Abstract

Low-cost, nontoxic, and earth-abundant photovoltaic materials are long-sought targets in the solar cell research community. Perovskite-inspired materials have emerged as promising candidates for this goal, with researchers employing materials design strategies including structural, dimensional, and compositional transformations to avoid the use of rare and toxic elemental constituents, while attempting to maintain high optoelectronic performance. These strategies have recently been invoked to propose Ti-based vacancy-ordered halide perovskites (A2TiX6; A = CH3NH3, Cs, Rb, or K; X = I, Br, or Cl) for photovoltaic operation, following the initial promise of Cs2SnX6 compounds. Theoretical investigations of these materials, however, consistently overestimate their band gaps, a fundamental property for photovoltaic applications. Here, we reveal strong excitonic effects as the origin of this discrepancy between theory and experiment, a consequence of both low structural dimensionality and band localization. These findings have vital implications for the optoelectronic application of these compounds while also highlighting the importance of frontier-orbital character for chemical substitution in materials design strategies.
Perovskite-inspired materials aim to replicate the exceptional optoelectronic performance of lead halide perovskites (LHPs), while avoiding issues of toxicity and operational stability.1 For decades, the standard materials design approach for identifying novel inorganic semiconductors has been chemical substitution, in which the undesirable elemental constituents (e.g., toxic Pb2+ in LHPs) are replaced by more favorable counterparts, while retaining the same structural motifs. For example, in the diamond-cubic crystal family, research moved from group IV elements Si and Ge to II–VI compounds like CdTe, to yield direct rather than indirect electronic band gaps, and then further splitting into the I–III–VI2 (e.g., CuInSe2) and I2–II–IV–VI4 families (e.g., Cu2ZnSnS4), to give earth-abundant compositions. While strategies such as dimensional modification2 and disorder engineering3,4 have recently gained in popularity, elemental substitution remains the prevailing design approach.
Strategies for replacing the divalent B-site cation in halide perovskites, while retaining the BX6 octahedral motif, have led to the exploration of A2BB′X6 double perovskites with a pair of monovalent and trivalent cations at the B and B′ sites,1,5,6 as well as the A3B2X9 “vacancy-ordered perovskites”, in which a trivalent B cation is combined with a 1/3 vacancy of the B site to satisfy electroneutrality.7−9 Issues of indirect and/or large band gaps in these materials have led to the emergence of A2BX6 vacancy-ordered double perovskites (VODPs), in which now the combination of a tetravalent cation and a 50% vacancy of the B site is employed, giving a checkerboard arrangement (Figure 1).10−12 Also known as defective or tetravalent perovskites, these compounds are actually some of the decomposition products of conventional ABX3 perovskites, for example, CsSnI3, which breaks down to form Cs2SnI6.13
Figure 1.

Crystal structure of Cs2BX6 vacancy-ordered perovskites, in the conventional cubic unit cell (space group Fm3̅m). Cs atoms are colored green, M cations blue, and halide anions (X) purple.
A2BX6 vacancy-ordered perovskites have shown promise for optoelectronic applications, with good stability under air, moisture, light, and thermal stresses, as well as being solution-processable and nontoxic.11,14,15 As with the single and double perovskites, the frontier orbitals of the B cation and X anion govern the conduction and valence band-edge properties, respectively. This combination allows tunability in the energy gap, absorption profile, and carrier effective masses, for example. The A-site species, on the contrary, is a large monovalent cation such as Cs+, which behaves as a spectator, dictating the spacing between BX6 octahedra but with no direct contribution to the band edges. A crucial difference from the ABX3 perovskite template is the lack of corner-sharing BX6 octahedra caused by vacancy introduction. Consequently, the crystal structure is comprised of isolated octahedra and thus an effective zero-dimensional (0D) framework, with this dramatic reduction in connectivity being a key factor in the properties of this material family.10,11,14 Research efforts in this area initially focused on the Sn-based compounds (A2SnX6)16 but have since expanded so that a range of tetravalent species have successfully been implemented in A2BX6 materials, including Te, Pd, Zr, and Pt.17−20 While some of these compounds have shown promise as potential white light and tunable emitters, the Sn- and Ti-based materials have shown the most promising results in the context of solar photovoltaic applications and thus received a majority of the research attention. Cs2SnI6 was originally used as a hole-transporting layer in dye-sensitized solar cells, for instance, achieving efficiencies of 8%,16 while a Cs2TiBr6 photovoltaic device demonstrated a modest efficiency of 3%.14 The poor performance of these materials has been attributed to relatively weak visible light absorption and indirect band gaps.21
As issues of defect intolerance and operational instability are becoming apparent for Cs2SnI6,22 there is growing interest in Ti-based compounds. The effects on structure, stability, and electronic properties in going from the group 14 d10s0 Sn4+ to group 4 d0s0 Ti4+ cations have been probed;10,11,23 however, the performance limits of these materials remain an open question. Notably, while theoretical methods are found to successfully reproduce the experimental electronic structure of the Te- and Sn-based compounds,11 a major discrepancy exists for the d0 Ti-based compounds,10,14,21,23−29 with severe overestimation of the experimental band gap by both hybrid density functional theory (DFT) and Green’s function (GW) methods. So extreme is the error that these theoretical methods actually yield qualitatively incorrect relative band gap energies for the Sn versus Ti compounds, as we show in this study.
Through in-depth computations including explicit electron–hole interactions via the Bethe–Salpeter equation (BSE), we resolve the Ti perovskite discrepancy and reveal strong excitonic effects as the origin. Electron–hole interactions result in significant renormalization of the lowest-energy electronic excitation, as well as qualitative reshaping of the optical absorption spectrum, finally reconciling computational predictions with experimental measurements. We elucidate the origins of this behavior and highlight the implications of strong exciton binding for applications of these materials in optoelectronic devices.
The crystal structure of the Cs2BX6 (B = Sn or Ti; X = Cl, Br, or I) family of vacancy-ordered perovskites is shown in Figure 1. The low structural dimensionality of this family is expected to produce behavior similar to that of the corresponding [BX6]2– molecular salts.16,24 One consequence of this “molecular” crystal structure is the possibility for intermolecular interactions, such as London dispersion, between the localized octahedra. Table 1 corroborates this hypothesis, showing contraction of the calculated lattice parameters upon inclusion of dispersion corrections in the model, demonstrating the presence of important van der Waals (vdW) bonding contributions. Indeed, geometry optimization with hybrid DFT excluding dispersion corrections consistently overestimates the experimental lattice parameters by ∼3%, whereas inclusion of vdW effects gives lattice constants with errors of <1% in all cases. Semilocal DFT including dispersion corrections (PBE+D3) was also found to accurately reproduce the experimental lattice constants (Table S1). The change in the lattice parameter (ΔaD3) is consistent within each halide subclass, irrespective of the B-site identity (Sn or Ti), reflecting the expected interoctahedral (BX6–BX6) rather than intra-octahedral (B–X) origin of these vdW interactions. Moreover, we demonstrate the importance of dispersion interactions between the BX6 molecular blocks on the electronic properties, showing the calculated energy band gap to shift by 0.04–0.31 eV in the optimized crystal structure. There is an increasing sensitivity of the band gap to the lattice parameter as we move down the halogen group (Cl → Br → I), as the through-space B–X and X–X interactions in the conduction and valence bands strengthen with larger X p orbitals, also explaining the reduced band gap shifts for B = Ti due to the more localized d orbitals. We further note a sensitivity of the electronic band gap on the DFT functional choice for geometry optimization, with a 0.4 eV lower (−40%) band gap obtained for Cs2SnI6 using semilocal DFT (PBEsol) for structure relaxation.30 Hybrid DFT including dispersion corrections was employed for all further DFT calculations in this study.
Table 1. Calculated Cubic Lattice Parameters and Electronic Band Gap Shifts (ΔEg, D3) for Cs2BX6 (B = Sn or Ti; X = Cl, Br, or I) Using Hybrid DFT Including Spin–Orbit Coupling (HSE06+SOC), with and without Explicit Inclusion of vdW Dispersion Interactions (D3 correction)a.
| Cs2SnCl6 | Cs2SnBr6 | Cs2SnI6 | Cs2TiCl6 | Cs2TiBr6 | Cs2TiI6 | |
|---|---|---|---|---|---|---|
| aHSE06 (Å) | 10.65 | 11.15 | 11.95 | 10.51 | 10.99 | 11.76 |
| ΔaHSE06 (%) | 2.8 | 3.5 | 2.7 | 2.6 | 2.9 | 2.3 |
| aHSE06+D3 (Å) | 10.32 | 10.78 | 11.54 | 10.18 | 10.62 | 11.32 |
| ΔaHSE06+D3 (%) | –0.4 | 0.1 | –0.9 | –0.6 | –0.6 | –1.5 |
| aExp (Å) | 10.36 | 10.77 | 11.64 | 10.24 | 10.68 | 11.5 |
| ΔaD3 (Å) | –0.33 | –0.37 | –0.41 | –0.33 | –0.37 | –0.44 |
| ΔEg, D3 (eV) | –0.14 | –0.23 | –0.31 | –0.04 | –0.08 | –0.15 |
Lattice parameter errors (Δa) given with respect to experimental values. Experimental values taken from refs (31) and (32) for Cs2SnCl6, refs (32) and (33) for Cs2SnBr6, refs (30), (32), and (34−36) for Cs2SnI6, refs (23) and (37) for Cs2TiCl6, refs (21), (23), and (37) for Cs2TiBr6, and ref (38) for Cs2TiI6, matching with our measured values (Section S1.6)
The electronic band structures, densities of states, and charge densities at the valence band maximum (VBM) and conduction band minimum (CBM) for Cs2TiI6 and Cs2SnI6 are shown in Figures 2 and 3. While Cs2SnX6 compounds exhibit direct electronic band gaps at Γ, Cs2TiX6 compounds have indirect gaps with the CBM at the X high-symmetry k-point and the VBM remaining at Γ, in agreement with experimental studies.15,21 The direct/indirect gap energy difference (Δ) is relatively small, however, with Δ values of 0.06, 0.07, and 0.04 eV for the I, Br, and Cl isomorphs, respectively, calculated using HSE06+SOC. As previously noted,10,11 the VBM and CBM electronic levels follow that predicted by BX62– crystal field splitting molecular-orbital diagrams, with a t2g*(π) Ti d–X p CBM for Cs2TiX6 (dxy, dxz, dyz; 3-fold degenerate at Γ) and an eg(σ) band just above (dz2 and dx2–y2), a single a1g*(σ) Sn s–X p CBM for Cs2SnX6, and nonbonding X pt2g(π) states at the VBM in all cases (Figures 2 and 3c). The centrosymmetric crystal structure and equal (gerade) parity with respect to inversion for the VBM and CBM states (Figures 2 and 3b,c) result in a dipole-forbidden transition at the direct band gap. Consequently, the symmetry-allowed direct band gap (Eg, allowed) corresponds to the vertical transition from the second-highest valence band at Γ [t1u (Γ15) symmetry; ψVBM–1] to the CBM.
Figure 2.
(a) Electronic band structure of Cs2TiI6 calculated with hybrid DFT including spin–orbit coupling (HSE06+SOC), alongside a vertical plot of the orbital-projected electronic density of states. Faded gray and green arrows indicate the lowest-energy symmetry-forbidden and allowed electronic transitions, respectively (ΔEt1g/t1u = 0.02 eV). Valence band in blue, conduction band in orange, and valence band maximum (VBM) set to 0 eV. Ti d conduction bands are labeled with their crystal field orbital symmetries. Charge densities at the (b) conduction band minimum (CBM) and (c) VBM. Unoccupied states in yellow and occupied states in blue.
Figure 3.
(a) Electronic band structure of Cs2SnI6 calculated with hybrid DFT including spin–orbit coupling (HSE06+SOC), alongside a vertical plot of the orbital-projected electronic density of states. Faded gray and green arrows indicate the lowest-energy symmetry-forbidden and allowed electronic transitions, respectively (ΔEt1g/t1u = 0.38 eV). Valence band in blue, conduction band in orange, and VBM set to 0 eV. Charge densities at the (b) conduction band minimum (CBM) and (c) valence band maximum (VBM), using the same isosurface levels that were used for Cs2TiI6. Unoccupied states in yellow and occupied states in blue.
The halide p valence band is similar for both compounds, though with a slightly wider bandwidth (∼0.5 eV) for the Ti analogues (Figures 2 and 3 and Figures S3–S8) due to a significantly reduced anion–anion distance (dI–I = 4.03 Å for Cs2SnI6 vs dI–I = 3.87 Å for Cs2TiI6) and wider interaction range between the cation valence orbitals (Ti s and d) with anion p states in the lower valence band, compared to those of Sn p. This is a consequence of reduced M–X bond lengths (2.73 Å vs 2.85 Å) and lattice parameters for B = Ti versus Sn (Table 1), aided by the reduced ionic radius of Ti4+ versus that of Sn4+, resulting in a much lower energy difference between the t1g ψVBM and t1u ψVBM–1 for B = Ti versus Sn, with ΔEt1g/t1u = 0.02 eV/0.38 eV, 0.07 eV/0.30 eV, and 0.07 eV/0.07 eV for B = Ti/Sn and X = I, Br, Cl (using HSE06+SOC). Another consequence is that, in contrast to the electron masses, the hole effective masses are actually larger for Cs2SnX6 than for Cs2TiX6 (Table 2). Unlike conventional perovskites and many other “perovskite-inspired” materials that retain the partially oxidized, filled valence subshell of the B cation (yielding antibonding character at the VBM1,39,40), the fully oxidized B4+ in A2BX6 means we have a less dispersive, nonbonding VBM,4 yielding heavier hole masses (particularly for X = Br or Cl) and aiding carrier localization. In contrast, the conduction band of the Sn analogues is relatively disperse with low electron effective masses (Table 2) due to strong mixing and delocalization of the Sn s and X p states, while extremely flat bands are found for B = Ti due to weak Ti d–X p mixing and localized, isolated Ti d states. The band structures of the bromide and chloride isomorphs are included in Figures S3–S8, showing similar results, though with larger band gaps and reduced dispersion as X changes from I to Br to Cl. Further analysis of the electronic structure is provided in Section S2.
Table 2. Calculated Direct (Eg, direct) and Direct-Allowed Band Gaps (Eg, allowed), Effective Masses (
),a High-Frequency
Dielectric Constants (ε∞), and Wannier–Mott
Model Exciton Binding Energies (Eex, Wannier) for Cs2BX6 (B = Sn or Ti; X = Cl, Br, or
I), Using Hybrid DFT Including Spin–Orbit Coupling (HSE06+SOC)
and Comparison to Experimentally Reported Band Gap Rangesb.
| Cs2SnCl6 | Cs2SnBr6 | Cs2SnI6 | Cs2TiCl6 | Cs2TiBr6 | Cs2TiI6 | |
|---|---|---|---|---|---|---|
| Eg, direct (eV) | 4.10 | 2.39 | 0.71 | 3.68 | 2.75 | 1.69 |
| Eg, allowed (eV) | 4.38 | 2.70 | 1.09 | 3.79 | 2.84 | 1.71 |
| Eg, exp (eV) | 4.4–4.9 | 2.7–3.3 | 1.25–1.3 | 2.8–3.4 | 1.8–2.3 | 1.0–1.2 |
(m0) |
0.55 | 0.38 | 0.26 | 3.5 | 2.7 | 1.8 |
(m0) |
2.2 | 1.3 | 0.78 | 2.2 | 0.90 | 0.55 |
| ε∞ | 2.86 | 3.37 | 4.54 | 3.26 | 3.84 | 5.08 |
| Eex, Wannier (eV) | 0.73 | 0.35 | 0.13 | 1.73 | 0.62 | 0.22 |
values are computed from the harmonic mean
over directions and light/heavy bands for the effective masses. Values
of >1 are given to one decimal place.
Experimental band gap values taken from refs (31), (32), (42), and (43) for Cs2SnCl6, refs (32), (35), (36), and (42−44) for Cs2SnBr6, refs (16), (17), (31), (32), (35), (42), and (45) for Cs2SnI6, refs (23) and (37) for Cs2TiCl6, refs (14), (15), (21), (23), (24), (37), (38), and (46) for Cs2TiBr6, and refs (24) and (37) for Cs2TiI6.
The electronic properties of the Cs2BX6 family
are listed in Table 2. To illustrate the expected trends in exciton binding based on band
structure and dielectric screening, the Wannier–Mott model
binding energies are also included, calculated using the average carrier
effective masses (
) and high-frequency dielectric constants
(
) from hybrid DFT (HSE06+SOC) according
to41
| 1 |
where
is the reduced mass of the electron–hole
pair, Ry is the Rydberg energy (13.6 eV), and m0 is the electron rest mass.
From Table 2, we witness the typical trend of a larger band gap with smaller and more electronegative halogen anions [Eg(Cl) > Eg(Br) > Eg(I)], as observed across the perovskite (-inspired) family.1,10 Typically, the smaller the B-site atom, the smaller the band gap in the A2BX6 family.10,47 This is the case experimentally here, with all Ti isomorphs having experimentally measured band gaps that are smaller than those of their Sn counterparts. The opposite trend is found in the computed band gaps, for which hybrid DFT incorrectly predicts larger gaps for the Ti compounds (except for X = Cl). While the direct-allowed gaps computed by hybrid DFT mostly coincide with the lower end of experimental ranges for Cs2SnX6, neglecting Wannier–Mott-predicted exciton binding, the entirely opposite trend is found for each Cs2TiX6 isomorph, with a consistent severe overestimation of the experimental band gap. Notably, screened hybrid DFT (HSE06) tends to slightly underestimate rather than overestimate semiconductor band gaps, with this underestimation typically worsening with larger band gaps.48,49 The error in predicted band gaps for Cs2TiX6 results in qualitatively incorrect relative band gap energies for Cs2SnX6 versus Cs2TiX6 (X = I or Br).
A dielectric-dependent hybrid functional approach was also tested, which can improve the description of dielectric screening from hybrid DFT with fixed exchange (e.g., HSE06) and give reduced band gap prediction errors,49−51 though this only slightly reduced the hybrid DFT gap for Cs2TiI6 by 0.04 eV (αSCF = 24%), still giving a significantly overestimated band gap with a relative error ΔEg of ∼70%. Even using the computationally intensive GW approximation, typically a gold standard for predicting band gaps,52,53 the calculated quasiparticle gaps in fact show far worse overestimation [for both Cs2TiX6, as previously noted by Cucco et al.,10 and Cs2SnX6 (Section S3)]. These observations suggest the presence of physical interactions in Cs2TiX6 that are not captured in these single-particle electronic models.
This major experiment–-theory discrepancy is witnessed in reported values across the literature10,14,21,23−29 but has not been addressed until now. In many cases, semilocal DFT (known to severely underestimate semiconductor band gaps)48 has been employed to yield fortuitous error cancellation and thus theoretical values closer to experiment. As we show in this work, however, semilocal DFT predicts qualitatively incorrect relative band gaps [even finding Cs2SnI6 to be metallic, for example (Table S3)], alongside incorrect absorption spectra, thus being unsuitable for modeling the electronic structure of A2BX6 compounds.
Using the Wannier–Mott effective mass model (Table 2),54 we find large exciton binding energies, particularly for the Ti compounds (due to flat bands and heavy carrier masses), suggesting strong electron–hole interactions in these materials. Moreover, the large Wannier–Mott binding energies, isolated octahedra, low dielectric screening, and highly localized d orbitals suggest that strongly bound excitons may be formed for Cs2TiX6.41 To explore this hypothesis, we extend our model using GW to calculate quasiparticle eigenvalues and include electron–hole interactions via the Bethe–Salpeter equation (BSE). The optical absorption spectrum calculated with this approach, alongside that obtained from hybrid DFT and the experimental data, is shown in Figure 4.
Figure 4.

Optical absorption spectra of Cs2TiI6, Cs2TiBr6, and Cs2TiCl6 (from left to right, respectively), calculated with both hybrid DFT (dotted violet) and the G0W0 + BSE method (solid blue), alongside the experimental data from ultraviolet–visible spectroscopy (dashed green). To directly compare the spectral shapes, calculated curves have been rigidly shifted to match the experimental absorption onset (unshifted results shown in Section S3).
Electron–hole interactions are found to dramatically red-shift (Tables S2 and S3 and Figure S10) and qualitatively alter the absorption spectra for Cs2TiX6, now yielding excellent agreement with the peaked experimental spectra. Analysis of the electronic states reveals the lowest-energy bright exciton peak to originate from the t1u(π + σ) → t2g(d) electronic transition as expected, i.e., from the second-highest valence band (ψVBM–1) at Γ (Figure 2) to the CBM. The lowest-energy dark excitonic state, indicated by the arrows in Figure 4 and located 0.3–0.4 eV below the first bright peak, corresponds to the symmetry-forbidden t1g(π) ψVBM → t2g(d) ψCBM transition mentioned previously. While improved agreement with the experimental spectra is found for all Cs2TiX6 isomorphs, smaller changes in the spectral shapes are noticed when X = Br and Cl. This results from the low band dispersion in these compounds (demonstrated by the large effective masses in Table 2), resulting in similar strong excitonic downshifting of the low-energy excitations; t1u(π + σ) → t2g(d) and t2u(π) → t2g(d) corresponding to ⟨ψVBM–1|H′|ψCBM⟩ and ⟨ψVBM–2|H′|ψCBM⟩ transitions. Indeed, in their recent paper, Grandhi et al.37 refer to the absorption onset of Cs2TiBr6 as an exciton peak, with our calculations revealing in fact both low-energy peaks to be excitonic in nature. This strong renormalization of the low-energy excitations and the lack of a band-like absorption onset rule out standard spectrum fitting techniques (such as the Tauc and Elliott models) for extracting band gap and exciton binding energies.55,56 Notably, the experimental spectra for TiBr62– and TiCl62– salts reported by Brisdon et al.57 closely resemble the results for Cs2TiBr6 and Cs2TiCl6 reported here and in the literature,37,38 evidencing the conclusion of molecular crystal behavior, the orbital assignments of the absorption peaks, and the presence of strong electron–hole interactions. Improved agreement between the calculated and experimental spectra is also found for the Sn compounds upon inclusion of electron–hole interactions, as weaker exciton interactions modify transition intensities and shift spectral weights to give more peak-like absorption onsets. The small residual mismatch in some cases between the GW+BSE and experimental spectra could be a result of temperature effects (vibrations can lower the symmetry restriction of dark excitonic transitions in this range), quasiparticle lifetime broadening, or the neglect of higher-order terms in GW.56,58,59
We highlight that the low-energy absorption peaks for Cs2TiX6 correspond to charge-transfer Frenkel excitons, with the electron wave function localized on the Ti t2gd orbitals (dxy, dxz, dyz) and the hole localized on the surrounding X p orbitals of the BX6 octahedron. This form of exciton is commonly witnessed in organic and molecular crystals60 and has been well-established in other 3d0 (Ti4+, Sc3+, and Ca2+) halides,61 arising here as a consequence of the 0D polyhedral connectivity. The strongly bound nature of these excitons is further demonstrated by the band contributions (“fatband plot”) shown in Figure 6, where the delocalization of the exciton wave functions in reciprocal space for the Ti compounds corresponds to real-space localization of the exciton wavepacket.41 The large Stokes shifts (∼0.5 eV) and broad PL emission observed for this family of materials in both this study and other studies15,21,23 are other characteristic results of localized bound excitons, as well as strong exciton–phonon coupling and low-energy dark excitons, with the photogenerated electron–hole pair readily localizing within the lattice to yield emissive self-trapped exciton (STE) states.
Figure 6.
Band contributions to the brightest exciton state at the absorption onset in Cs2SnX6 (top) and Cs2TiX6 (bottom), calculated using the BSE approach. Band eigenvalues are indicated by the black dots, with filled circles weighted by their contributions to the exciton state and gray interpolating bands. The average of the three degenerate brightest states at the absorption onset is used, with the sum area of the filled circles normalized across all compositions. Hole and electron states are colored blue and orange, respectively, and the VBM is set to 0 eV.
In contrast, the reduced effective masses of Cs2SnX6 yield a weakly bound exciton as expected, demonstrated by dominant band contributions at the Γ point to the first bright exciton state. Figure 6 also illustrates the differing trends in band structure when B = Sn and Ti, as X changes from I to Br to Cl, with a greater band gap increase and a reduction in conduction band dispersion for Cs2SnX6 than Cs2TiX6 [such that the relative band gap energies of B = Sn and Ti change from X = I (Eg, Sn < Eg, Ti) to X = Cl (Eg, Ti < Eg, Sn)], due to the greater localization of the Ti d states. As discussed in Section S3, quasiparticle band gaps and thus exciton binding energies from GW(+BSE) remain overestimated for these vacancy-ordered compounds, consistent with recent studies that attribute this behavior to underscreening errors within the random phase approximation (RPA) employed within GW.10,62−64 As such, to avoid this issue and obtain a reasonable estimate of the exciton binding energies in these systems, we also employed a constrained-supercell approach in which an exciton state is generated by controlling spin initialization and band occupation. Here we calculate the exciton binding energy using hybrid DFT for multiple supercell sizes of ≤972 atoms and then extrapolate to the dilute limit using the relevant scaling relationship to avoid supercell-size effects.65,66 With this approach, we obtain localized Frenkel exciton states as expected for each Cs2TiX6 (Figure S13, TOC), with extrapolated binding energies of 0.44, 0.52, and 0.72 eV for X = I, Br, and Cl, respectively (Figure S14), which when subtracted from the HSE06+SOC direct-allowed transition energies in Table 2 brings the hybrid DFT optical transition energy into agreement with the experimental values in each case. For Cs2SnX6, the electron and hole remain delocalized across the supercell with this approach [under a maximum cell length of 23.1 Å (Figure S13)], yielding extrapolated binding energies close to zero (Figure S15). Further details are provided in Figure S3.3.
Crucially,
these results demonstrate the presence of qualitatively
different electronic behavior in the Cs2SnX6 and Cs2TiX6 families, where despite retaining
the same cation valence, the change in frontier-orbital character
upon substitution of Sn4+ with Ti4+ dramatically
alters the electronic structure and optical absorption. From the band
structures in Figure 6 and values in Table 2, it is clear that the electron effective masses (
) dictate the exciton behavior in this family,
with the weak dispersion and strong real-space localization of the
flat d-orbital conduction bands in Cs2TiX6,
aided by the 0D crystal structure, yielding strong electron–hole
interactions. This strong excitonic renormalization of the optical
absorption in Cs2TiX6 explains the origin of
long-standing discrepancies between experiment and theoretical models
of their electronic structure. Moreover, these findings serve as a
warning of the changes that can occur when employing ionic substitution
as a materials design approach, when such strategies involve changes
in valence orbital character.
The presence of strong excitonic interactions in this material family is unsurprising, given the low structural and electronic dimensionality (Figures 1 and 2), weak band dispersion, and large carrier effective masses (Table 2) discussed above. We find the exciton binding strength to be governed by the conduction band character in these compounds, giving the expectation for similar strongly bound Frenkel excitons in A2BX6 compounds with isoelectronic (d0) B4+ cations, such as Zr and Hf. Indeed, strong excitonic interactions have been recently been reported in Cs2ZrX6, promising white light emitters,20 and a distinct excitonic feature is seen at the absorption onset in Cs2HfCl6,67,68 which has emission and radiation detection applications. Moreover, the bound excitonic behavior in this material class is very similar to that witnessed in the double perovskites,55,69,70 which despite a greater structural connectivity, exhibit a low effective electronic dimensionality due to orbital mismatch between the B-site cations.5,71 Likewise, extension of theoretical models to include explicit electron–hole interactions was required to reproduce the experimental spectrum,55,70,72 explaining the excitonic origin of the direct absorption onset.
In conclusion, by revealing strongly bound excitonic behavior in the cesium titanium halide vacancy-ordered perovskites (Cs2TiX6), we reconcile long-standing discrepancies between theoretical predictions and experimental measurements for this material class. While previous theoretical studies have found semilocal DFT to yield band gaps matching those from experiment, we show this to be the result of fortuitous error cancellation with qualitatively incorrect absorption spectra and relative band gaps for Cs2SnX6 versus Cs2TiX6 (X = I or Br). Our results show that electron–hole interactions are crucial to obtaining the correct polarizability and dielectric screening between octahedra within many-body perturbation theory (MBPT) in these low-electronic-dimensionality systems. A range of optical, photoelectron, and polarization spectroscopies could be employed to further study the behavior of excitons in this material class, including Stark spectroscopy, temperature-dependent optical measurements, and excitation-dependent terahertz and electromodulation spectroscopies.58,61,69,73 Moreover, a majority of previous theoretical studies have not included vdW dispersion interactions when modeling these systems, yet here we demonstrate their importance in obtaining accurate crystal and electronic structure predictions, calling for their inclusion in future computational studies of these and related low-dimensional and “molecular” crystals such as the A4BX6 family.
These findings have important implications for optoelectronic applications. Strong exciton binding can significantly reduce charge separation and open-circuit voltages (Voc) in solar cells, likely one of the key origins of the poor photovoltaic performance achieved thus far in this material class. Our results show the key role of structural dimensionality and octahedral connectivity, alongside orbital chemistry, in determining the effective electronic dimensionality and optoelectronic properties of inorganic perovskite-inspired materials. More generally, these findings illustrate the importance of considering frontier-orbital character when employing atomic substitution in materials engineering and design strategies, here resulting in qualitatively different electronic behavior despite equal cation valence and similar band gaps.
Computational Methods
Calculations were performed using both DFT and quasiparticle Green’s function (GW) approaches within periodic boundary conditions, through the Vienna Ab Initio Simulation Package (VASP).74 Scalar-relativistic pseudopotentials were employed to describe the interaction between core and valence electrons, via the projector-augmented wave (PAW) method.75 Specifically, the Cs_sv, Sn_d, Ti_pv, I, Br, and Cl VASP PAW potentials were used. The effect of pseudopotential choice and DFT starting point (semilocal vs hybrid) on the GW results was tested and found to give qualitatively similar results, with the same large overestimation of band gaps relative to those from experiment (details in Section S3).
Initial guesses for the crystal structures were obtained from the Materials Project, before relaxing the geometry using the HSE06 screened hybrid DFT functional with dispersion corrections.76 The plane-wave energy cutoff and Γ-centered k-point mesh were sequentially increased using vaspup2.0(77) until the total energies from static calculations were converged to 0.1 meV/atom, giving values of 300 eV and 3 × 3 × 3, respectively (for the nine-atom primitive unit cell, equivalent to a k-point density of 0.33 Å–1 in reciprocal space). During structural optimization, a convergence criterion of 0.01 eV/Å was imposed on the forces on each atom and the plane-wave energy cutoff was increased to 500 eV, and the final geometries were re-relaxed, to avoid Pulay stress effects. As discussed at the beginning of the results section, the effect of dispersion corrections (Grimmes D3)78 on the structural relaxation was tested and shown to be important; therefore, the HSE06+D3 (using the recommended PBE0+BJ parametrizations)79 relaxed unit cells were used for all further calculations in this work.
Electronic band structures and independent-particle optical absorption spectra were initially calculated with the HSE06 hybrid DFT functional, including spin–orbit coupling effects (HSE06+SOC) due to the presence of heavy-atom elements (see results and Section S4). For density of states (DOS) and optical calculations, the k-point mesh for the primitive unit cell was increased to 8 × 8 × 8 (reciprocal-space density of 0.12 Å–1), and the tetrahedron smearing method was used. The number of virtual states in the optical calculations was increased using vaspup2.0(77) until the high-frequency dielectric constant ε∞ was converged to a precision of 0.01. Electronic band structure diagrams were generated using sumo.80 Carrier effective masses were determined using nonparabolic fitting of the band edges through the effmass(81) package.
Wave functions calculated using HSE06+SOC were used as input orbitals for the G0W0(+BSE) calculations. While only having a modest effect on the band gap energies, SOC was found to have a relatively significant effect on the spectral shape above the absorption onset, as shown in Section S4. Convergence with respect to the number of virtual states/empty bands, imaginary frequency and time grid points, and electron–hole excitation pairs was confirmed in each case. Given the lack of symmetry reduction, the requirement for large numbers of virtual states/empty bands (particularly when including spin–orbit coupling effects) and rapid scaling of the computational cost (memory) with k-point density in the G0W0+BSE calculations, a 3 × 3 × 3 mesh (for the nine-atom primitive unit cell, equivalent to a k-point density of 0.33 Å–1 in reciprocal space) was the maximum tractable k-point mesh for Cs2BX6 with our computational resources. While converged for the Br and Cl compounds, the greater band dispersion of the iodides (Cs2BI6) means the spectra are not well converged for this k-point density; thus, the “model BSE” approach82,83 was employed to reach converged k-point meshes of 4 × 4 × 4 and 8 × 8 × 8 for Cs2TiI6 and Cs2SnI6, respectively. Further details about the G0W0+BSE calculations are given in Section S3.
Details of the experimental synthesis and absorption measurements are provided in Section S1.
Figure 5.

Optical absorption spectra of Cs2SnI6, Cs2SnBr6, and Cs2SnCl6 (from left to right, respectively), calculated with both hybrid DFT (dotted violet) and the G0W0+BSE method (solid blue), alongside the experimental data from ultraviolet–visible spectroscopy (dashed green). To directly compare the spectral shapes, calculated curves have been rigidly shifted to match the experimental absorption onset (unshifted results shown in Section S3). For Cs2SnCl6, the absorption spectrum recorded by Karim et al.32 is also shown for comparison (downshifted by 0.5 eV as discussed in Section S3.4).
Acknowledgments
The authors thank Alex M. Ganose for useful discussions regarding the electronic structure of A2BX6 vacancy-ordered perovskites. S.R.K. acknowledges the EPSRC Centre for Doctoral Training in the Advanced Characterisation of Materials (CDT-ACM) (EP/S023259/1) for funding a Ph.D. studentship. C.N.S. is grateful to the Department of Chemistry at UCL and the Ramsay Memorial Fellowship Trust for the funding of a Ramsay fellowship. D.O.S. acknowledges support from the EPSRC (EP/N01572X/1) and from the European Research Council, ERC (Grant 758345). The authors acknowledge the use of the UCL Kathleen High Performance Computing Facility (Kathleen@UCL), the Imperial College Research Computing Service, and associated support services in the completion of this work. Via membership of the UK’s HEC Materials Chemistry Consortium, which is funded by the EPSRC (EP/L000202, EP/R029431, and EP/T022213), this work used the ARCHER2 UK National Supercomputing Service (www.archer2.ac.uk) and the UK Materials and Molecular Modelling (MMM) Hub (Thomas EP/P020194 and Young EP/T022213). The authors acknowledge financial support from the European Research Council (ERC) under the European Unions Horizon 2020 research and innovation program (Grant Agreement 725165) as well as from the European Unions Horizon 2020 research and innovation program under Marie Skodowska-Curie Grant Agreement 713729. This project has received funding also from the Spanish State Research Agency, through the Severo Ochoa Center of Excellence (CEX2019-000910-S), the CERCA Programme/Generalitat de Catalunya and Fundacio Mir-Puig. The authors also acknowledge funding by the Fundacio Joan Ribas Araquistain (FJRA). This project was funded also by EQC2019-005797-P (AEI/FEDER UE).
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.2c02436.
Experimental methods, DFT functional dependence of lattice parameters, additional electronic structure analysis, further GW calculation details and analysis, and spin–orbit coupling analysis (PDF)
Author Contributions
Conceptualization: S.R.K., G.K., A.W., and D.O.S. Investigation and methodology: S.R.K., S.M.L., and C.N.S. Writing of the original draft: S.R.K. Review and editing: S.R.K., C.N.S., S.M.L., G.K., A.W., and D.O.S. Supervision, resources, and funding acquisition: G.K., A.W., and D.O.S. These author contributions are defined according to the CRediT contributor roles taxonomy.
The authors declare no competing financial interest.
Notes
Data produced during this work is freely available at doi.org/10.5281/zenodo.7275359.
Supplementary Material
References
- Huang Y.-T.; Kavanagh S. R.; Scanlon D. O.; Walsh A.; Hoye R. L. Z. Perovskite-Inspired Materials for Photovoltaics and beyond—from Design to Devices. Nanotechnology 2021, 32, 132004. 10.1088/1361-6528/abcf6d. [DOI] [PubMed] [Google Scholar]
- Aubrey M. L.; Saldivar Valdes A.; Filip M. R.; Connor B. A.; Lindquist K. P.; Neaton J. B.; Karunadasa H. I. Directed Assembly of Layered Perovskite Heterostructures as Single Crystals. Nature 2021, 597, 355–359. 10.1038/s41586-021-03810-x. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Kavanagh S. R.; Burgués-Ceballos I.; Walsh A.; Scanlon D.; Konstantatos G. Cation Disorder Engineering Yields AgBiS2 Nanocrystals with Enhanced Optical Absorption for Efficient Ultrathin Solar Cells. Nat. Photonics 2022, 16, 235–241. 10.1038/s41566-021-00950-4. [DOI] [Google Scholar]
- Huang Y.-T.; et al. Strong Absorption and Ultrafast Localisation in NaBiS2 Nanocrystals with Slow Charge-Carrier Recombination. Nat. Commun. 2022, 13, 4960. 10.1038/s41467-022-32669-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Savory C. N.; Walsh A.; Scanlon D. O. Can. Pb-Free Halide Double Perovskites Support High-Efficiency Solar Cells?. ACS Energy Lett. 2016, 1, 949–955. 10.1021/acsenergylett.6b00471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nie R.; Sumukam R. R.; Reddy S. H.; Banavoth M.; Seok S. I. Lead-Free Perovskite Solar Cells Enabled by Hetero-Valent Substitutes. Energy Environ. Sci. 2020, 13, 2363–2385. 10.1039/D0EE01153C. [DOI] [Google Scholar]
- Krajewska C. J.; Kavanagh S. R.; Zhang L.; Kubicki D. J.; Dey K.; Galkowski K.; Grey C. P.; Stranks S. D.; Walsh A.; Scanlon D. O.; Palgrave R. G. Enhanced Visible Light Absorption in Layered Cs3Bi2Br9 through Mixed-Valence Sn(Il)/Sn(Iv) Doping. Chem. Sci. 2021, 12, 14686–14699. 10.1039/D1SC03775G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peng Y.; Li F.; Wang Y.; Li Y.; Hoye R. L. Z.; Feng L.; Xia K.; Pecunia V. Enhanced Photoconversion Efficiency in Cesium-Antimony-Halide Perovskite Derivatives by Tuning Crystallographic Dimensionality. Appl. Mater. Today 2020, 19, 100637. 10.1016/j.apmt.2020.100637. [DOI] [Google Scholar]
- Roy M.; Ghorui S.; Bhawna; Kangsabanik J.; Yadav R.; Alam A.; Aslam M. Enhanced Visible Light Absorption in Layered Cs3Bi2Br9 Halide Perovskites: Heterovalent Pb2+ Substitution-Induced Defect Band Formation. J. Phys. Chem. C 2020, 124, 19484–19491. 10.1021/acs.jpcc.0c05880. [DOI] [Google Scholar]
- Cucco B.; Bouder G.; Pedesseau L.; Katan C.; Even J.; Kepenekian M.; Volonakis G. Electronic Structure and Stability of Cs2TiX6 and Cs2ZrX6 (X = Br, I) Vacancy Ordered Double Perovskites. Appl. Phys. Lett. 2021, 119, 181903. 10.1063/5.0070104. [DOI] [Google Scholar]
- Maughan A. E.; Ganose A. M.; Scanlon D. O.; Neilson J. R. Perspectives and Design Principles of Vacancy-Ordered Double Perovskite Halide Semiconductors. Chem. Mater. 2019, 31, 1184–1195. 10.1021/acs.chemmater.8b05036. [DOI] [Google Scholar]
- Sa R.; Zhang Q.; Luo B.; Liu D. Exploring the Electronic and Optical Properties of Vacancy-Ordered Double Perovskites Cs2PtX6 (X = Cl, Br, I). J. Solid State Chem. 2021, 304, 122602. 10.1016/j.jssc.2021.122602. [DOI] [Google Scholar]
- Wu J.; Fang F.; Zhao Z.; Li T.; Ullah R.; Lv Z.; Zhou Y.; Sawtell D. Fluorine Ion Induced Phase Evolution of Tin-Based Perovskite Thin Films: Structure and Properties. RSC Adv. 2019, 9, 37119–37126. 10.1039/C9RA07415E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen M.; Ju M.-G.; Carl A. D.; Zong Y.; Grimm R. L.; Gu J.; Zeng X. C.; Zhou Y.; Padture N. P. Cesium Titanium(IV) Bromide Thin Films Based Stable Lead-free Perovskite Solar Cells. Joule 2018, 2, 558–570. 10.1016/j.joule.2018.01.009. [DOI] [Google Scholar]
- He Y.; Guo X.; Zheng H.; Xv L.; Li S. Stability Investigation of the Titanium-Based Eco-Friendly Perovskite-like Antifluorite Cs2TiBr6. J. Mater. Chem. C 2022, 10, 9301–9309. 10.1039/D2TC00729K. [DOI] [Google Scholar]
- Lee B.; Stoumpos C. C.; Zhou N.; Hao F.; Malliakas C.; Yeh C.-Y.; Marks T. J.; Kanatzidis M. G.; Chang R. P. H. Air-Stable Molecular Semiconducting Iodosalts for Solar Cell Applications: Cs2SnI6 as a Hole Conductor. J. Am. Chem. Soc. 2014, 136, 15379–15385. 10.1021/ja508464w. [DOI] [PubMed] [Google Scholar]
- Maughan A. E.; Ganose A. M.; Bordelon M. M.; Miller E. M.; Scanlon D. O.; Neil-son J. R. Defect Tolerance to Intolerance in the Vacancy-Ordered Double Perovskite Semiconductors Cs2SnI6 and Cs2TeI6. J. Am. Chem. Soc. 2016, 138, 8453–8464. 10.1021/jacs.6b03207. [DOI] [PubMed] [Google Scholar]
- Evans H. A.; Fabini D. H.; Andrews J. L.; Koerner M.; Preefer M. B.; Wu G.; Wudl F.; Cheetham A. K.; Seshadri R. Hydrogen Bonding Controls the Structural Evolution in Perovskite-Related Hybrid Platinum(IV) Iodides. Inorg. Chem. 2018, 57, 10375–10382. 10.1021/acs.inorgchem.8b01597. [DOI] [PubMed] [Google Scholar]
- Sakai N.; Haghighirad A. A.; Filip M. R.; Nayak P. K.; Nayak S.; Ramadan A.; Wang Z.; Giustino F.; Snaith H. J. Solution-Processed Cesium Hexabromopalla-date(IV), Cs2PdBr6, for Optoelectronic Applications. J. Am. Chem. Soc. 2017, 139, 6030–6033. 10.1021/jacs.6b13258. [DOI] [PubMed] [Google Scholar]
- Abfalterer A.; Shamsi J.; Kubicki D. J.; Savory C. N.; Xiao J.; Divitini G.; Li W.; Macpherson S.; Galkowski K.; MacManus-Driscoll J. L.; Scanlon D. O.; Stranks S. D. Colloidal Synthesis and Optical Properties of Perovskite-Inspired Cesium Zirconium Halide Nanocrystals. ACS Mater. Lett. 2020, 2, 1644–1652. 10.1021/acsmaterialslett.0c00393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Euvrard J.; Wang X.; Li T.; Yan Y.; Mitzi D. B. Is Cs2TiBr6 a Promising Pb-free Perovskite for Solar Energy Applications?. J. Mater. Chem. A 2020, 8, 4049–4054. 10.1039/C9TA13870F. [DOI] [Google Scholar]
- Saparov B.; Sun J.-P.; Meng W.; Xiao Z.; Duan H.-S.; Gunawan O.; Shin D.; Hill I. G.; Yan Y.; Mitzi D. B. Thin-Film Deposition and Characterization of a Sn-Deficient Perovskite Derivative Cs2SnI6. Chem. Mater. 2016, 28, 2315–2322. 10.1021/acs.chemmater.6b00433. [DOI] [Google Scholar]
- Kong D.; Cheng D.; Wang X.; Zhang K.; Wang H.; Liu K.; Li H.; Sheng X.; Yin L. Solution Processed Lead-Free Cesium Titanium Halide Perovskites and Their Structural, Thermal and Optical Characteristics. J. Mater. Chem. C 2020, 8, 1591–1597. 10.1039/C9TC05711K. [DOI] [Google Scholar]
- Ju M.-G.; Chen M.; Zhou Y.; Garces H. F.; Dai J.; Ma L.; Padture N. P.; Zeng X. C. Earth-Abundant Nontoxic Titanium(IV)-Based Vacancy-Ordered Double Perovskite Halides with Tunable 1.0 to 1.8 eV Bandgaps for Photovoltaic Applications. ACS Energy Lett. 2018, 3, 297–304. 10.1021/acsenergylett.7b01167. [DOI] [Google Scholar]
- Li W.; Zhu S.; Zhao Y.; Qiu Y. Structure, Electronic and Optical Properties of Cs2Ti(Br1–xYx)6 (Y = Cl, I; x = 0, 0.25, 0.5, 0.75, 1) Perovskites: The First Principles Investigations. J. Solid State Chem. 2020, 284, 121213. 10.1016/j.jssc.2020.121213. [DOI] [Google Scholar]
- Mahmood Q.; Hassan M.; Yousaf N.; AlObaid A. A.; Al-Muhimeed T. I.; Morsi M.; Albalawi H.; Alamri O. A. Study of Lead-Free Double Perovskites Halides Cs2TiCl6, and Cs2TiBr6 for Optoelectronics, and Thermoelectric Applications. Mater. Sci. Semicond. Process 2022, 137, 106180. 10.1016/j.mssp.2021.106180. [DOI] [Google Scholar]
- Zhao P.; Su J.; Guo Y.; Wang L.; Lin Z.; Hao Y.; Ouyang X.; Chang J. Cs2TiI6: A Potential Lead-Free All-Inorganic Perovskite Material for Ultrahigh-Performance Photovoltaic Cells and Alpha-Particle Detection. Nano Res. 2022, 15, 2697. 10.1007/s12274-021-3801-5. [DOI] [Google Scholar]
- Natik A.; Abid Y.; Moubah R.; Abid M.; Lassri H. Ab-Initio Investigation of the Structural, Electronic and Optical Properties of Lead-Free Halide Cs2TiI6 Double Perovskites. Solid State Commun. 2020, 319, 114006. 10.1016/j.ssc.2020.114006. [DOI] [Google Scholar]
- Liu D.; Zha W.; Yuan R.; Chen J.; Sa R. A First-Principles Study on the Optoelectronic Properties of Mixed-Halide Double Perovskites Cs2TiI6–xBrx. New J. Chem. 2020, 44, 13613. 10.1039/D0NJ02535F. [DOI] [Google Scholar]
- Maughan A. E.; Ganose A. M.; Candia A. M.; Granger J. T.; Scanlon D. O.; Neil-son J. R. Anharmonicity and Octahedral Tilting in Hybrid Vacancy-Ordered Double Perovskites. Chem. Mater. 2018, 30, 472–483. 10.1021/acs.chemmater.7b04516. [DOI] [Google Scholar]
- Zhu W.; Xin G.; Wang Y.; Min X.; Yao T.; Xu W.; Fang M.; Shi S.; Shi J.; Lian J. Tunable Optical Properties and Stability of Lead Free All Inorganic Perovskites (Cs2SnIxCl6–x). J. Mater. Chem. A 2018, 6, 2577–2584. 10.1039/C7TA10040J. [DOI] [Google Scholar]
- Karim M. M. S.; Ganose A. M.; Pieters L.; Winnie Leung W. W.; Wade J.; Zhang L.; Scanlon D. O.; Palgrave R. G. Anion Distribution, Structural Distortion, and Symmetry-Driven Optical Band Gap Bowing in Mixed Halide Cs2SnX6 Vacancy Ordered Double Perovskites. Chem. Mater. 2019, 31, 9430–9444. 10.1021/acs.chemmater.9b03267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ketelaar J. A. A.; Rietdijk A. A.; van Staveren C. H. Die Kristallstruktur von Ammonium-, Kalium-, Rubidium- und Cäsiumstannibromid. Recl. Trav. Chim. Pays-Bas 1937, 56, 907–908. 10.1002/recl.19370560913. [DOI] [Google Scholar]
- Stoumpos C. C.; Malliakas C. D.; Kanatzidis M. G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019–9038. 10.1021/ic401215x. [DOI] [PubMed] [Google Scholar]
- Lee B.; Krenselewski A.; Baik S. I.; Seidman D. N.; Chang R. P. H. Solution Processing of Air-Stable Molecular Semiconducting Iodosalts, Cs2SnI6–xBrx, for Potential Solar Cell Applications. Sustain. Energy Fuels 2017, 1, 710–724. 10.1039/C7SE00100B. [DOI] [Google Scholar]
- Kaltzoglou A.; Antoniadou M.; Kontos A. G.; Stoumpos C. C.; Perganti D.; Sir-anidi E.; Raptis V.; Trohidou K.; Psycharis V.; Kanatzidis M. G.; Falaras P. Optical-Vibrational Properties of the Cs 2 SnX 6 (X = Cl, Br, I) Defect Perovskites and Hole-Transport Efficiency in Dye-Sensitized Solar Cells. J. Phys. Chem. C 2016, 120, 11777–11785. 10.1021/acs.jpcc.6b02175. [DOI] [Google Scholar]
- Grandhi G.; Matuhina A.; Liu M.; Annurakshita S.; Ali-Löytty H.; Bautista G.; Vivo P. Lead-Free Cesium Titanium Bromide Double Perovskite Nanocrystals. Nanomaterials 2021, 11, 1458. 10.3390/nano11061458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liga S. M.; Konstantatos G. Colloidal Synthesis of Lead-Free Cs2TiBr6–xIx Perovskite Nanocrystals. J. Mater. Chem. C 2021, 9, 11098–11103. 10.1039/D1TC01732B. [DOI] [Google Scholar]
- Kavanagh S. R.; Savory C. N.; Scanlon D. O.; Walsh A. Hidden Spontaneous Polarisation in the Chalcohalide Photovoltaic Absorber Sn2SbS2I3. Mater. Horiz. 2021, 8, 2709–2716. 10.1039/D1MH00764E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang G.; Ghosez P.; Hong J. Band-Edge Orbital Engineering of Perovskite Semiconductors for Optoelectronic Applications. J. Phys. Chem. Lett. 2021, 12, 4227–4239. 10.1021/acs.jpclett.0c03816. [DOI] [PubMed] [Google Scholar]
- Fox M.Optical Properties of Solids, 2nd ed.; Oxford Master Series in Physics; Oxford University Press: Oxford, U.K., 2010. [Google Scholar]
- Dalpian G. M.; Liu Q.; Stoumpos C. C.; Douvalis A. P.; Balasubramanian M.; Kanatzidis M. G.; Zunger A. Changes in Charge Density vs Changes in Formal Oxidation States: The Case of Sn Halide Perovskites and Their Ordered Vacancy Analogues. Phys. Rev. Mater. 2017, 1, 025401. 10.1103/PhysRevMaterials.1.025401. [DOI] [Google Scholar]
- Zhou J.; Luo J.; Rong X.; Wei P.; Molokeev M. S.; Huang Y.; Zhao J.; Liu Q.; Zhang X.; Tang J.; Xia Z. Lead-Free Perovskite Derivative Cs2SnCl6–xBrx Single Crystals for Narrowband Photodetectors. Adv. Opt. Mater. 2019, 7, 1900139. 10.1002/adom.201900139. [DOI] [Google Scholar]
- Bonomi S.; Patrini M.; Bongiovanni G.; Malavasi L. Versatile Vapor Phase Deposition Approach to Cesium Tin Bromide CsSnBr3, CsSn2Br5 and Cs2SnBr6. RSC Adv. 2020, 10, 28478–28482. 10.1039/D0RA04680A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang J.; Yu C.; Wang L.; Li Y.; Ren Y.; Shum K. Energy Barrier at the N719-dye/CsSnI3 Interface for Photogenerated Holes in Dye-Sensitized Solar Cells. Sci. Rep. 2015, 4, 6954. 10.1038/srep06954. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mendes J. L.; Gao W.; Martin J. L.; Carl A. D.; Deskins N. A.; Granados-Focil S.; Grimm R. L. Interfacial States, Energetics, and Atmospheric Stability of Large-Grain Antifluorite Cs2TiBr6. J. Phys. Chem. C 2020, 124, 24289–24297. 10.1021/acs.jpcc.0c08719. [DOI] [Google Scholar]
- Cai Y.; Xie W.; Ding H.; Chen Y.; Thirumal K.; Wong L. H.; Mathews N.; Mhaisalkar S. G.; Sherburne M.; Asta M. Computational Study of Halide Perovskite-Derived A2BX6 Inorganic Compounds: Chemical Trends in Electronic Structure and Structural Stability. Chem. Mater. 2017, 29, 7740–7749. 10.1021/acs.chemmater.7b02013. [DOI] [Google Scholar]
- Chen W.; Pasquarello A. Band-Edge Levels in Semiconductors and Insulators: Hybrid Density Functional Theory versus Many-Body Perturbation Theory. Phys. Rev. B 2012, 86, 035134. 10.1103/PhysRevB.86.035134. [DOI] [Google Scholar]
- Chen W.; Miceli G.; Rignanese G.-M.; Pasquarello A. Nonempirical Dielectric-Dependent Hybrid Functional with Range Separation for Semiconductors and Insulators. Phys. Rev. Mater. 2018, 2, 073803. 10.1103/PhysRevMaterials.2.073803. [DOI] [Google Scholar]
- Zheng H.; Govoni M.; Galli G. Dielectric-Dependent Hybrid Functionals for Heterogeneous Materials. Phys. Rev. Mater. 2019, 3, 073803. 10.1103/PhysRevMaterials.3.073803. [DOI] [Google Scholar]
- Skone J. H.; Govoni M.; Galli G. Self-Consistent Hybrid Functional for Condensed Systems. Phys. Rev. B 2014, 89, 195112. 10.1103/PhysRevB.89.195112. [DOI] [Google Scholar]
- Shishkin M.; Marsman M.; Kresse G. Accurate Quasiparticle Spectra from Self-Consistent GW Calculations with Vertex Corrections. Phys. Rev. Lett. 2007, 99, 246403. 10.1103/PhysRevLett.99.246403. [DOI] [PubMed] [Google Scholar]
- Leppert L.; Rangel T.; Neaton J. B. Towards Predictive Band Gaps for Halide Per-ovskites: Lessons from One-Shot and Eigenvalue Self-Consistent G W. Phys. Rev. Mater. 2019, 3, 103803. 10.1103/PhysRevMaterials.3.103803. [DOI] [Google Scholar]
- Wannier G. H. The Structure of Electronic Excitation Levels in Insulating Crystals. Phys. Rev. 1937, 52, 191–197. 10.1103/PhysRev.52.191. [DOI] [Google Scholar]
- Biega R.-I.; Filip M. R.; Leppert L.; Neaton J. B. Chemically Localized Resonant Excitons in Silver-Pnictogen Halide Double Perovskites. J. Phys. Chem. Lett. 2021, 12, 2057–2063. 10.1021/acs.jpclett.0c03579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Radha S. K.; Lambrecht W. R. L.; Cunningham B.; Grüning M.; Pashov D.; van Schilfgaarde M. Optical Response and Band Structure of LiCoO2 Including Electron-Hole Interaction Effects. Phys. Rev. B 2021, 104, 115120. 10.1103/PhysRevB.104.115120. [DOI] [Google Scholar]
- Brisdon B. J.; Lester T. E.; Walton R. A. Complex Halides of Transition Metals–III Electronic Absorption Spectra of Hexahalotitanates(IV), Vanadates(IV), and Zirconates(IV). Spectrochim. Acta, Part A 1967, 23, 1969–1976. 10.1016/0584-8539(67)80083-7. [DOI] [Google Scholar]
- Qiu D. Y.; da Jornada F. H.; Louie S. G. Optical Spectrum of MoS2: Many-Body Effects and Diversity of Exciton States. Phys. Rev. Lett. 2013, 111, 216805. 10.1103/PhysRevLett.111.216805. [DOI] [PubMed] [Google Scholar]
- Marini A.; Del Sole R. Dynamical Excitonic Effects in Metals and Semiconductors. Phys. Rev. Lett. 2003, 91, 176402. 10.1103/PhysRevLett.91.176402. [DOI] [PubMed] [Google Scholar]
- Lanzani G.The Photophysics behind Photovoltaics and Photonics; John Wiley & Sons, 2012. [Google Scholar]
- de Boer D. K. G.; Haas C.; Sawatzky G. A. Exciton Satellites in Photoelectron Spectra. Phys. Rev. B 1984, 29, 4401–4419. 10.1103/PhysRevB.29.4401. [DOI] [Google Scholar]
- Acharya S.; Pashov D.; Rudenko A. N.; Rösner M.; van Schilfgaarde M.; Katsnel-son M. I. Importance of Charge Self-Consistency in First-Principles Description of Strongly Correlated Systems. npj Comput. Mater. 2021, 7, 208. 10.1038/s41524-021-00676-5. [DOI] [Google Scholar]
- Cunningham B.; Gruening M.; Pashov D.; van Schilfgaarde M. QSGW: Quasiparticle Self Consistent GW with Ladder Diagrams in W.. arXiv 2021, 2106.05759. 10.48550/arXiv.2106.05759. [DOI] [Google Scholar]
- Cunningham B.; Grüning M.; Azarhoosh P.; Pashov D.; van Schilfgaarde M. Effect of Ladder Diagrams on Optical Absorption Spectra in a Quasiparticle Self-Consistent GW Framework. Phys. Rev. Mater. 2018, 2, 034603. 10.1103/PhysRevMaterials.2.034603. [DOI] [Google Scholar]
- Freysoldt C.; Neugebauer J.; Van de Walle C. G. Fully Ab Initio Finite-Size Corrections for Charged-Defect Supercell Calculations. Phys. Rev. Lett. 2009, 102, 016402. 10.1103/PhysRevLett.102.016402. [DOI] [PubMed] [Google Scholar]
- Swift M. W.; Peelaers H.; Mu S.; Morton J. J. L.; Van de Walle C. G. First-Principles Calculations of Hyperfine Interaction, Binding Energy, and Quadrupole Coupling for Shallow Donors in Silicon. npj Comput. Mater. 2020, 6, 181. 10.1038/s41524-020-00448-7. [DOI] [Google Scholar]
- Buryi M.; Babin V.; Ligthart R. A. M.; Nagorny S. S.; Mikhailik V. B.; Vaněček V.; Prouzová Prochazková L.; Kandel R.; Nahorna V. V.; Wang P. Correlation of Emission, Scintillation and Charge Trapping Properties in Cs2HfCl6 and Cs2ZrCl6 Single Crystals. J. Mater. Chem. C 2021, 9, 2955. 10.1039/D0TC05482H. [DOI] [Google Scholar]
- Liu S.; Yang B.; Chen J.; Zheng D.; Tang Z.; Deng W.; Han K. Colloidal Synthesis and Tunable Multicolor Emission of Vacancy-Ordered Cs2HfCl6 Perovskite Nanocrystals. Laser Photonics Rev. 2022, 16, 2100439. 10.1002/lpor.202100439. [DOI] [Google Scholar]
- Wright A. D.; Buizza L. R. V.; Savill K. J.; Longo G.; Snaith H. J.; Johnston M. B.; Herz L. M. Ultrafast Excited-State Localization in Cs2AgBiBr6 Double Perovskite. J. Phys. Chem. Lett. 2021, 12, 3352–3360. 10.1021/acs.jpclett.1c00653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palummo M.; Berrios E.; Varsano D.; Giorgi G. Optical Properties of Lead-Free Double Perovskites by Ab Initio Excited-State Methods. ACS Energy Lett. 2020, 5, 457. 10.1021/acsenergylett.9b02593. [DOI] [Google Scholar]
- Xiao Z.; Meng W.; Wang J.; Mitzi D. B.; Yan Y. Searching for Promising New Perovskite-Based Photovoltaic Absorbers: The Importance of Electronic Dimensionality. Mater. Horiz. 2017, 4, 206–216. 10.1039/C6MH00519E. [DOI] [Google Scholar]
- Li Z.; Kavanagh S. R.; Napari M.; Palgrave R. G.; Abdi-Jalebi M.; Andaji-Garmaroudi Z.; Davies D. W.; Laitinen M.; Julin J.; Isaacs M. A.; Friend R. H.; Scanlon D. O.; Walsh A.; Hoye R. L. Z. Bandgap Lowering in Mixed Alloys of Cs2Ag(SbxBi1-x)Br6 Double Perovskite Thin Films. J. Mater. Chem. A 2020, 8, 21780–21788. 10.1039/D0TA07145E. [DOI] [Google Scholar]
- Feldmann S.; Gangishetty M. K.; Bravić I.; Neumann T.; Peng B.; Winkler T.; Friend R. H.; Monserrat B.; Congreve D. N.; Deschler F. Charge Carrier Localization in Doped Perovskite Nanocrystals Enhances Radiative Recombination. J. Am. Chem. Soc. 2021, 143, 8647–8653. 10.1021/jacs.1c01567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kresse G.; Furthmüller J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Blöchl P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
- Heyd J.; Scuseria G. E.; Ernzerhof M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. 10.1063/1.1564060. [DOI] [Google Scholar]
- Kavanagh S. R.Vaspup2.0; 2022.
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Ini-tio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Ehrlich S.; Goerigk L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
- Ganose A. M.; Jackson A. J.; Scanlon D. O. Sumo: Command-line Tools for Plotting and Analysis of Periodic Ab Initio Calculations. Journal of Open Source Software 2018, 3, 717. 10.21105/joss.00717. [DOI] [Google Scholar]
- Whalley L. D. Effmass. An Effective Mass Package. Journal of Open Source Software 2018, 3, 797. 10.21105/joss.00797. [DOI] [Google Scholar]
- Varrassi L.; Liu P.; Yavas Z. E.; Bokdam M.; Kresse G.; Franchini C. Optical and Excitonic Properties of Transition Metal Oxide Perovskites by the Bethe-Salpeter Equation. Phys. Rev. Mater, 2021, 5, 074601. 10.1103/PhysRevMaterials.5.074601. [DOI] [Google Scholar]
- Tal A.; Liu P.; Kresse G.; Pasquarello A. Accurate Optical Spectra through Time-Dependent Density Functional Theory Based on Screening-Dependent Hybrid Functionals. Phys. Rev. Res. 2020, 2, 032019. 10.1103/PhysRevResearch.2.032019. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





