Introduction

Hybrid organic–inorganic perovskites1,2,3,4 recently surged to worldwide attention following the report of highly efficient solar cells based on methylammonium lead halides5,6,7,8. In fact, within less than 2 years of development, the photovoltaic energy-conversion efficiency of these devices increased from the initial 10.9% (ref. 5) to the current record of 17.9% (NREL certified) and 19.3% (uncertified)9. This impressive trend fuelled the expectation that perovskite photovoltaics may soon revolutionize the solar energy landscape.

Given the unusually long lifetimes of the photoexcited carriers10,11,

Figure 2: Map** the band gap landscape of metal-halide perovskites.
figure 2

(a) Two-dimensional map of the DFT band gap of the Platonic model of PbI3-based perovskites as a function of the apical and equatorial metal–halide–metal bond angles. The Pb–I bond length is set to the experimental value for MA-PbI3 (ref. 17). The angles αa and αe are indicated in Fig. 1d. The calculations for the Platonic model were performed within scalar-relativistic DFT (fully relativistic calculations are shown in Fig. 3b, see Methods). The dashed line at 45° is a guide to the eye and shows the approximate symmetry of the map with respect the the exchange of the apical and equatorial angles. The dark discs represent the angular coordinates (αa, αe) of realistic models of PbI3-based perovskites including the cation (indicated by the chemical formula), with structures fully optimized within DFT. The apical and equatorial angles are the averages among all the inequivalent angles of the same type. The size of each circle represents the deviation of the Pb–I bond length from that of MA–PbI3 (3.18 Å). This quantity is displayed because the metal–halide bond length has a small but non-negligible effect on the band gap, which we quantify as a linear shift of ~3 eV/Å (refs 27, 31). For the same angular coordinates, the band gap difference between the largest and the smallest circle in the figure is 0.2 eV. The white discs correspond to perovskites already synthesized13,17,18. This map shows that the band gap can be modulated over a much wider range than currently possible. (b,c) Atomistic models of the hypothetical metal-halide perovskites identified in this work exhibiting the smallest (LiPbI3) and the largest (NCl4PbI3) metal–halide–metal bond angles, respectively.

The variation of the angular coordinates in Fig. 2 clearly induces a substantial modulation of the band gap, from the top of the mid-infrared (1.1 eV) to the beginning of the visible spectrum (1.9 eV). This trend is in line with earlier calculations on simpler two-dimensional (2D) Sn-I perovskites for solution-processable electronics31.

The qualitative trend shown in Fig. 2 can be interpreted using elementary tight-binding arguments. If we consider for definiteness the bottom of the conduction band at Γ, which is most affected by bond angles, in the case of αa=αe=180° there are three degenerate electronic states derived from the 6p orbitals of Pb (without considering spin–orbit coupling for simplicity; this coupling is fully included in all our final results presented in Fig. 3b). It is intuitive that the energy of these states results mainly from ppσ bonding integrals since all the Pb-I bonds are collinear32. By moving away from the bottom left corner, the degeneracy is lifted as the bond integrals acquire ppπ components weighted by the angles αa and αe. Both components tend to raise the energy levels; therefore, the conduction band bottom results from the competition between the antibonding orbitals with the lowest energy along the apical or the equatorial direction. This is consistent with the approximate symmetry of the map in Fig. 2 with respect to the line corresponding to αa=αe. More importantly, this reasoning suggests that the band gap of the Platonic perovskite model is governed by the largest Pb–I–Pb bond angle in the metal–halide network, consistent with the trend in Fig. 2. This relation between band gap and bond angles holds unchanged when considering fully relativistic calculations including spin–orbit corrections. In fact, as shown in Supplementary Fig. 1, relativistic effects induce a large but slowly varying red shift of 0.8–1.1 eV across the family of compounds considered in this work. For the sake of completeness we also checked that excitonic effects play only a minor role, with exciton-binding energies never exceeding 80 meV as shown by Supplementary Fig. 2. Details on the calculation of exciton binding energies are given in the Methods, and the results are discussed in the Supplementary Note 1.

Figure 3: Tuning the band gap of metal-halide perovskites via the steric size of the cation.
figure 3

(a) Correlation between the apical and equatorial bond angles of PbI3-based perovskites and the steric radius of the cation (discs). As the size of the cation increases the metal–halide–metal bonds (see Fig. 1d) tend to become collinear. A linear least-square fit to the data (straight line) yields a slope of 18°/Å. (b) Correlation between the DFT band gap and the largest metal–halide–metal bond angle in the structure. We show calculations for the Platonic model (circles) and fully optimized structures within scalar-relativistic (grey discs) and fully relativistic (blue discs) DFT. The arrows point to the scales corresponding to each set of data. The significant data dispersion at large angles is due to the distortions of the octahedra. The calculations based on the Platonic model include the effect of the average Pb–I bond length on the band gap, as discussed in Fig. 2. (c) Calculated band gaps (in eV) of all the PbI3-based perovskites considered in this work. The band gaps were obtained after full structural optimization within scalar relativistic DFT (see Methods).

After taking into account the presence of the cation and the variations in the Pb-I bond lengths and Pb–I–Pb angles, the calculated band gaps deviate from the idealized predictions in Fig. 2 by 0.1–0.3 eV (see Fig. 3 and discussion below). Nonetheless, the general trend, which spans a range of up to 1 eV, is robust. This picture is also confirmed by relativistic GW calculations on a few selected compounds, as shown in Supplementary Fig. 4. The trend identified here suggests that in order to make low-gap perovskites for optimum photovoltaic efficiency we need to engineer structures with minimal octahedral tilt. On the other hand, in order to make large-gap perovskites for light-emitting diodes operating in the visible, we need to design structures with maximum tilt. The remaining question is how the tilt angle can be controlled.

Controlling metal–halide–metal bond angles via the steric size of the cation

The Pb–I–Pb bond angles define the volume of the cuboctahedral cavity; therefore, it is intuitive that the use of larger cations may lead to values of αa and αe closer to straight angles33. In the search for such cations we perform DFT calculations for existing structures as well as for many hypothetical structures not considered hitherto, including full structural optimizations and spin–orbit interactions, in the presence of the cation (see Methods). In order to relate these structures and their properties to the predictions of the Platonic model, we consider the average values of the Pb–I–Pb bond angles and of the Pb–I bond length over a crystalline unit cell. These values are used as the coordinates needed for locating each structure on the map of Fig. 2.

Previously reported cations are methylammonium (ref. 17), formamidinium (refs 13, 18), the alkali metal Cs+18 (synthesised) and ammonium (refs 34, 35, 36) (proposed). As shown in Fig. 2 all these cations cluster around the centre of the map; therefore, they are not expected to yield a significant modulation of the band gap. In line with this result their measured optical absorption onsets are very similar, within 0.25 eV (refs 13, 18).

In order to explore a wider portion of the map we consider four different families of cations generated from those above. The first family consists of secondary, tertiary and quaternary ammonium cations, namely di-, tri- and tetra-methylammonium. These are large molecules obtained by replacing hydrogen atoms bonded to the N atom with methyl groups. The second family is generated from ammonium by descending the pnictogen column in the periodic table, that is by substituting N for P, As or Sb. Members of this family include phosphonium , arsonium , stibonium , and similarly the methylammonium analogues , and . The third family of cations that we consider is obtained by replacing hydrogen in ammonium by halogen atoms, and possibly nitrogen by another pnictogen: in this group we have , , , and . The fourth and final family simply consists of the alkali metals Li+, Na+, K+ and Rb+. In total, we consider 22 structures, including 18 hypothetical compounds not reported to date.

In order to quantify the steric size of each cation, we use the radius of the sphere that contains 95% of the DFT electron density. This choice ensures that the steric radii of the alkali metals are in agreement with their standard ionic radii37. For completeness the calculated steric sizes are reported in Supplementary Table 1. Figure 3 shows that the steric radii of the cations considered in this work span a wide range, from 0.7 Å (Li+) to .

Our structural optimizations indicate that for some of the large cations the three-dimensional (3D) perovskite network is significantly distorted. This is in line with previous studies showing that large molecular cations determine a reorganization of the 3D perovskite network into a 2D layered structure38,39. These large molecules include tertiary and quaternary methylammonium cations, as well as the tetrafluoride and the tetrachloride . As these structures depart substantially from the 3D metal–halide network considered in this work, a separate analysis would be required. Accordingly, we do not consider them further.

Figure 3a shows that the largest Pb–I–Pb angles correlate strongly with the steric size of the cation (Spearman correlation 88%), in line with our initial expectation. The newly proposed structures span a range of angles from 130° to 170°, thereby covering a much wider portion of the band gap map in Fig. 2 than presently possible. In addition, Fig. 3b shows that, in agreement with the predictions of the Platonic model, the band gap correlates strongly with the largest Pb–I–Pb angle in the unit cell (Spearman correlation 91%). For completeness in Fig. 3b we report the band gaps obtained from this model as well as those calculated using the fully optimized structures within fully relativistic DFT calculations. The scatter of the data, which is clearly visible in Fig. 3b, is mostly because of the variation of the Pb–I bond lengths in the metal–halide network. We calculated the correlation coefficient between band gap and bond lengths using the same data, and we obtained a very weak correlation (Spearman coefficient 9%). This test clearly indicates that the role of bond lengths represents a second-order effect, thereby providing further support to our Platonic model. Finally, Fig. 3c shows our main finding, namely that by suitably choosing the cation it is possible, at least in principle, to fine-tune the band gap almost continuously over a very wide range of photon energies.

Our analysis highlights several new promising structures, in particular perovskites with phosphonium , arsonium and stibonium , all with band gaps ranging between 1.2 and 1.4 eV. These values fall in the middle of the range for optimum photovoltaic efficiency, and suggest that our new structures hold potential as novel solar cells materials. Besides the low gap, these perovskites are expected to exhibit higher mobilities than MA–PbI3. In fact, as shown in Supplementary Fig. 2a,b, the carrier-effective masses decrease towards larger bond angles, consistent with the concomitant reduction of the band gap. In addition to lower gaps and effective masses, large cations have the advantage that the filling of the cuboctahedral cavity should counter the known tendency of perovskite solar cells to degrade by water incorporation14.

In order to evaluate the practical feasibility of the hypothetical perovskites identified in this work we searched for possible dynamical instabilities in one of the structures with the smallest band gap, AsH4PbI3 ( is less interesting because of safety concerns related to chlorides). This structure is especially interesting since the preparation could proceed through the corresponding iodide AsH4I (ref. 40), similarly to the case of methylammonium iodide CH3NH3I. As shown by Supplementary Fig. 3, the vibrational density of states calculated using density-functional perturbation theory41 does not exhibit any soft modes, thereby indicating that the 3D perovskite structure of AsH4PbI3 should be stable against distortions. More generally, in order to assess the stability of all the structures considered here, we move to a simple semi-empirical approach based on the Goldschmidt tolerance factor t (refs 33, 42) (see Methods for the definition of t). It is known empirically that 3D perovskites should form when the tolerance factor falls within a narrow range t=0.7–1.1 (refs 33, 42, 43, 44). Figure 4a shows that the tolerance factors derived from our calculated steric sizes falls within this range for most of the structures considered here. In particular, according to this criterion, all the cations up to Rb and K should be stable.

Figure 4: Material stability and optical measurements.
figure 4

(a) Calculated Goldschmidt tolerance factors for all the perovskite structures considered in this work. The open circles are obtained using the steric sizes calculated from DFT (Supplementary Table 1). For comparison we also show the tolerance factors obtained by using the ionic radii reported in ref. 37 (filled circles). The grey area bound by the solid grey lines correspond to the geometric criterion for the Goldschmidt tolerance factor42,43, while the dashed lines correspond to the empirical range proposed in ref. 44. (b) Measured absorbance spectra of the dark phase of RbxCs1−xPbI3 perovskite thin films, with x=0, x=0.1 and x=0.2. When increasing the Rb content a continuous blue shift of the absorption onset is observed, from 720 nm (x=0) to 690 nm (x=0.2). This blue shift corresponds to an increase in band gap, and is assigned to the more pronounced octahedral tilt in the PbI3 network, as illustrated in the inset.

In order to demonstrate the band gap tunability we attempted the synthesis of perovskites based on elemental cations, building on our experience with CsPbI3 (ref. 18) (see Methods). As deposited, we observed that CsPbI3 formed a pale phase with a wide bandgap. Upon annealing this structure turned into a dark phase with a band gap of 1.73 eV, which we identify as the 3D perovskite structure discussed in this work. The synthesis of films with cations smaller than Cs (Rb, K, Na and Li) yielded similar pale phases. However, upon annealing, some of the films degraded via sublimation before we could observe a dark phase as for CsPbI3, suggesting that 3D perovskites were not formed. This finding is in agreement with our analysis of the Goldschmidt tolerance factor shown in Fig. 4a. While the proposed 3D perovskites could possibly be synthesised using higher pressures or longer annealing runs at lower temperatures, here for the sake of simplicity we did not pursue this direction. Instead we considered an alternative approach and explored mixtures of cations, as suggested in ref. 45: if mixtures of Cs and smaller cations could gradually shift the bandgap towards the larger values, this would confirm our predicted tunabiliy via steric effects.

As a proof-of-concept we prepared perovskites with Rb/Cs cation mixtures, namely RbxCs1−xPbI3 with x=0, x=0.1 and x=0.2 (never hitherto reported). Optical characterization of these films, shown in Fig. 4b, indicates that the absorption edge blue shifts from 720 to 690 nm in going from Cs to Rb. This result is in excellent agreement with our calculations, which predict a corresponding blue shift of 40 nm.

Our findings are also in line with a recent study (which appeared while the present manuscript was under review) where the authors demonstrated band gap tunability via the mixing of mixtures46. Taken together, our present results and those of refs 13, 18, 47 confirm the tunability of the band gap via the steric size of the cation over almost half of our predicted range, thereby providing strong support to our theory.

Finally, our theory is based on very general considerations and its predictions are not linked to the underlying calculation methods. Therefore, it is expected that this model will carry similar predictive power across other families of metal–halide perovskites. More fundamentally, the ability to control the band gap via the steric size of the cation, as proposed in this work, may become a new paradigm in the solution deposition of solar cells, light-emitting diodes and photonic structures with layer-by-layer control of their optical properties.

Note added in proof. While this work was under peer review, a related work exploring the relation between the size of cations and the electronic and optical properties of metal-halide perovskites was submitted and published63. The key difference between the present work and that of ref. 63 is that we develop a predictive universal theory of band gap tuning via steric effects, and we perform experiments that confirm a posteriori our predictions.