28776
28790
(
2014
).
33.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
, “
Hybrid functionals based on a screened Coulomb potential
,”
Chem. Phys.
118
,
8207
8215
(
2003
).
34.
A. V.
Krukau
,
O. A.
Vydrov
,
A. F.
Izmaylov
, and
G. E.
Scuseria
, “
Influence of the exchange screening parameter on the performance of screened hybrid functionals
,”
J. Chem. Phys.
125
,
224106
(
2006
).
35.
D.
Broberg
,
B.
Medasani
,
N. E.
Zimmermann
,
G.
Yu
,
A.
Canning
,
M.
Haranczyk
,
M.
Asta
, and
G.
Hautier
, “
PyCDT: A Python toolkit for modeling point defects in semiconductors and insulators
,”
Comput. Phys. Commun.
226
,
165
179
(
2018
).
36.
S. P.
Ong
,
W. D.
Richards
,
A.
Jain
,
G.
Hautier
,
M.
Kocher
,
S.
Cholia
,
D.
Gunter
,
V. L.
Chevrier
,
K. A.
Persson
, and
G.
Ceder
, “
Python materials genomics (pymatgen): A robust, open-source Python library for materials analysis
,”
Comput. Mater. Sci.
68
,
314
319
(
2013
).
37.
C.
Freysoldt
,
J.
Neugebauer
, and
C. G.
Van de Walle
, “
Fully ab initio finite-size corrections for charged-defect supercell calculations
,”
Phys. Rev. Lett.
102
,
016402
(
2009
).
38.
S.
Curtarolo
,
W.
Setyawan
,
S.
Wang
,
J.
Xue
,
K.
Yang
,
R. H.
Taylor
,
L. J.
Nelson
,
G. L. W.
Hart
,
S.
Sanvito
,
M. B.
Nardelli
,
N.
Mingo
, and
O.
Levy
, “
AFLOWLIB.ORG: A distributed materials properties repository from high-throughput ab initio calculations
,”
Comput. Mater. Sci.
58
,
227
235
(
2012
).
39.
A.
Jain
,
S. P.
Ong
,
G.
Hautier
,
W.
Chen
,
W. D.
Richards
,
S.
Dacek
,
S.
Cholia
,
D.
Gunter
,
D.
Skinner
,
G.
Ceder
, and
K. A.
Persson
, “
Commentary: The materials Project: A materials genome approach to accelerating materials innovation
,”
APL Matter
1
,
011002
(
2013
).
40.
C.
Freysoldt
,
B.
Grabowski
,
T.
Hickel
,
J.
Neugebauer
,
G.
Kresse
,
A.
Janotti
, and
C. G.
Van de Walle
, “
First-principles calculations for point defects in solids
,”
Rev. Mod. Phys.
86
,
253
(
2014
).
41.
S.
Saha
,
T. P.
Sinha
, and
A.
Mookerjee
, “
Electronic structure, chemical bonding, and optical properties of paraelectric BaTiO3
,”
Phys. Rev. B
62
,
8828
8834
(
2000
).
42.
G.
Giorgi
,
K.
Yamashita
, and
M.
Palummo
, “
Nature of the electronic and optical excitations of Ruddlesden–Popper hybrid organic–inorganic perovskites: The role of the many-body interactions
,”
J. Phys. Chem. Lett.
9
,
5891
5896
(
2018
).
43.
Y.
Cho
and
T. C.
Berkelbach
, “
Optical properties of layered hybrid organic–inorganic halide perovskites: A tight-binding GW-BSE study
,”
J. Phys. Chem. Lett.
10
,
6189
6196
(
2019
).

Supplementary Material