Seok Yeong Hong
a,
David Sunghwan Lee
a,
Hyong Joon Lee
a,
Ki-Ha Hong
*b and
Sang Hyuk Im
*ac
aDepartment of Chemical and Biological Engineering, Korea University, Seoul, 02841, Republic of Korea. E-mail: imromy@korea.ac.kr
bDepartment of Materials Science and Engineering, Hanbat National University, Daejeon 34158, Republic of Korea. E-mail: kiha.hong@hanbat.ac.kr
cPerolight, New Engineering Hall, 45 Anam-ro, Seongbuk-gu, Seoul 02841, Republic of Korea
First published on 21st July 2025
All-inorganic metal halide perovskites have attracted increasing attention as promising materials for inorganic perovskite solar cells, owing to their superior thermal and chemical stability and wider bandgap tunability compared to organic–inorganic hybrid perovskites. However, inorganic perovskite has the problem of phase instability, as these materials are prone to undesirable phase transitions induced by both intrinsic and extrinsic factors. Such instability significantly degrades the long-term stability and efficiency of inorganic perovskite solar cells. In this review, we provide a comprehensive overview of Pb-, Sn-, and Pb–Sn-based inorganic perovskite solar cells. We systematically address various intrinsic and extrinsic factors that induce phase transitions and degrade the stability of inorganic perovskite. In particular, we focus on polymorphic transitions and examine how α/β/γ to δ transformations determine the optoelectronic properties in CsPbX3, CsSnX3 and their Pb–Sn alloys. In addition, we introduce recent strategies to suppress or mitigate these factors in order to develop long-term stable and highly efficient inorganic perovskite solar cells and inorganic perovskite-based tandem solar cells.
Broader contextThis review explores recent advances and key challenges related to inorganic perovskite materials. Inorganic perovskite materials, which eliminate volatile organic cations such as methylammonium (MA) and formamidinium (FA), have attracted considerable interest in the field of perovskite photovoltaics due to their superior thermal stability compared to hybrid perovskites. However, despite these advantages, issues such as phase transitions and poor phase stability continue to limit the long-term durability and power conversion efficiency of inorganic perovskite solar cells. In this review, we comprehensively analyze the intrinsic and extrinsic factors that trigger phase transitions and instability in Pb-, Sn-, and Pb–Sn-based inorganic perovskites-including lattice distortion, tolerance factor mismatch, ion migration, moisture exposure, and light-induced degradation. We further examine recent advances in mitigating these challenges through strategies such as compositional engineering, additive incorporation, surface and interface passivation, and heterostructure engineering. By categorizing and discussing these approaches in the context of both long-term stability and efficiency enhancement, this review aims to provide valuable insight for the development of robust inorganic perovskite solar cells and all-inorganic tandem solar cells with improved operational performance. |
Replacing organic components with the inorganic cesium (Cs) cation to create all-inorganic CsPbX3 perovskites (where X = Cl, Br, I, or mixed halogens) represents a viable and effective strategy to overcome decomposition problems.14 All-inorganic CsPbX3 perovskites demonstrate significantly improved heat resistance, with thermal decomposition occurring only above 400 °C,15,16 unlike their hybrid counterparts. Given its suitable bandgap (∼1.7 eV, which is the closest among CsPbX3 compositions to the Shockley–Queisser limit for single-junction solar cells) and low-cost solution processability, CsPbI3 has been recognized as one of the most promising candidates to overcome the instability of hybrid perovskites and drive perovskite photovoltaics toward industrial viability (Fig. 1).16,17
Despite its potential, CsPbI3 faces a critical challenge in that its perovskite black phases are thermodynamically metastable at room temperature. CsPbI3 crystallizes in a cubic (α) phase at high temperature (∼320 °C) but under ambient conditions (especially in the presence of moisture), freshly formed black-phase CsPbI3 readily undergoes a phase transition to the yellow δ-phase, which is a non-perovskite structure with a larger bandgap and inferior charge-transport properties.17,18
Parallel to CsPbI3, significant attention has also been directed toward its lead-free analog, CsSnI3. CsSnI3 addresses environmental and toxicity concerns associated with lead-based materials and offers a narrower bandgap closer to the ideal Shockley–Queisser (SQ) limit (∼1.3 eV).19 While CsSnI3 also experiences the undesirable δ-phase transformation, its stability issues are further exacerbated by the easy oxidation of Sn2+ to Sn4+, causing deep trap formation and rapid performance degradation in practical environments.20–22
Using Cs(Pb,Sn)I3 solid solutions has emerged as a strategy to overcome these limitations by combining the enhanced stability imparted by lead-based frameworks with the favorable bandgap tuning provided by Sn substitution in the range of 1.35–1.76 eV.23 Recent studies, particularly using compositions such as CsPb0.7Sn0.3I3, have demonstrated significant advancements, achieving a record efficiency of up to 17.55%, the highest reported to date for inorganic Pb–Sn alloyed perovskite solar cells with bandgaps below 1.35 eV. This notable efficiency underscores their potential for excellent spectral matching in tandem solar cell configurations, thus driving intense research interest.24 Although Cs(Pb,Sn)I3 demonstrates improved phase stability and oxidation resistance relative to pure CsSnI3, significant challenges remain, particularly in managing Sn2+ oxidation and ensuring long-term operational stability.25,26
Most previous reviews have treated Pb-, Sn-, or Pb–Sn-based inorganic perovskite individually, with few providing a comprehensive discussion that encompasses all three compositions. In addition, the effects of various degradation factors on the performance of inorganic perovskite solar cells have generally been addressed briefly, with previous reviews primarily focusing on specific solution strategies.
In this review, we systematically discuss the issue of the phase transition, which is one of the primary causes of reduced stability and efficiency in inorganic perovskites, along with various intrinsic and extrinsic factors that affect stability. We provide the first cross-cutting roadmap that systematically maps the α/β/γ ↔ δ transformations of Pb-, Sn-, and Pb–Sn-alloys to strain, defect chemistry, and interface design. Furthermore, we classify and discuss recent strategic approaches for improving the performance of Pb-, Sn-, and Pb–Sn-based inorganic perovskite solar cells from the perspectives of long-term stability enhancement and efficiency enhancement. In addition, we summarize research on tandem solar cells employing inorganic perovskites to overcome the efficiency limit of single-junction solar cells. Finally, we present our outlook on the future development direction of inorganic perovskite solar cells.
![]() | ||
Fig. 2 Phase transitions in inorganic halide perovskites. (a) Structural phase transition of CsPbI3 (ref. 27) (reproduced with permission from Science, 2019, 365, 679–684. Copyright 2019 AAAS), (b) Absorbance spectra of black and yellow phases of CsPbI3 thin films29 (reproduced with permission from the Journal of Materials Chemistry A, 2015, 3, 19688–19695. Copyright 2015 Royal Society of Chemistry), (c) diagram of the formation enthalpy of α- and δ-CsPbI3 and the enthalpy change of the δ to α phase transition (reproduced with permission from the Journal of the American Chemical Society, 2019, 37, 14501–14504. Copyright 2019 American Chemical Society), and (d) phase transition of CsPbI3 at varying P–T28 (reproduced with permission from Nature Communications, 2021, 12, 461. Copyright 2021 Springer Nature).30 (e) Thermodynamic changes of phase transitions for the formation and stabilization processes of FAPbI3 and CsPbI3 metastable perovskite phases31 (reproduced with permission from Cell Reports Physical Science, 2024, 5, 101825. Copyright 2024 Cell Press). |
In CsPbI3, the cubic α-phase (black polymorph) is only stable above ∼330 °C. As temperature decreases, CsPbI3 transitions from α (cubic) to β (tetragonal) at ∼281 °C and then to γ (orthorhombic perovskite) at ∼184 °C. All three α, β, and γ phases are photoactive black phases with band gaps ∼1.7–1.75 eV, suitable for optoelectronics.32 However, at room temperature the perovskite γ-phase is metastable and will slowly convert to the δ-phase (yellow), which is a lower-symmetry non-perovskite structure with a larger band gap and poor electronic properties (Fig. 2b).29,33
Similarly, CsSnI3 was experimentally reported to exhibit polymorphism with three distinct metastable black-phase structures. Specifically, CsSnI3 undergoes two-step phase transitions during cooling: from cubic (α-CsSnI3, space group Fmm, stable above 440 K) to tetragonal (β-CsSnI3, space group P4/mbm, stable between 362 and 440 K) and then to orthorhombic (γ-CsSnI3, space group Pnma, stable below 362 K).34 Below room temperature, the material eventually transforms into a yellow non-perovskite δ-phase upon prolonged exposure to ambient conditions or moisture. In essence, both CsPbI3 and CsSnI3 share the trait that their perovskite-structured phases are metastable at room temperature, in contrast to analogous inorganic perovskites with smaller anions (e.g. CsPbBr3 and CsPbCl3) which remain perovskites under ambient conditions (albeit with distorted orthorhombic symmetry).35–37
From a thermodynamic perspective, phase stability is determined by the free energy G(T,P) of each polymorph. The cubic α-phase, while entropically favored at high T, has a higher enthalpy at room temperature than the δ-phase, making α metastable (local minimum) and δ the global minimum under ambient conditions. Using RT solution calorimetry and differential scanning calorimetry, Wang et al. reported the formation enthalpies (ΔH) of the α and δ phases of CsPbI3 (ΔHα = −2.83 ± 0.90 kJ mol−1, ΔHδ = −16.93 ± 0.87 kJ mol−1), as depicted in Fig. 2c.30 The conversion of the δ phase to the α phase is associated with an enthalpy change of 14.10 ± 0.24 kJ mol−1 and an entropy change of 23.78 J (mol K)−1. The phase transition from α (or β/γ) to δ is typically first-order, involving a discontinuous change in the structure and volume and often a latent heat. Kinetics play a key role: fast cooling (thermal quenching) can kinetically trap the material in a metastable black phase at room temperature by preventing the rearrangement required for the δ-phase.17,32 Indeed, quenching CsPbI3 from high temperature can retain β/γ-CsPbI3 at room temperature, as the rapid change induces internal strain that raises the energy barrier for the transition.27 However, given sufficient time or slight stimuli, the trapped β/γ phase will still relax to δ-CsPbI3 as this allows strain release. Pressure can also influence these thermodynamics: applying external pressure generally favors denser phases and can alter the transition temperature. Orthorhombic γ-CsPbI3 has been successfully stabilized through the combined application of moderate pressure (0.1–0.6 GPa) and heat treatment, with the phase remaining intact upon decompression under ambient conditions. The interrelation of phase transitions as a function of pressure and temperature parameters is illustrated in Fig. 2d.28 Thus, by controlling temperature, pressure, and composition, researchers can tune both the energy barriers and relative phase energies, enabling access to different phases and stabilizing high-temperature phases that are metastable under room conditions. Composition (such as mixing Pb/Sn or halide content) similarly shifts phase stabilities, as discussed next, by altering the lattice dimensions and energetics (Fig. 2e).31,38,39
![]() | ||
Fig. 3 Goldschmidt's tolerance factor, geometrical effects and phonon effects on stability. (a) The structural description of an ideal ABX3-type perovskite structure40 (reproduced with permission from Solar RRL, 2020, 4, 2000513. Copyright 2020 Wiley-VCH GmbH). (b) The stable area prediction for Sn-based halide perovskites with regard to the tolerance factor (TF, 0.875 ≤ TF ≤ 1.06) and octahedral factor (m, 0.41 ≤ μ ≤ 0.895)40 (reproduced with permission from Solar RRL, 2020, 4, 2000513. Copyright 2020 Wiley-VCH GmbH). (c) Classification using the machine learned new tolerance factor43 (reproduced with permission from Science Advances, 2019, 5, eaav0693. Copyright 2019 AAAS). (d) Phonon-driven polymorphic phase transitions among black phases (α, β, and γ) of CsPbI3, highlighting small energy differences that govern structural stability45 (reproduced with permission from ACS nano, 2018, 12, 3477–3486. Copyright 2018 American Chemical Society). (e) Comparison of the impact of different exchange correlations to calculate phase transition temperatures46 (reproduced with permission from Chemistry of Materials, 2022, 34, 8561–8576. Copyright 2022 American Chemical Society). |
Despite these advancements, approaches based on tolerance factors continue to demonstrate fundamental limitations. Inconsistencies arise from different research groups employing varied ionic radius values and stability criteria, creating confusion in predicting perovskite structural stability. Indeed, several studies have erroneously predicted CsPbI3 stability based on tolerance factor calculations alone.47 Most notably, the approach relies purely on geometric and ionic considerations, disregarding important electronic and chemical effects such as orbital hybridization, covalency, defect chemistry, and anharmonic lattice dynamics. These chemical and electronic factors critically influence stability in real systems, especially for compounds with significant covalent bonding character or polarizable ions, thus limiting the universal applicability of tolerance factors alone.
Within the harmonic approximation, lattice vibrations are assumed to be small oscillations around equilibrium positions with linear restoring forces. In this idealized scenario, phonon dispersion calculations using density functional theory (DFT) reveal soft-mode instabilities characterized by imaginary frequencies, especially prominent at the Brillouin zone boundaries.45,48 Such instabilities indicate that the cubic high-symmetry phases (α-phase for CsPbI3 and CsSnI3) are inherently unstable at low temperatures, predicting spontaneous distortions to lower-symmetry orthorhombic or tetragonal structures.
However, the purely harmonic description does not fully capture temperature-dependent stability. This discrepancy is resolved by considering anharmonic phonon effects—nonlinearities in lattice vibrations arising from significant atomic displacements at finite temperatures. Anharmonicity manifests notably in phonon–phonon interactions, double-well potentials, and temperature-dependent frequency shifts, critically influencing the thermodynamic stability of phases as shown in Fig. 3d.45,49,50 This dynamic stabilization explains the persistence of metastable black phases under conditions far from equilibrium, such as rapid quenching or lattice strain. Understanding phonon-driven stability mechanisms offers practical strategies for enhancing phase stability. For instance, compositional tuning through doping or alloying can modify phonon spectra, reducing anharmonic instabilities and suppressing undesirable phase transitions. Additionally, controlled lattice strain and microstructural engineering can selectively stabilize desired perovskite phases by altering vibrational modes and entropy contributions.51 Accurate modeling of phonon properties is essential; recent research combining random-phase approximation (RPA) calculations with machine learning potentials (MLPs) has demonstrated significant sensitivity in predicted phase transition temperatures. For example, employing different exchange–correlation (XC) functionals or inadequate structural optimizations can cause deviations in predicted transition temperatures exceeding 100 K. Specifically, precise calculations using RPA combined with MLP corrections yielded transition temperatures within 50 to 100 K of experimental values, highlighting the necessity of carefully chosen computational methods for accurate stability predictions (Fig. 3e).46
![]() | ||
Fig. 4 Role of point defects. (a) Layer by layer phase transition processes between the γ-phase and MS1-phase (intermediate states having higher energy barriers)52 (reproduced with permission from Advanced Functional Materials, 2024, 34, 2308246. Copyright 2024 Wiley-VCH GmbH). (b) Transition barrier lowering by the formation of an iodine vacancy52 (reproduced with permission from Advanced Functional Materials, 2024, 34, 2308246. Copyright 2024 Wiley-VCH GmbH). (c) The conversion of subdomains from cubic to orthorhombic phases in defect-rich or fewer NCs when cooling down54 (reproduced with permission from Chemistry of Materials, ACS Materials Letters, 2019, 1, 185–191. Copyright 2019 American Chemical Society). (d) Schematic illustration of reducing the energy change from the d to the a phase by controlling Cs+ and Pb2+ vacancies55 (reproduced with permission from Chemistry of Materials, The Journal of Physical Chemistry C, 2019, 123, 9735–9744. Copyright 2019 American Chemical Society). |
Experimental observations strongly support the above picture. Iodine vacancies tend to form readily under iodine-poor or destabilizing conditions (owing to their relatively low formation energy), and their presence correlates with accelerated phase degradation. For example, moisture is known to catalyze the black-to-yellow transition in CsPbI3 by leaching iodide from the surface (via HI or I2 formation), thereby amplifying the surface VI concentration.53 In contrast, vacancies on the cation sublattices (A-site Cs or B-site Pb vacancies) may have quite different effects. Cs vacancies, in particular, tend to stabilize the perovskite framework rather than destabilize it (Fig. 4d). Thermodynamic calculations by Kye et al. showed that the introduction of cation vacancies weakens the coupling between Cs+ ions and the PbI6 octahedral network, thereby lowering the energy difference between the cubic/orthorhombic perovskite phase and the δ-phase.55 However, Kye's theoretical study extrapolates macroscopic phase stability from single-defect formation energies, neglecting defect–defect interactions; moreover, the very high vacancy concentrations required for full stabilization have yet to be experimentally verified.
Lin et al. quantified the kinetics of this moisture-induced phase change: higher relative humidity dramatically increases the nucleation rate of the yellow phase, accelerating the black → yellow conversion (Fig. 5a and b).58 Crucially, water itself is not consumed into new compounds (no permanent hydrates are formed); instead, it acts as a catalyst. Dastidar et al. measured a black → yellow phase-change enthalpy of ∼14.2 kJ mol−1 and found that under dry inert conditions the trapped black phase remains stable up to ∼100 °C.56 The presence of water vapor did not alter the phase equilibrium enthalpy, but sped up the transformation, confirming that moisture primarily catalyzes the phase transition rather than shifting the thermodynamics.56 Interestingly, δ-CsPbI3 can be easily converted into a perovskite phase by annealing, which is the distinguished feature of inorganic perovskites with hybrid perovskites. This reversible characteristic of CsPbI3 enables its application in thermochromic solar cells (Fig. 5c and d).59
![]() | ||
Fig. 5 Effect of moisture on inorganic perovskites. (a) PL micrographs of CsPbI3 at 90% RH over time under 375 nm laser excitation. Dotted lines: a guide to the eye highlighting the emergence of low-T phase crystals. Scale bars, 20 μm (ref. 58) (reproduced with permission from Matter, 2021, 4, 2392–2402. Copyright 2021 Cell Press). (b) Dependence of the nucleation rate (JN) on RH. Inset: the nucleation rate shown in the log plot58 (reproduced with permission from Matter, 2021, 4, 2392–2402. Copyright 2021 Cell Press). (c) Photograph of the low-T phase and high-T phase (orange-red-colored) thin films59 (reproduced with permission from Nature Materials, 2018, 17, 261–267. Copyright 2018 Springer Nature). (d) The stable and reversible switching of the absorption (550 nm) of the three CsPbIBr2 thin films over 100 phase transition cycles. (e) Schematic of VI and H2O driven fast δ-CsPbI3 growth59 (reproduced with permission from Nature Materials, 2018, 17, 261–267. Copyright 2018 Springer Nature). (f) Variation of H (enthalpy) during the γ-to-δ phase transition in the presence of H2O, H+, and OH− within CsPbI3. The shaded curve represents the pristine case60 (reproduced with permission from Chemistry of Materials, 2023, 35, 2321–2329. Copyright 2023 American Chemical Society). |
Surface iodide vacancies serve as nucleation sites for δ-CsPbI3 growth, with moisture amplifying these vacancies by strongly solvating halide ions at the interface (Fig. 5e).53 Wylie et al. showed that surface treatment strategies are effective in improving moisture resistance, with CsI and CdI2 treatments reducing the phase transition rate by approximately 5-fold by filling surface iodide vacancies.53
Theoretical investigation revealed that H2O and OH− largely affect the kinetics by lowering the activation energy, whereas H+ has a negligible effect on the kinetics by calculating the phase transition barrier with DFT (Fig. 5f).60 For CsSnI3, moisture effects have not been thoroughly investigated as much as in the case of CsPbI3, yet. Yang et al.'s DFT calculations presented that water adsorbed on the surface of γ-phase CsSnI3 could induce charge transfers between the H atom in the water and adjacent I atoms and between the O atom and Cs+, leading to distortion of SnI6 octahedra and structural deformation, which can be the plausible origin of phase instability.61
The recognition of water as a catalyst (rather than a reactant) was key to devising mitigation strategies: theory indicated that simply keeping films dry or introducing water-resistant interfaces could preserve the metastable black phase. Indeed, encapsulation and surface passivation (to prevent H2O adsorption) emerged as effective measures to maintain black-phase stability in humid environments, aligning with the computational predictions.
By comparison, tin-based perovskites are extremely oxygen-sensitive. Oxygen exposure induces a chemical oxidation: Sn2+ (in the black perovskite) oxidizes to Sn4+, which destabilizes the perovskite lattice and generates a yellow or amorphous phase. This manifests as rapid degradation of the optoelectronic properties of CsSnI3 in air. The pronounced oxygen sensitivity of CsSnI3 stems from the synergy between redox chemistry and defect formation in its Sn-based lattice. Frost diagrams indicates that Pb2+ is in a deep asymmetric thermodynamic sink while Sn2+ is in a shallow sink (Fig. 6a), which can lead to oxidation pathways for Sn2+.63 Therefore, adsorbed O2 or O2/H2O rapidly extracts electrons and converts Sn2+ to Sn4+, which can result in VSn and lower the defect formation energy barrier. Tin vacancies raise the native p-doping level and accommodate the oxidized Sn4+ ions.
![]() | ||
Fig. 6 Effect of oxygen and light irradiation on inorganic perovskites. (a) Frost diagrams for Pb and Sn under standard conditions63 (reproduced with permission from ACS Materials Letters, 2021, 3, 299–307. Copyright 2021 American Chemical Society). (b) Cyclic degradation mechanism of a tin iodide perovskite under ambient air exposure64 (reproduced with permission from Nature Communications, 2021, 12, 2853. Copyright 2021 Springer Nature). (c) Growth of Cs2SnI6 from CsSnI3 (ref. 65) (reproduced with permission from Energy & Environmental Science, 2019, 12, 1495–1511. Copyright 2019 Royal Society of Chemistry). (d) Energy levels of Pb- and Sn–iodide perovskites. Adapted with minor modifications66 (reproduced with permission from Nature Communications, 2019, 10, 2560. Copyright 2019 Springer Nature). (e) Confocal false-color images with the red and blue colors corresponding to emission channels >600 and <600 nm (ref. 67) (reproduced with permission from Matter, Matter, 2022, 5, 1455–1465. Copyright 2022 Cell Press). (f) The total area of the I-rich regions plotted as a function of time at carrier generate rates of G1 (blue line) and G2 (red line), respectively, in comparison with the experimental data (black dots)68 (reproduced with permission from Matter, Matter, 2023, 6, 2052–2065. Copyright 2023 Cell Press). |
Oxygen can trigger cyclic degradation of tin-based perovskites. Haque et al. revealed that oxidized Sn4+ complexes with I−, releasing SnI4; SnI4 hydrolyses to HI + SnO2, and HI + O2 ⇒ I2 + H2O. I2 is itself a strong oxidant toward remaining Sn2+, creating a cyclic degradation loop (Fig. 6b).64 Depletion of Sn destabilizes the already metastable black perovskite and drives it toward yellow δ-CsSnI3 and then vacancy-ordered Cs2SnI6 or direct conversion into yellow Cs2SnI6 (Fig. 6c).65 The high-lying Sn 5s and I 5p valence bands also facilitate hole transfer to O2, so easy Sn oxidation, self p-doping, iodine autocatalysis, lattice metastability and favorable band energetics together make CsSnI3 far more air-labile than its Pb counterpart (energy levels can be found in Fig. 6d).66
Another phenomenon observed under continuous illumination is halide redistribution in mixed-halide perovskites. In CsPb(I1−xBrx)3, strong light causes iodine-rich and bromine-rich regions to segregate, leading to local bandgap changes.69,70 Peng et al. conclusively showed that light-induced Br/I segregation in single-crystalline CsPbBr2.1I0.9 proceeds by spinodal decomposition, quantifying both the carrier-dependent kinetics and the nanometre-scale coarsening length using in situ cryo-STEM cathodoluminescence and phase-field modelling (Fig. 6f).68
Although this halide segregation is not directly connected with phase transformation, the ensuing I-rich micro-domains promote iodide-vacancy accumulation and local strain, which under severe operating conditions can nucleate the yellow δ-phase, so halide segregation—though structurally discrete—may indirectly accelerate phase degradation.
As mentioned above, phase transitions and instability in inorganic perovskites are induced by a variety of intrinsic and extrinsic factors, which in turn contribute to the degradation of both long-term operational stability and device efficiency. Accordingly, mitigating these instability-inducing factors, such as lattice distortions, defect formation, and environmental stressors, is essential for enhancing the long-term stability and photovoltaic performance of inorganic perovskite solar cells.
No. | Strategy | Structure | Type | PCE | Stability | Ref. |
---|---|---|---|---|---|---|
1 | A-site engineering | FTO/TiO2/CsPbI3−xBrx:TrMA/P3HT/Ag | NIP | 20.59 | 84% for 192 h@T: 85 °C/ambient air/non-encapsulated | 71 |
91% for 3055 h@ambient air/non-encapsulated | ||||||
97% for 2071 h@N2 | ||||||
2 | B-site engineering | FTO/TiO2/PCBA/CsPb0.95Ge0.05I3/spiro-OMeTAD/Au | NIP | 19.52 | 85.5% for 3000 h@MPP/1 sun illumination/N2 | 72 |
92% for 1400 h@ambient air/non-encapsulated | ||||||
3 | FTO/PCBM/CsSn0.5Ge0.5I3/spiro-OMeTAD/Au | NIP | 7.11 | 90% for 500 h@MPP/1 sun light illumination/N2 | 73 | |
4 | FTO/c-TiO2/mp-TiO2/CsSn1−xZnxI3/Al2O3/NiOx/carbon | PIN | 8.27 | 86% for 216 h@ambient air/non-encapsulated | 74 | |
90% for 30 d@N2 | ||||||
5 | ITO/SnO2/CsPb0.7Sn0.3IBr2:ZnC2O4/spiro-OMeTAD/Au | PIN | 14.1 | 450 h@T: 80 °C/N2 | 75 | |
75% for 240 h@ambient air/non-encapsulated | ||||||
1350 h@N2 | ||||||
6 | X-site engineering | FTO/TiO2/Cs2PbI2Cl2/CsPbI3/spiro-OMeTAD/Au | NIP | 20.6 | 80% for 1000 h@MPP/1 sun illumination/N2 | 76 |
7 | ITO/SnO2/CsPbI3:NH4PbCl2.8Br0.2/spiro-OMeTAD/MoO3/Ag | NIP | 20.2 | 95% for 1000 h@MPP/1 sun illumination/ | 77 | |
80% for 250 h@T: 85 °C RH: 85% | ||||||
8 | FTO/NiOx/CsPb0.6Sn0.4I3:SnF2*3FACl/4AMPI2/PCBM/BCP/Ag | PIN | 13.37 | 70% for 1045 h@MPP/1 sun illumination/T: 45 °C/N2 | 78 | |
77% for 100 h@T: 85 °C/N2 | ||||||
100% for 2800 h@N2 | ||||||
9 | ITO/PEDOT:PSS/CsPb0.55Sn0.45I2Br:CsCl/PbSO4/PCBM/BCP/Ag | PIN | 10.39 | 93% for 300 h@MPP/1 sun illumination/N2 | 79 | |
80% for 300 h@MPP/1 sun illumination/T: 25 °C RH: 40–50%/non-encapsulated | ||||||
95% for 300 h@T: 85 °C/N2 | ||||||
92.5% for 2000 h@N2 |
![]() | ||
Fig. 7 The review of stability enhancement via composition engineering with (a) trimethylammonium in CsPbI3−xBrx (reproduced with permission from ACS Energy Lett., 8(5), 2284–2291. Copyright 2023 American Chemical Society),71 (b) NH4PbX3 in CsPbI3 perovskite (reproduced with permission from Nano Energy, 132, 110396. Copyright 2024 Elsevier)77 and (c) CsCl in CsPbI3 perovskite (reproduced with permission from Cell Rep. Phys. Sci., 5(5), 101935. Copyright 2024 Elsevier).76 |
![]() | ||
Fig. 8 The review of stability enhancement via composition engineering with (a) GeI2 in CsPbI3 (reproduced with permission from Adv. Energy Mater., 12(10), 2103690),72 (b) Zn in CsSnI3 (reproduced with permission from J. Mater. Chem. A, 10, 23204–23211. Copyright 2022 Royal Society of Chemistry),74 and (c) GeI2 in CsPbI3 (reproduced with permission from Adv. Energy Mater., 12(10), 2103690. Copyright 2022 Wiley-VCH GmbH).73 |
Wang et al. proposed a facile and effective galvanic displacement reaction approach for stable CsSnI3 perovskite solar cells.74 Metallic Zn powder was introduced into the perovskite precursor film, which behaves as a favorable redox couple with Zn/Zn2+ and Sn2+/Sn4+ ions. The introduction of zinc powder thus successfully reduces the undesired Sn4+ species within the precursor solution and in turn results in reduced trap states in the final perovskite film. Moreover, the galvanic displacement reaction simultaneously incorporates small fractions of Zn2+ ions, saturating at 5% in the final perovskite lattice, affording enhanced inherent stability, as shown in Fig. 8b. The zinc incorporated CsSnI3 perovskite film achieved 8.27% PCE with a carbon electrode and retained 86.3% PCE after 216 hours of storage in air without encapsulation.
Padture et al. demonstrated a stable lead-free perovskite light absorber through Sn and Ge alloys. A solid solution perovskite of CsSnxGe1−xI3 was prepared to stabilize the Goldschmidt tolerance factor and octahedral factor.73 The optimal inherent structural and chemical stability was achieved in CsSn0.5Ge0.5I3, which retained the perovskite crystal structure even after exposure to humid air for 72 hours as shown in Fig. 8c. The CsSn0.5Ge0.5I3 film was prepared through vacuum thermal evaporation of the alloy powder, and the successful Ge incorporation in the resulting film was confirmed with the optical bandgap reaching 1.50 eV. Notably, the Ge incorporated film spontaneously forms native GeO2 oxide at the surface when exposed to air, providing a passivation layer against detrimental oxygen and moisture. Accordingly, the CsSn0.5Ge0.5I3 alloy perovskite solar cell exhibited promising storage stability in air without encapsulation, maintaining 91% of its initial PCE after 100 hours, and demonstrated excellent operational stability, maintaining 92% of its initial PCE after 500 hours of continuous operation in an inert environment.
Min et al. also adopted halide compositional engineering by introducing CsCl within the precursor solution for stability enhancement of CsPbI3 perovskite solar cells.76 According to the time-of-flight secondary ion mass spectrometry spectra and X-ray diffraction pattern, the chloride ions remain at the bottom buried layer during the film formation stage, establishing a spontaneous Cs2PbI2Cl2/CsPbI3 interface. The as-grown CsPbI3 film exhibits suppressed trap density and superior film crystallinity, contributing to the film stability. Moreover, the superior thermodynamic stability of the Cs2PbI2Cl2 interlayer inhibits the δ-CsPbI3 phase transition predominant at the buried interface during exposure to humid environments as shown in Fig. 7c. Thus, the CsPbI3 perovskite solar cell with CsCl incorporation exhibits 20.6% PCE, maintaining approximately 80% of its initial performance after 1000 hours of light exposure in a N2 environment.
No. | Strategy | Structure | Type | PCE | Stability | Ref. |
---|---|---|---|---|---|---|
1 | Antioxidant | FTO/TiO2/CsPbI3:FBTH/spiro-OMeTAD/Au | NIP | 21.41 | 85.63% for 200 h@MPP/1 sun illumination/encapsulated | 80 |
93.1% for 280 h@T: 85 °C/N2 | ||||||
92.6% for 900 h@ambient air/non-encapsulated | ||||||
2 | ITO/NiOx/CsSnI3:CBZ/ZnO/PCBM/Ag | PIN | 11.21 | 90% for 650 h@MPP/1 sun illumination/T: 65 °C/N2 | 81 | |
100% for 60d@N2 | ||||||
3 | FTO/NiOx/CsPb0.6Sn0.4I3:4AMPI2:acetylhydrazine/PCBM/BCP/Ag | PIN | 15.04 | 90% for 1000 h@MPP/1 sun illumination/N2 | 82 | |
100% for 850 h@pure O2 T![]() ![]() |
||||||
3500 h@N2 | ||||||
4 | FTO/TiO2/CsPbI3−xBrx/OMe-SP/spiro-OMeTAD/Au | NIP | 22.2 | 90.70% for 1000 h@MPP/1 sun illumination/N2 | 83 | |
94.45% for 170 h@UV/ambient air/non-encapsulated | ||||||
95.03% for 840 h@ambient air/non-encapsulated | ||||||
5 | FTO/PEDOT:PSS/CsPb0.55Sn0.45I2.4Br0.6:acrylamide/PCBM/BCP/Ag | PIN | 14.17 | 90% for 200 h@MPP/1 sun illumination/T: 65 °C/N2 | 84 | |
70% for 500 h@T: 65 °C/N2 | ||||||
6 | Crystallization regulation | ITO/PEDOT:PSS/Me-4PACz/CsPbI3:EAL/PCBM/C60/BCP/Ag | PIN | 21.08 | 98% for 600 h@MPP/1 sun illumination/N2 | 85 |
90% for 140 h@T: 60 °C/N2 | ||||||
95% for 1680 h@N2 | ||||||
7 | FTO/TiO2/CsPbI3:[PPN][TFSI]/spiro-OMeTAD/Au | NIP | 20.64 | 88% for 1000 h@MPP/1 sun illumination/N2 | 86 | |
72% for 1000 h@T: 85 °C/N2 | ||||||
82% for 1000 h@ambient air/T: 25 °C RH:20–30%/non-encapuslated | ||||||
8 | FTO/TiO2/CsPbI3:GDY//spiro-OMeTAD/Au | NIP | 20.49 | 96.7% for 500 h@MPP/1 sun illumination/N2 | 87 | |
97.5% for 500 h@T: 65 °C/N2 | ||||||
9 | FTO/TiO2/CsPbI2.85Br0.15:PHSF/spiro-OMeTAD/Au | NIP | 20.16 | 82.9% for 600 h@MPP/1 sun illumination/N2 | 88 | |
81.7% for 65 h@T: 70 °C/ambient air/non-encapsulated | ||||||
87.7% for 60d/ambient air/non-encapsulated | ||||||
10 | ITO/NiiOx/CbzNaph/CsPb(IxBr(1−x))3:(AQS:FPEA)/C60/BCP/Ag | PIN | 18.59 | 94% for 1000 h@MPP/1 sun illumination/T: 45 °C/N2 | 89 | |
11 | ITO/PEDOT:PSS/CsPb0.5Sn0.5I2Br:PbS/PCBM/Ag | PIN | 8.03 | 90% for 400 h@Ar | 90 | |
12 | ITO/PEDOT:PSS/CsPb0.7Sn0.3I3:1-4FP/PCBM/PEI/Ag | PIN | 17.19 | 91% for 48 h@RH: 50–60%/T: ambient RT/non-encapsulated | 91 | |
Around 100% for 4000 h@N2/non-encapsulated | ||||||
13 | FTO/c-TiO2/mp-TiO2/CsSnI3:EMIMAc/Al2O3/NiOx/carbon | PIN | 8.54 | 64% for 78 h@T: RT RH: 60%/non-encapsulated | 92 | |
94% for 2160 h@N2 | ||||||
14 | Defect passivation | FTO/TiO2/CsPbI3:DED/spiro-OMeTAD/Au | NIP | 21.15 | 92.8% for 250 h@1 sun illumination/N2 | 93 |
94.9% for 1000 h@ambient air/non-encapsulated | ||||||
15 | ITO/SnO2/CsPbI2.85Br0.15:PC/ODADI/spiro-OMeTAD/Au | NIP | 22.07 | 90.89% for 450 h@MPP/1 sun illumination/T: 25 °C RH: 25%/non-encapsulated | 94 | |
92.79% for 600 h@RH 15–25%/T: 25 °C | ||||||
90.10% for 100 h@T: 65 °C/N2 | ||||||
96.34% for 1000 h@N2 | ||||||
16 | FTO/TiO2/CsPbI3:RBITC/spiro-OMeTAD/Au | NIP | 20.95 | 93.6% for 100 h@MPP/1 sun illumination/encapsulated | 95 | |
86.02% for 500 h@T: 60 °C/N2 | ||||||
17 | ITO/NiOx/CsPbI2.85Br0.15:5-MVA/PCBM/BCP/Ag | PIN | 20.82 | 90% for 500 h@T: 85 °C/N2 | 96 | |
90% for 2000 h@ambient air/T: 25 °C RH: 25%/non-encapsulated | ||||||
18 | FTO/TiO2/CsPbI3:EMIMHSO4/spiro-OMeTAD/Au | NIP | 20.01 | 95% for 1000 h@ambient air/non-encapsulated | 97 | |
19 | FTO/SnO2/CsPbI3:Zn(C6F5)2/spiro-OMeTAD/Au | NIP | 19 | 98% for 700 h@ambient air/non-encapsulated | 98 | |
20 | ITO/NiOx/CsPb0.6Sn0.4I2Br:DCD/ZnO/PCBM/Ag | PIN | 14.17 | 92% for 600 h@MPP/1 sun illumination/N2 | 99 | |
60% for 1200 h@MPP/1 sun illumination/T: 65 °C | ||||||
80% for 45 d@T: 85 °C/N2 | ||||||
95% for 60d@N2 | ||||||
21 | ITO/NiOx/CsPb0.7Sn0.3I3:(1-4FP)/PCBM/ZrAcac/Ag | PIN | 17.551 | 90% for 700 h@1 sun illumination | 24 | |
100% for 400 h@N2 | ||||||
22 | ITO/PEDOT:PSS/CsSnI3:PTM/ICBA/BCP/Ag | PIN | 10.1 | 81.3% for 2000 min@MPP/1 sun illumination/T: 70 °C, RH: 30%/encapsulated | 100 | |
83.4% for 45 d@T: RT, RH: 30%/non-encapsulated | ||||||
94.3% for 60 d@N2 | ||||||
23 | FTO/c-TiO2/mp-TiO2/CsSnI3:MBAA/P3HT/Au | NIP | 7.5 | 58.4% for 120 h@1 sun illumination/T: 45 °C, RH: 10%/non-encapsulated | 101 | |
25.2% for 600 h@T: RT, RH: 10%/non-encapsulated | ||||||
76.5% for 1440 h@N2 | ||||||
24 | FTO/c-TiO2/mp-TiO2/CsSnI3:CPT/Al2O3//NiOx/carbon | NIP | 8.03 | 86% for 120 h@T: RT, RH: 60%/encapsulated | 102 | |
90% for 3000 h@N2 | ||||||
25 | Moisture barrier | FTO/TiO2/CsPbI3:CSE/spiro-OMeTAD/Au | NIP | 21.8 | 97% for 440 h@MPP/1 sun illumination/encapsulated | 103 |
90% for 1500 h@ambient air/non-encapsulated |
![]() | ||
Fig. 9 The review of stability enhancement via additive engineering with (a) carbazide molecule in CsSnI3 (reproduced with permission from Adv. Mater., 35(26), 2300503. Copyright 2023 Wiley-VCH GmbH),81 (b) 4-fluorobenzothiohydrazide in CsPbI3 (Adv. Funct. Mater., 34(10), 2312638. Copyright 2023 Wiley-VCH GmbH),80 and (c) acetylhydrazine in CsPb0.6Sn0.4I3 (reproduced with permission from J. Energy Chem., 72, 487–494. Copyright 2022 Elsevier).82 |
Similarly, Liu et al. incorporated a 4-fluorobenzothiohydrazide (FBTH) molecule in CsPbI3 perovskite solar cells to stabilize the redox reaction during aging and to passivate defects in the final perovskite film.80 Owing to the Lewis acidic F, S, and N atoms, the FBTH additive stabilizes the precursor solution, maintaining uniform colloidal size distribution during storage and regulating the crystallization process, enhancing film morphology with larger grain size. Moreover, the FBTH molecule acts as an effective redox agent to suppress I2 and Pb0 formation during aging of the CsPbI3 perovskite film. In addition, the enhanced hydrophobicity of FBTH further contributes to the phase stability of the resulting film as shown in Fig. 9b. As a result, the CsPbI3 perovskite solar cell prepared with TBFH demonstrated 92.6% and 85.63% PCE retention during 900 hours of storage and during 200 hours of operation at the MPP, respectively.
Yang et al. reported an acetylhydrazine additive in the CsPb0.6Sn0.4I3 perovskite precursor solution to demonstrate an antioxidative solution processing method.82 The Sn2+ ions in the precursor solution are susceptible to spontaneous oxidation to the Sn4+ oxidation state, resulting in abundant deep trap states in the control film. The acetylhydrazine addition successfully reduces the Sn4+ oxidation states to desirable Sn2+ states and further prevents oxidation under thermal stress, as shown in Fig. 9c. The acetylhydrazine incorporation successfully suppresses trap states in the resulting film and further protects the resulting CsPb0.6Sn0.4I3 film from oxidation in ambient air owing to the favorable adsorption energy on the perovskite surface compared to oxygen molecules. Accordingly, the CsPb0.6Sn0.4I3 perovskite solar cell with acetylhydrazine demonstrates excellent storage stability in an N2 environment for 3500 hours and in an ambient environment for 85 hours and further exhibits remarkable operational stability, maintaining 90% of its initial PCE for 1000 hours of operation.
![]() | ||
Fig. 10 The review of stability enhancement via additive engineering with (a) CSE in CsPbI3 perovskite (reproduced with permission from ACS Energy Lett., 9(10), 4817–4826. Copyright 2024 American Chemical Society),103 (b) [PPN][TFSI] in CsPbI3,86 and (c) DED in CsPbI3 (reproduced with permission from Adv. Mater., 35(12), 2210223. Copyright 2023 Wiley-VCH GmbH).93 |
In general, post-treatment with inorganic salts104,105 and molecules106–112 or organic ionic salt113–115 containing functional groups such as NH3, S–COO−, SO−, etc. enhances the stability by passivating defects of the perovskite surface. Additionally, surface reconstruction and the formation of an environmental protection layer through post-treatment can further improve the stability of inorganic perovskites (Table 3).
No. | Strategy | Structure | Type | PCE | Stability | Ref |
---|---|---|---|---|---|---|
1 | Surface defect passivation | FTO/NiO2/MeO-2PACs/CsPbI3−xBrx/Yb(TFSI)3/PCBM/BCP/Ag | PIN | 21.4% | 90% for 1260 h@MPP/1 sun illumination/ambient air | 104 |
86% for 350 h@T: 65 °C/N2 | ||||||
2 | FTO/TiO2/CsPbI3−xBrx/BMBC/spiro-OMeTAD/Au | NIP | 21.75% | 89.3% for 120 h@T: 85 °C/N2/non-encapsulated | 107 | |
91.8% for 720 h@RH: 20–35%/T: 65 °C | ||||||
3 | FTO/TiO2/CsPbI3−xBrx/TFA/spiro-OMeTAD/Au | NIP | 21.35% | % For 120 h@T: 65 °C/N2 | 108 | |
% For 960 h@RH: 20–35%/T: ambient RT | ||||||
4 | FTO/NiOx/2PACz/CsPbI2.85Br0.15/LiF/PD/PCBM/BCP/Ag | PIN | 20.24% | 89.66% for 500 h@T: 65 °C/N2/non-encapsulated | 110 | |
97.84% for 500 h@ambient air/non-encapsulated | ||||||
96.48% for 1000 h@N2/non-encapsulated | ||||||
5 | ITO/NiOx/P3CT-N/CsPbI2.85Br0.15/BPFz/PCBM/BCP/Ag | PIN | 20.22% | 83.85% for 100 h@1 sun illumination/ambient air/non-encapsulated | 111 | |
86.45% for 500 h@T: 65 °C/N2/non-encapsulated | ||||||
89.66% for 600 h@ambient air/non-encapsulated | ||||||
96.15% for 1000 h@N2/non-encapsulated | ||||||
6 | FTO/NiOx/CsPbI3−xBrx/MMI/PCBM/BCP/Ag | PIN | 20.6% | 91% for 100 h@MPP/1 sun illumination/RH: 15–25%/T: ambient/encapsulated | 113 | |
90% for 265 h@T: 65 °C/N2/non-encapsulated | ||||||
90% for 1000 h@RH: 15–25%/T: ambient/non-encapsulated | ||||||
7 | FTO/TiO2/CsPbI3/F3EAI/spiro-OMeTAD/Au | NIP | 20.5% | 90% for 1000 h@RH: ∼20%/T: ambient/non-encapsulated | 114 | |
8 | ITO/PEDOT:PSS/CsPb0.5Sn0.5I2Br/F-TBA/PCBM/BCP/Ag | PIN | 14.01% | 90% for 900 h@MPP/1 sun illumination/T: 55 °C/Ar/non-encapsulated | 112 | |
72% for 100 h@RH: 60%/T: ambient RT/non-encapsulated | ||||||
94% for 1000 h@Ar/non-encapsulated | ||||||
9 | Surface reconstruction | ITO/SnO2/CsPbI3−xBrx/CsF/spiro-OMeTAD/Au | NIP | 21.02% | 86% for 400 h@MPP/1 sun illumination/T: 40 °C/N2/non-encapsulated | 116 |
10 | FTO/TiO2/CsPbI3/BTABr/spiro-OMeTAD/Au | NIP | 21.31% | 90% for 1000 h@RH: 10–15%/T: 25 °C/non-encapsulated | 117 | |
Around 100% for 400 h@0.85 V/LED(6500 K) illumination/N2/non-encapsulated | ||||||
11 | ITO/PEDOT:PSS/CsPb0.7Sn0.3I3/1-4FP/PCBM/PEI/Ag | PIN | 17.19% | 91% for 48 h@RH: 50–60%/T: ambient RT/non-encapsulated | 91 | |
Around 100% for 4000 h@N2/non-encapsulated | ||||||
12 | Environmental protection layer | FTO/TiO2/CsPbI3/APTES/spiro-OMeTAD/Au | NIP | 21.42% | 81% for 500 h@RH: 30%/non-encapsulated | 118 |
95% for 500 h@RH: 85%/encapsulated | ||||||
13 | FTO/NiOx/MeO-2PACz/CsPbI3−xBrx/MPTS/PCBM/BCP/Ag | PIN | 21.00% | 86% for 800 h@MPP/1 sun illumination/encapsulated | 119 | |
86% for 300 h@T: 65 °C/N2/non-encapsulated | ||||||
91% for 1000 h@RH: 40–50%/T: 25–30 °C/non-encapsulated | ||||||
14 | ITO/PTAA/CsPbI3/Omxene/CPTA/BCP/Ag | PIN | 19.69% | 93.01% for 1000 h@T: 85 °C/N2/non-encapsulated | 120 | |
72.19% for 1000 h@RH: 85%/T: 25 °C/non-encapsulated | ||||||
90.38% for 1000 h@RH: 85%/T: 85 °C/encapsulated |
![]() | ||
Fig. 11 The review of stability enhancement via surface post-treatment. (a) Post-treatment with Yb(TFSi)3 on an inorganic perovskite surface (reproduced with permission from Energy Environ. Sci., 2024, 17, 7271–7280. Copyright 2024 The Royal Society of Chemistry),104 (b) passivating surface defects by applying post-treatment with Boc-S-4-methoxy-benzyl-L-cysteine (BMBC) (reproduced with permission from Adv. Mater., 2023, 35, 2301140. Copyright 2023 Wiley-VCH GmbH),107 (c) suppressing Sn2+ oxidation and passivating defects in inorganic Pb–Sn-based perovskite through post-treatment with 4-fluorophenylcarbothioamide (F-TBA) (reproduced with permission from ACS Appl. Mater. Interfaces, 2023, 15(30), 36594–36601. Copyright 2023 American Chemical Society),112 (d) surface defect passivation through CsF post-treatment induced surface reconstruction (reproduced with permission from Nature Energy, 2023, 8, 372–380. Copyright 2023 Springer Nature),116 (e) suppression of Sn2+ oxidation and defect passivation in Pb–Sn-based inorganic perovskites through surface reconstruction induced by 1-(4-fluorophenyl) piperazine (1-4FP) additive incorporation and post-treatment (reproduced with permission from Chemical Engineering Journal, 2024, 479, 147554. Copyright 2023 Elsevier B.V.),91 and (f) formation of a moisture protection layer via surface treatment with 3-aminopropyltriethoxysilane (APTES) (reproduced with permission from Chemical Engineering Journal, 2024, 497, 154706. Copyright 2024 Elsevier B.V.).118 |
As shown in Fig. 11b, Zhang et al. proposed a method for passivating surface defects by applying post-treatment with Boc-S-4-methoxy-benzyl-L-cysteine (BMBC), a molecule containing multiple functional groups.107 The multiple Lewis bases of –NH, –S, and –CO in BMBC formed strong interaction with undercoordinated Pb2+ through Lewis base–acid reactions and suppressed the formation of halide vacancies. Additionally, the tert-butyl group, which exhibits hydrophobic properties, was uniformly distributed on the surface, effectively preventing moisture penetration. The BMBC-treated unencapsulated device exhibited high thermal stability, maintaining 89.3% of its initial efficiency for 120 h at 85 °C in a nitrogen-filled glovebox. Additionally, it demonstrated excellent storage stability, retaining 91.8% of its initial efficiency for 720 h at a relative humidity of 20–35% at ambient temperature.
As shown in Fig. 11c, Zhang et al. suggested a strategy to suppress Sn2+ oxidation and passivate defects in inorganic Pb–Sn perovskite through post-treatment with 4-fluorophenylcarbothioamide (F-TBA).112 The CS and NH2 groups of F-TBA formed strong coordinated interactions with Sn2+, suppressing its oxidation and consequently reducing defects formed by Sn4+. In addition, the hydrophobic nature of fluorine enhanced moisture resistance. The F-TBA-modified cell retained 90% of its original efficiency after continuous operation at the MPP at 55 °C for 900 h and maintained over 72% of its initial performance for 100 h at 60% relative humidity at ambient temperature.
Tan et al. proposed a method for passivating surface defects by applying post-treatment with benzyl trimethylammonium bromide (BTABr).117 After post-treating the CsPbI3 perovskite surface with BTABr and applying low-temperature annealing, Br− ions diffuse into the CsPbI3 film, passivating undercoordinated Pb2+ and inhibiting Pb cluster formation in both the bulk and on the surface of CsPbI3. Moreover, through I/Br ion exchange, a gradient BTA+–CsPbI3−xBrx heterostructure was formed on the surface, improving energy level alignment. The BTABr-treated device exhibited good storage stability, maintaining over 90% of its initial efficiency for 1000 h under ambient conditions with 10–15% RH and 25 °C. Furthermore, it retained nearly 100% of its performance for 400 h under continuous white LED (6500 K) illumination and at a bias voltage of 0.85 V in a nitrogen-filled glove box.
As shown in Fig. 11e, Zhang et al. suggested a method to suppress Sn2+ oxidation and passivate defects by simultaneously applying 1-(4-fluorophenyl) piperazine (1-4FP) additive incorporation and post-treatment to Pb–Sn based inorganic perovskites.91 The surface post-treatment with 1-4FP reconstructed the film, transforming the pinhole-rich and poorly covered Pb–Sn inorganic perovskite film into a more uniform and smoother film. Additionally, 1-4FP passivated the Sn vacancy defect by inhibiting the oxidation of Sn2+ through strong interaction with Sn2+. The 1-4FP-treated perovskite film exhibited improved humidity and thermal stability, while the device demonstrated enhanced long-term storage stability.
As shown in Fig. 11f, Li et al. demonstrated an in situ surface reconstruction using siloxane surfactants to form a protective layer that inhibits moisture infiltration.118 When 3-aminopropyltriethoxysilane (APTES) was surface-treated on the perovskite surface, siloxanes reacted with moisture in the air to form a hydrophobic Si–O–Si cross-linked network layer, which impeded the infiltration of moisture and oxygen. The water contact angle of the APTES-treated CsPbI3 film significantly increased from 52.25° to 76.5°, demonstrating enhanced hydrophobicity. Furthermore, the –NH2 groups in APTES were converted to –NH3+, interacting with I− and anchoring to non-coordinated Pb2+, which demonstrated defect passivation and the suppression of ion migration. The APTES-treated inorganic perovskite film maintained the black phase for 216 h when exposed to air at 20–25 °C and 30–40% relative humidity. Additionally, the unencapsulated device retained over 80% of its initial efficiency for 500 h at around 30% relative humidity, while the encapsulated device maintained 95% of its initial efficiency for 500 h at 85% relative humidity, demonstrating improved stability.
Hea et al. synthesized OMXene plates by oxidizing Ti3C2Tx MXene and applied them as a surface post-treatment on CsPbI3 to form a physical protective layer that prevents moisture ingress.120 By adjusting the oxidation time of MXene, OMXene with a TiO2 electron-transporting shell on its surface was synthesized. As a result, the electron selectivity was enhanced and the formation of a strong electric field at the OMXene interface improved charge extraction. The OMXene-treated CsPbI3 mini-module encapsulation demonstrated robust stability maintaining 90% of its initial efficiency at 85 °C and 85% relative humidity environments during 1 sun light soaking for 1000 h.
At the CTL/perovskite interface, various factors contribute to the degradation of the perovskite film and device performance. These include defects such as oxygen vacancies in the CTL and uncoordinated Pb2+ in the perovskite interface, chemical reactions with the CTL and perovskite, nonuniform perovskite film growth caused by CTL surface roughness and aggregation, lattice mismatch, inadequate adhesion between the CTL and perovskite layer, and energy-level mismatch leading to charge accumulation. To address these issues occurring at the CTL/perovskite interfaece, it is necessary to optimize the interface through various strategies (Table 4).
No. | Strategy | Structure | Type | PCE | Stability | Ref |
---|---|---|---|---|---|---|
1 | ETL interface | FTO/TiO2/DCB(SA)/CsPbI3/spiro-OMeTAD/Ag | NIP | 21.86% | 91.1% for 300 h@1 sun illumination/non-encapsulated | 122 |
87.1% for 720 h@RH: 35 ± 5%/non-encapsulated | ||||||
2 | FTO/TiO2/ATFC/CsPbI3/spiro-OMeTAD/Au | NIP | 21.11% | 99.37% for 350 h@MPP/1 sun illumination/RH: ∼30%/N2/encapsulated | 123 | |
95.48% for 120 h@T: 65 °C/N2/non-encapsulated | ||||||
92.47% for 800 h@RH: 30–35%/T: 25 °C/non-encapsulated | ||||||
3 | FTO/TiO2/MOCs/CsPbI3/spiro-OMeTAD/Au | NIP | 20.67% | 94% for 800 h@RH: 20–30%/non-encapsulated | 124 | |
4 | FTO/TiO2/18C6/CsPbI3/spiro-OMeTAD/Au | NIP | 22.14% | 95% for 1500 h@MPP/1 sun illumination/T: 50–55 °C/N2/non-encapsulated 95% for 1500 h@RH: <10%/T: 25 °C/non-encapsulated | 125 | |
5 | FTO/TiO2/SPA/CsPbI3/OAI/spiro-OMeTAD/Au | NIP | 20.98% | 80% for 400 h@1 sun illumination/N2/non-encapsulated | 126 | |
>90% for 200 h@T: 80 °C/N2/non-encapsulated | ||||||
91% for 1500 h@RH: 20–30%/T: 20–30 °C/non-encapsulated | ||||||
6 | FTO/NiOx/CsPb0.6Sn0.4I3:4AMPI2/SnOx/PCBM/BCP/Ag | PIN | 16.79% | >90% for 958 h@MPP/1 sun illumination/N2/non-encapsulated | 127 | |
7 | HTL interface | FTO/NiOx/Br-2PACz/CsPbI3/PCBM/BCP/Ag | PIN | 19.34% | 89% for 284 h@MPP/1 sun illumination/N2/non-encapsulated | 128 |
73% for 48 h@T: 60 °C/non-encapsulated | ||||||
80% for 960 h@RH: 30–35%/non-encapsulated | ||||||
8 | ITO/NiOx/2PACz/TPAI/CsPbI3/PCBM/BCP/Ag | PIN | 21.60% | 96.71% for 1400 h@MPP/1 sun illumination/RH: ∼30%/encapsulated | 129 | |
95.88% for 2000 h@RH: 30–35%/non-encapsulated | ||||||
9 | ITO/NiOx/MeO-2PACz/NbCl5/CsPbI3/PCBM/BCP/Ag | NIP | 21.24% | 92.27% for 1000 h@MPP/1 sun illumination/RH: 30–35%/encapsulated | 130 | |
97.61% for 1000 h@RH: 35%/T: 25 °C/non-encapsulated | ||||||
10 | FTO/TiO2/CsPbI3−xBrx/spiro-OMeTAD:TA/Au | NIP | 21.80% | 83% for 250 h@T: 65 °C/N2/non-encapsulated | 131 | |
90% for 800 h@RH: 25–35%/non-encapsulated | ||||||
11 | FTO/TIO2/CsPbI3/spiro-OMeTAD:BPFPDS/Au | NIP | 21.95% | 98% for 1200 h@MPP/1 sun illumination/RH: 30%/T: 85 °C/encapsulated | 132 | |
89% for 300 h@T: 65 °C/non-encapsulated | ||||||
96% for 3000 h@RH: 15–20%/non-encapsulated | ||||||
12 | Low-dimension heterostructure | FTO/TiO2/Al2O3/CsPbI3/Cs2PbI2Cl2/CuSCN/Cr/Au | NIP | 17.4% | 80% for 2100 h@MPP/120 mW cm−2/RH: 65 ± 26%/T: 110 °C/encapsulated | 133 |
13 | ITO/PTAA/CsPbI3/Cs4PbBr6/CPTA/BCP/Ag | PIN | 21.03 | 92.48% for 1000 h@MPP/1 sun illumination/RH: 85%/T: 85 °C/encapsulated | 134 | |
76.33% for 1000 h@MPP/1 sun illumination/non-encapsulated | ||||||
66.03% for 1000 h@MPP/1 sun illumination/RH: 85%/non-encapsulated | ||||||
14 | FTO/c-TiO2/CsPbI2.75Br0.25/NPA 2D-RP/P3HT/Ag | NIP | 19.77 | 93.2% for 1200 h@MPP/1 sun illumination/RH: 25%/T: 25 °C/encapsulated | 135 | |
87.8% for 1200 h@MPP/1 sun illumination/N2/T: 20 °C/encapsulated | ||||||
88.9% for 1200 h@MPP/1 sun illumination/T: 80 °C/N2/encapsulated | ||||||
15 | Overcoming the hygroscopic additive (HTM free/carbon electrode) | FTO/TiO2/m-Al2O3/CsPbI3/CF3-PAI/carbon | NIP | 18.33 | 87.7% for 200 h@MPP/1 sun illumination/N2/non-encapsulated | 136 |
90% for 1000 h@RH: <20%, T: 10–30 °C/non-encapsulated | ||||||
16 | FTO/TiO2/CsPbI3/TChI/carbon | NIP | 19.08 | 91.5% for 300 h@MPP/1 sun illumination/N2/non-encapsulated | 137 | |
93.2% for 100 h@MPP/1 sun illumination/RH: 65 ± 5%/encapsulated | ||||||
91.5% for 300 h@T: 65 °C/N2/non-encapsulated |
As shown in Fig. 12a, Qiu et al. adapted dipolar chemical bridge (DCB) materials at the ETL/inorganic perovskite interface.122 The DCB material, 3-amino-1-propane-sulfonic acid (SA), generated a strong electric field at the interface, optimizing the interfacial energy-level alignment and facilitating charge extraction. SA passivated defects through strong interactions with uncoordinated Ti4+ on the TiO2 surface, I− ions and uncoordinated Pb2+ in the buried perovskite interface. Furthermore, by enhancing the contact with the perovskite, it effectively reduced non-radiative recombination. Furthermore, the interface treatment with SA improved the crystallinity and relieved tensile strain of the inorganic perovskite, thereby enhancing the durability of the inorganic perovskite film. The SA-treated unencapsulated device maintained 87.1% of its initial efficiency after 720 h of storage at 35 ± 5% RH. Additionally, it retained 91.1% of its initial efficiency after 300 h of operation under continuous 100 mW cm−2 illumination.
![]() | ||
Fig. 12 The review of stability enhancement via interface optimization. (a) ETL/inorganic perovskite interface optimization via TiO2 surface treatment with 3-amino-1-propane-sulfonic acid (SA) (reproduced with permission from Angew. Chem. Int. Ed. 2024, 63, e202401751. Copyright 2024 Wiley-VCH GmbH),112 (b) suppression of Sn2+ oxidation and in situ formation of a SnO2 ETL via the surface Sn(IV) hydrolysis (SSH) technique in Pb–Sn-based inorganic perovskite solar cells (reproduced with permission from ACS Energy Lett., 2023, 8(2), 1035–1041. Copyright 2023 American Chemical Society),127 (c) modification of the SAM surface with 4-(aminomethyl)-N,N-diphenylaniline iodide (TPAI) to suppress SAM aggregation and improve molecular ordering (reproduced with permission from Angew. Chem., 2025, e202502221. Copyright 2025 Wiley-VCH GmbH),129 (d) suppression of Li+ ion migration and passivation of perovskite interface defects by incorporating the biocompatible molecule tryptamine (TA) into the HTL (reproduced with permission from Adv. Mater., 2024, 36, 2306982. Copyright 2023 Wiley-VCH GmbH),131 (e) surface defect passivation, lattice strain buffering, and environmental barrier effect achieved by forming a 3D/0D heterostructure through deposition of 0D Cs4PbBr6 nanocrystals on CsPbI3 (reproduced with permission from Adv. Mater., 2024, 36, 2408387. Copyright 2024 Wiley-VCH GmbH),134 and (f) optimization of the CsPbI3/carbon interface in HTM-free devices via CF3-PAI interface treatment (reproduced with permission from Small, 2024, 20, 2402061. Copyright 2024 Wiley-VCH GmbH).136 |
As shown in Fig. 12b, Hu et al. applied the surface Sn(IV) hydrolysis (SSH) technique in Pb–Sn based inorganic perovskite solar cells to suppress Sn2+ oxidation and form an in situ SnO2 ETL, thereby enhancing charge transport and extraction.127 As a first step, SnF2 was coated onto the perovskite surface to reduce Sn4+ to Sn2+, while also inducing Sn2+ to fill defects on the surface. Subsequently, a second surface treatment was performed using an IPA solution containing a trace amount of H2O, forming an ultrathin SnO2 film. The in situ SnO2 film formed through secondary treatment suppressed the oxidation of Sn2+ in Pb–Sn based inorganic perovskites and improved energy-level alignment, resulting in enhanced device stability and charge extraction capability. The unencapsulated device employing the SSH technique demonstrated long-term stability, maintaining over 90% of its initial efficiency for 958 h under 1 sun continuous illumination in a nitrogen-filled glovebox.
Recently, to address these issues, strategies have been reported that either replace conventional HTL materials with self-assembly materials (SAMs)140 or enhance stability by suppressing chemical reactions through SAM surface treatment.128 However, despite these advantages, the uneven distribution of the SAM layer due to agglomeration causes defects to form at the HTL/perovskite interface, ultimately degrading stability.129,130
As shown in Fig. 12c, Liu et al. demonstrated that the modification of 4-(aminomethyl)-N,N-diphenylaniline iodide (TPAI) to the SAM surface enhances stability.129 TPAI bonded to the SAM through π–π interactions, suppressing SAM aggregation and improving SAM order. In addition, TPAI passivated defects at the CsPbI3 surface by interacting with Pb2+ through its –NH3 group and optimized the energy-level alignment at the HTL/perovskite interface, enhancing charge extraction and suppressing charge recombination. The TPAI-treated perovskite solar cell maintained 96.71% of its initial efficiency after 1400 h of maximum power point (MPP) tracking and retained 95.88% of its initial efficiency even after 2000 h of storage in air, demonstrating improved stability.
Another approach is to modify the HTM itself. In the case of n–i–p structured devices, the commonly used HTL, spiro-OMeTAD, contains additives such as LiTFSi to enhance its conductivity. The LiTFSi doped spiro-OMeTAD is highly sensitive to humidity and temperature variations, leading to Li+ ion migration. Additionally, due to the high moisture affinity of Li+ ions, it can absorb humidity and react with the perovskite layer, causing its degradation.131,132
As shown in Fig. 12d, Liu et al. proposed a method to suppress Li+ ion migration and passivate perovskite interface defects by incorporating the biocompatible molecule tryptamine(TA) into the HTL.131 The –NH2 group of TA formed strong bonds with undercoordinated Pb2+ on the perovskite surface, reducing the interface defect density. Furthermore, TA formed TA:Li+ complexes, reducing moisture absorption and suppressing Li+ ion migration, thereby enhancing the stability of the perovskite layer. The device incorporating the TA-doped HTM maintained 90% of its initial efficiency after 800 h of storage at 25–35% RH. Additionally, in the thermal stability test, it retained 83% of its initial efficiency after 250 h at 65 °C.
Zhao et al. formed a 2D Cs2PbI2Cl2 capping layer on the CsPbI3 surface by applying CsCI post-treatment.133 Through CsCl surface treatment, Cl− ions passivated I vacancy defects, while the 2D Cs2PbI2Cl2 capping layer increased the activation energy to 0.536 eV, thereby suppressing iodide ion migration from the perovskite to the HTL under thermal and light stress, ultimately enhancing stability. The encapsulated 3D CsPbI3/2D Cs2PbI2Cl2 device demonstrated a remarkable T80 lifetime of 2100 h using the ISOS-L-3 protocol measured at the MPP under simultaneous light exposure (120 mW cm−2) and in a heating environment at 110 °C.
As shown in Fig. 12e, Heo et al. synthesized 0D Cs4PbBr6 nanocrystals on the CsPbI3 surface and subsequently applied them via the spray coating process, forming a 3D/0D heterostructure.134 0D Cs4PbBr6 consists of isolated PbBr64− octahedra surrounded by Cs+ ions, effectively passivating Pb2+ defects on the perovskite surface. Also, the Cs4PbBr6 lattice spacing, similar to that of 3D perovskites, acts as a lattice strain holder, improving both phase and thermal stability. Furthermore, it has been shown to function as a barrier against environmental stress. The encapsulated device with 0D Cs4PbBr6 treatment retained 92.48% of its initial efficiency under simultaneous 1 sun illumination and damp heat (85 °C/85% relative humidity) for 1000 h.
In contrast, the chemical inertness of carbon electrodes fundamentally prevents electrode corrosion while the HTL-free structure enhances moisture resistance, thereby improving stability. However, despite these advantages, the abundant surface defects of inorganic perovskites and the low hole selectivity of carbon electrodes lead to decreased durability and performance degradation.136,137,143,144
As shown in Fig. 12f, Zhang et al. demonstrated a method utilizing a dipole electric field at the CsPbI3/carbon interface to enhance device performance and stability.136 The –NH3 group of CF3-PAI bonded to the perovskite surface, while the –CF3 group interacted with the carbon electrode. This alignment generated a well-ordered dipole electric field within the perovskite/carbon electrode interface, enhancing hole selectivity and interfacial charge separation. In addition, CF3-PAI passivated defects by interacting with uncoordinated Pb2+, which suppressed non-radiative recombination and contributed to the improvement of phase stability. The unencapsulated CF3-PAI-treated HTM-free device maintained 90% of its initial efficiency over 1000 h under storage conditions with relative humidity below 20%. Moreover, under continuous illumination at 100 mW cm−2 for 200 h, it retained 87.7% of its performance, demonstrating enhanced operational stability.
Zhang et al.144 and Lin et al.137 treated the inorganic perovskite surface with choline iodide (ChI) and thiocholine iodide (TChI), respectively, forming a low-dimensional perovskite layer, which passivated the inorganic perovskite surface and enhanced charge extraction capability by improving energy-level alignment. In particular, Zhang et al.144 enhanced stability by applying ChI in two steps, first at the wet film intermediate phase and then as a second treatment on the dried film, effectively passivating both bulk and surface defects in the inorganic perovskite.
Another strategy for overcoming the low stability induced by hygroscopic additives to the HTL is to use self-standing inorganic HTLs such as CuSCN and phthalocyanine (CuPc) that do not require hygroscopic additives.145 Moreover, inorganic HTLs are generally more robust both mechanically and chemically, compared to soft organic HTLs such as spiro-OMeTAD, thus providing a significant advantage in long-term thermal and environmental stability.146 Similarly, inorganic ETLs such as SnO2 can also provide robust stability when processed on top of the perovskite layer for inverted perovskite solar cells.147 However, the number of reports on using inorganic CTLs for all-inorganic PSCs with high stability is extremely limited, largely due to the difficulty in processing inorganic CTLs on top of the inorganic perovskite surface, and many reports on “all-inorganic PSCs” unfortunately still rely on organic top-CTLs such as spiro-OMeTAD and PCBM. It is expected that there remains a significant amount of research to be explored in the development of “true” all-inorganic PSCs with inorganic CTLs for enhanced thermal and moisture stability.
![]() | ||
Fig. 13 The Shockley–Queisser (SQ) limit graph showing the record efficiencies of Pb-, Sn-, and Pb–Sn-based inorganic perovskites and hybrid perovskites according to bandgap energy. |
In particular, Sn- and Pb–Sn-based perovskites possess narrow bandgaps in the range of approximately 1.3 to 1.5 eV, which are close to the ideal bandgap of 1.34 eV for single-junction solar cells under AM1.5G illumination, and are well-suited as bottom cells in tandem architectures. Therefore, research on Sn- and Pb–Sn-based inorganic perovskites is essential not only for achieving optimal efficiency in single-junction solar cells but also for developing all-perovskite tandem solar cells that can overcome the efficiency limitations of single-junction devices while ensuring high thermal and operational stability.
No. | Strategy | Structure | Type | Jsc | Voc | FF | PCE | Ref. |
---|---|---|---|---|---|---|---|---|
1 | B-site engineering | ITO/NiOx/CsPb(0.6−x)Sn(0.4−x)MnxI(3−y)Bry:BHA/PCBM/ZrAcac/Ag | PIN | 26.9 | 0.83 | 76.7 | 17.12 | 149 |
2 | FTO/c-TiO2/CsPbI3−xBrx:CdI2/spiro-OMeTAD/Au | NIP | 20.64 | 1.21 | 83.2 | 20.8 | 150 | |
3 | FTO/c-TiO2/mp-TiO2/CsPb0.9Sn0.1IBr2/carbon | NIP | 14.3 | 1.26 | 63 | 11.33 | 151 | |
4 | FTO/SnO2/C60/CsPb0.75Sn0.25IBr2/spiro-OMeTAD/Au | NIP | 12.57 | 1.21 | 75.8 | 11.53 | 152 | |
5 | FTO/c-TiO2/mp-TiO2/Al2O3/CsPb0.5Sn0.5I2Br:TP/NiOx/carbon | NIP | 20.1 | 0.62 | 65 | 8.1 | 153 | |
6 | ITO/PEDOT:PSS/CsPb0.7Sn0.3I3/PCBM/BCP/Ag | PIN | 20.96 | 0.64 | 70.12 | 9.41 | 154 | |
7 | ITO/PEDOT:PSS/CsPb0.4Sn0.6I2Br/PCBM/CeOx/Ag | PIN | 19.02 | 0.86 | 75.1 | 12.34 | 155 | |
8 | ITO/NiOx/CsSnI2.6Br0.4:DMKO/PCBM/ZrAcac/Ag | PIN | 20.7 | 0.75 | 72.1 | 11.2 | 156 | |
9 | ITO/PEDOT:PSS/CsSn1−xZnxBr3/C60/BCP/Ag | PIN | 13.99 | 0.35 | 54 | 2.59 | 157 | |
10 | ITO/PEDOT:PSS/CsPb0.75Sn0.25I3/PCBM/bis-C60/Ag | PIN | 15.4 | 0.67 | 56 | 5.78 | 158 | |
11 | Phase engineering | FTO/TiO2/SnO2/β-CsPbI3:Zn(C6F5)2/γ-CsPbI3:GAI/P3HT + SMeTAPyr/Au | NIP | 21.72 | 1.22 | 81.5 | 21.59 | 159 |
12 | X-site engineering | FTO/c-TiO2/CsPbI3:HBr/spiro-OMeTAD/MoO3/IZO/Ag | NIP | 20.34 | 1.22 | 80.36 | 19.89 | 160 |
13 | ITO/PTAA/CsSnI2Br/ICBA/BCP/Ag | PIN | 22.2 | 0.75 | 71.29 | 11.87 | 37 | |
14 | ITO/PEDOT:PSS/CsSnI3:CsFa/ICBA/BCP/Ag | PIN | 24.94 | 0.75 | 74 | 13.68 | 161 | |
15 | FTO/TiO2/CsPbI3:CsFo/spiro-OMeTAD/Au | NIP | 20.57 | 1.25 | 82.5 | 21.23 | 162 | |
16 | FTO/TiO2/CsPbI3:DMAAc/spiro-OMeTAD/Au | NIP | 20.8 | 1.25 | 81.6 | 21.14 | 163 | |
17 | FTO/TiO2/CsPbI3:DMAAc/spiro-OMeTAD/Au | NIP | 20.81 | 1.23 | 82.2 | 21.04 | 164 | |
18 | ITO/SnO2/CsPbI2.75Br0.25:DMACl/CsI/spiro-OMeTAD/Au | NIP | 18.86 | 1.356 | 81.15 | 20.75 | 165 | |
19 | FTO/TiO2/CsPbI3:DMAFa/spiro-OMeTAD/Au | NIP | 20.49 | 1.21 | 82.02 | 20.4 | 166 | |
20 | ITO/P3CT-N/CsPbI3:PbAc2/PCBM/C60/BCP/Ag | PIN | 20.52 | 1.215 | 80.89 | 20.17 | 167 | |
21 | FTO/TiO2/CsPbI2Br/CsPbI3−xBrx/PTAA/Au | NIP | 18.08 | 1.15 | 80.85 | 16.81 | 168 | |
22 | FTO/PEDOT:PSS/PTAA/CsPb0.4Sn0.6I2.4Br0.6/C60/BCP/Ag | PIN | 22.96 | 0.793 | 78.77 | 13.98 | 169 | |
23 | ITO/PEDOT:PSS/CsPb0.5Sn0.5I2.7Br0.3/PCBM/BCP/Ag | PIN | 22.11 | 0.86 | 72.7 | 13.82 | 170 |
Cai et al. introduced a small amount of HBr acid into the CsPbI3 perovskite precursor solution as a source of Br− anions and phase regulator for preparing efficient semitransparent inorganic perovskite solar cells.160 The Br− additive induced a slight blue shift in the absorption edge of the perovskite film, confirming successful Br− halide incorporation. The HBr additive induced partial formation of α-CsPbI3 upon annealing at 100 °C, in contrast to γ-CsPbI3 phase formation at 40 °C in the control film, thus imparting mixed α and γ phases, as shown in Fig. 14a. The co-existing mixed CsPbI3 phases mitigate spontaneous strain and substrate-constrained strain from mismatching thermal expansion coefficients, thus securing photoactive CsPbI3 perovskite films. Moreover, the HBr treated CsPbI3 film is a uniform and dense film and demonstrated lower defect density, suppressing the detrimental non-radiative recombination. Accordingly, the CsPbI3 perovskite solar cell with a transparent IZO electrode demonstrated a remarkable 19.89% PCE compared to the 17.80% PCE of the control sample.
![]() | ||
Fig. 14 The review of efficiency enhancement via material composition and phase engineering with (a) HBr in CsPbI3 (reproduced with permission from Small, 21(14), 2500710. Copyright 2025 Wiley-VCH GmbH),160 (b) Br− ion in CsSnI3 (reproduced with permission from ACS Energy Lett., 8(12), 5061–5069. Copyright 2023 American Chemical Society),37 and (c) formate pseudohalide anion in CsSnI3 (reproduced with permission from ACS Energy Lett., 9(12), 5870–5878. Copyright 2024 American Chemical Society).161 |
Hong et al. modulated the I to Br halide ratio in CsSnI3−xBrx to optimize the performance and the inherent stability trade-off relation from composition engineering to fabricate efficient inorganic lead-free perovskite solar cells.37 The authors systematically investigated the optoelectronic characteristics of the CsSnI3−xBrx film with varying I and Br halide composition through computational and experimental methods. While the I-rich composition benefitted the wider range of light absorption from lower bandgaps, the Br-rich composition favored enhanced γ-phase stability, as shown in Fig. 14b. The optimized CsSnI2Br perovskite exhibited robust lattice stability and demonstrated excellent performance, reaching 11.97% PCE through the precursor purification process and compositional engineering.
Wang et al. reported an anion alloying strategy in CsSnI3 perovskite solar cells with pseudohalides by incorporating a cesium formate (CsFa) additive in the precursor solution.161 The introduced Fa− effectively suppressed the inorganic CsSnI3 perovskite film through regulating the crystallization process. The Fa−-alloyed perovskite film exhibited slower crystallization by forming a low-dimensional intermediate phase and, therefore, resulted in a larger grain size and a narrower surface potential difference compared to the control CsSnI3 film, as shown in Fig. 14c. Moreover, the Fa− anions effectively passivate defects by coordinating with undercoordinated Sn2+ that fills the X-site vacancies and by suppressing Sn4+ defective oxidation states. As a result, the CsSnI3 film demonstrated a remarkable certified 13.68% PCE through the pseudohalide anion alloying strategy.
Wang et al. investigated the ternary B-site co-doping strategy to enhance the performance of a CsPb1−x−ySnxByI3−zBrz perovskite composition system.149 The B-site co-doping strategy was performed by introducing Sr2+, Ge2+, Zn2+, and Mn2+ to enhance phase stability and promote light response from adjusting the optical bandgap. Among the various candidates, 4% Mn2+ incorporation into the B-site alloy system benefitted from the small ionic radius, increasing the Goldschmidt tolerance factor up to 0.82, favoring the photoactive black perovskite phase. Moreover, the Pb–Sn–Mn co-doped perovskite system exhibited a reduced bandgap to 1.36 eV closer to the ideal bandgap, contributing to enhanced performance reaching 14.34% PCE and phase stability, as shown in Fig. 15b. Upon further introduction of benzhydroxamic acid as an antioxidant and a passivation agent, the inorganic Pb–Sn–Mn based perovskite solar cell demonstrated excellent 17.12% PCE.
![]() | ||
Fig. 15 The review of efficiency enhancement via material composition and phase engineering with (a) Zn(C6F5)2 in CsPbI3 perovskite (reproduced with permission from Nat. Energy, 8, 989–1001. Copyright 2023 Springer Nature),159 (b) Mn2+ in CsPb1−x−ySnxByI3−zBrz perovskite (reproduced with permission from Adv. Mater., 36(14), 2309193. Copyright 2023 Wiley-VCH GmbH),149 and (c) CdI2 in CsPbI3−xBrx (reproduced with permission from Adv. Sci., 9(36), 2204486. Copyright 2022 Wiley-VCH GmbH).150 |
Liu et al. introduced CdI2 in a CsPbI3−xBrx perovskite solar cell for B-site compositional engineering and for surface defect passivation.150 The Cd2+ ions possess a smaller ionic radius than Pb2+ and, therefore, are partially substituted within the CsPbI3−xBrx lattice up to 2 mol%, which reduces detrimental lattice distortion of the inorganic [PbX6]4− framework. Interestingly, the excess CdI2 additive mainly remained at the surface grain boundary in the form of Cs2PbI4−xBrx, as shown in Fig. 15c, and contributed to the enhanced performance by reducing non-radiative recombination. Furthermore, the CdI2 additive slows down the nucleation rate from increased precursor colloidal distribution and leads to the formation of larger crystal grain size with surface defect states. Accordingly, the CsPbI3−xBrx perovskite solar cell with 8 mol% CdI2 demonstrated 20.8% PCE from lattice engineering and defect control.
Hong et al. proposed a phase heterojunction (PHJ) with β-CsPbI3 and γ-CsPbI3 to prepare highly efficient CsPbI3 perovskite solar cells by exploiting the CsPbI3 polymorphs induced by compositional engineering.159 The PHJ CsPbI3 film was prepared by stacking the front dimethylammonium iodide-assisted β-CsPbI3 perovskite along with a bis(pentafluorophenyl)zinc (Zn(C6F5)2) additive from the solution process and the rear guanidinium iodide-assisted γ-CsPbI3 perovskite from the vapor phase process. The CsPbI3 polymorph stack provides a stable coherent phase-heterojunction without gap states, exhibiting favorable interfacial adhesion energy reaching −0.213 eV Å−2 and facilitating charge extraction with favorable band alignment as shown in Fig. 15a. Accordingly, the PHJ perovskite solar cell demonstrated a remarkable maximum 21.59% PCE and 18.43% PCE from a 19.17 cm2 active area module.
No. | Strategy | Structure | Type | JSC [mA cm−2] | VOC [V] | FF [%] | PCE [%] | Ref. |
---|---|---|---|---|---|---|---|---|
1 | Surface defect passivation | FTO/TiO2/CsPbI3−xBrx/HA/spiro-OMeTAD/Au | NIP | 20.55 | 1.233 | 81.9 | 20.8 | 171 |
2 | FTO/TiO2/CsPbI3−xBrx/TFA/spiro-OMeTAD/Au | NIP | 20.74 | 1.239 | 83.07 | 21.35 | 108 | |
3 | FTO/TiO2/CsPbI3−xBrx/2,6-DAPy/spiro-OMeTAD/Au | NIP | 20.78 | 1.252 | 83.71 | 21.8 | 106 | |
4 | FTO/TiO2/CsPbI3−xBrx/BMBC/spiro-OMeTAD/Au | NIP | 20.83 | 1.249 | 83.57 | 21.75 | 107 | |
5 | FTO/TiO2/CsPbI3/BTABr/spiro-OMeTAD/Au | NIP | 20.81 | 1.2 | 85.4 | 21.31 | 117 | |
6 | Buried interface passivation | FTO/TiO2/SPA/CsPbI3/OAI/spiro-OMeTAD/Au | NIP | 20.51 | 1.229 | 83.8 | 20.98 | 126 |
7 | FTO/TiO2/ATFC/CsPbI3/spiro-OMeTAD/Au | NIP | 20.6 | 1.24 | 82.82 | 21.11 | 123 | |
8 | ITO/NiOx/4-ABSA/CsPb0.6Sn0.4I3−xBrx/PCBM/ZrAcac/Ag | PIN | 26.5 | 0.83 | 79.1 | 17.4 | 172 | |
9 | Bulk passivation | FTO/TiO2/CsPbI3:CsFo/spiro-OMeTAD/Au | NIP | 20.57 | 1.25 | 82.5 | 21.23 | 162 |
10 | FTO/TiO2/CsPbI3:EMIMHSO4/spiro-OMeTAD/Au | NIP | 20.6 | 1.17 | 83 | 20.01 | 97 | |
11 | FTO/TiO2/CsPbI3:DED/spiro-OMeTAD/Au | NIP | 20.6 | 1.244 | 82.52 | 21.15 | 93 | |
12 | ITO/SnO2/CsPbI2.85Br0.15:PC/ODADI/spiro-OMeTAD/Au | NIP | 20.4 | 1.342 | 80.59 | 22.07 | 94 | |
13 | Crystal growth control | FTO/TiO2/CsPbI3:RBITC/spiro-OMeTAD/Au | NIP | 20.65 | 1.24 | 82.44 | 20.95 | 95 |
14 | FTO/TiO2/CsPbI3:DMAAc/spiro-OMeTAD/Au | NIP | 20.8 | 1.25 | 81.6 | 21.14 | 163 | |
15 | ITO/P3CT-N/CsPbI3:PbAc2/PCBM/C60/BCP/Ag | PIN | 20.52 | 1.215 | 80.89 | 20.17 | 167 | |
16 | Lattice stress release | FTO/TiO2/CsPbI3:GDY/spiro-OMeTAD/Au | NIP | 20.49 | 1.191 | 83.96 | 20.49 | 87 |
17 | ITO/NiOx/CsPbI2.85Br0.15:5-MVA/PCBM/BCP/Ag | PIN | 20.08 | 1.23 | 84.31 | 20.82 | 96 | |
18 | ITO/NiOx/MeO-2PACz/NbCl5/CsPbI3/PCBM/BCP/Ag | PIN | 20.5 | 1.25 | 83.1 | 21.24 | 130 | |
19 | Crystal growth control + bulk passivation | FTO/TiO2/CsPbI3:DMAFa/spiro-OMeTAD/Au | NIP | 20.49 | 1.21 | 82.02 | 20.4 | 166 |
20 | Buried interface passivation + lattice stress release | FTO/TiO2/DCB/CsPbI3/spiro-OMeTAD/Ag | NIP | 20.71 | 1.26 | 83.8 | 21.86 | 122 |
21 | Sn-defect regulation (crystal growth control) | ITO/PEDOT:PSS/CsPb0.55Sn0.45I2Br:CsCl/PbSO4/PCBM/BCP/Ag | PIN | 20.57 | 0.7 | 71.82 | 10.39 | 173 |
22 | FTO/PEDOT:PSS/PTAA/CsPb0.4Sn0.6I2.4Br0.6:MAI/C60/BCP/Ag | PIN | 22.96 | 0.793 | 78.77 | 13.98 | 169 | |
23 | ITO/PEDOT:PSS/CsSnI3:CsFa/ICBA/BCP/Ag | PIN | 24.94 | 0.75 | 74 | 13.68 | 161 | |
24 | Sn-defect regulation (antioxidants) | ITO/SnO2/CsPb0.7Sn0.3IBr2:ZnC2O4/spiro-OMeTAD/Au | NIP | 15.5 | 1.18 | 76.7 | 14.1 | 75 |
25 | ITO/NiOx/CsPb0.6Sn0.4I2Br:DCD/ZnO/PCBM/Ag | PIN | 22.67 | 0.87 | 71.73 | 14.17 | 99 | |
26 | ITO/NiOx/CsSnI3:CBZ/ZnO/PCBM/Ag | PIN | 21.38 | 0.73 | 71.99 | 11.21 | 81 | |
27 | ITO/NiOx/CsSnI2.6Br0.4:DMKO/PCBM/ZrAcac/Ag | PIN | 20.7 | 0.75 | 72.1 | 11.2 | 156 | |
28 | ITO/PEDOT:PSS/CsSnI3:PTM/ICBA/BCP/Ag | PIN | 21.81 | 0.64 | 72.1 | 10.1 | 100 | |
29 | FTO/c-TiO2/mp-TiO2/CsSnI3:EMIMAc/Al2O3/NiOx/carbon | NIP | 22.1 | 0.56 | 69 | 8.54 | 92 | |
30 | ITO/PEDOT:PSS/CsSnI3:TSC/C60/BCP/Cu | PIN | 19.7 | 0.63 | 66.1 | 8.2 | 174 | |
31 | FTO/c-TiO2/mp-TiO2/CsSnI3:CPT/Al2O3/NiOx/carbon | NIP | 23.4 | 0.52 | 66 | 8.03 | 102 |
Post-treatment of the perovskite film surface is one of the most widely used methods of defect passivation, especially since the perovskite surface is most vulnerable to defect formation during thin film fabrication.106–108,117,171,175,176 Gu et al. identified iodine vacancies as one of the most common types of surface defects for all-inorganic perovskite thin films and demonstrated the use of histamine (HA) to passivate the under-coordinated Pb2+ defects on the perovskite surface.171 HA possesses an amine group and an imidazole ring, which could coordinate with the Pb2+ defects while simultaneously forming a H-bond with neighboring halides, effectively resulting in increased VOC and FF from 1.195 V and 80.9% to 1.233 V and 81.9% and a PCE increase from 19.5% to 20.8%. Similarly, many surface passivation studies thereafter have focused on using molecules that can effectively interact with under-coordinated Pb2+ defects, while adding extra functionalities to the passivator for added effects. Zhang et al. used a trifluoroacetamide (TFA) passivator that has two amine groups for stronger chelation with the Pb2+ defects.108 Moreover, the trifluoro group not only applies a strong dipole to the TFA passivator, but also gives hydrophobic passivation capabilities, greatly increasing the PCE to 21.35%. In 2023, Wang et al. additionally identified iodine interstitial defects as another source of solar cell performance inhibition and used 2,6-diaminopyridine (2,6-DAPy) to strongly passivate the surface defects as shown in Fig. 16a.106 The authors showed that the strong nucleophilicity of the pyridine ring is more effective in passivating the positively charged defects, effectively resulting in a large increase in VOC and FF from 1.189 V and 80.03% to 1.252 V and 83.71% and a significant PCE increase from 19.6% to 21.8%. Amine groups are not the only functional groups that can passivate perovskite surface defects, as shown by Zhang et al. in 2023.107 Here, the authors used a cysteine derivative, namely BMBC, that possesses multiple functional groups such as –NH, –S and –CO that can each interact with halide vacancy defects. Through the multiple passivation effect, the authors were able to confirm the effectiveness of this strategy through the enhanced VOC of 1.249 V and high FF of 83.57%, yielding a high PCE of 21.75%. On the other hand, Tan et al. demonstrated the passivation of the perovskite surface with extra halide anions, supplied from an organic halide salt, BTABr.117 The authors showed that the bromide ions could not only passivate the surface defects but also diffuse into the perovskite bulk film and passivate deeper defects, while most of the BTA+ ions remain at the surface. The effectiveness of this strategy can be seen by the high efficiency of 21.31%.
![]() | ||
Fig. 16 The review of efficiency enhancement via reducing trap-induced recombination losses. (a) Perovskite surface defect passivation with 2,6-diaminopyridine (2,6-DAPy) (reproduced with permission from Angewandte Chemie, 2023, 62, e202305815. Copyright 2023 Wiley-VCH GmbH).106 (b) Buried interface defect passivation with 3-sulphonatopropyl acrylate potassium (SPA) (reproduced with permission from Advanced Materials, 2023, 35, 2207172. Copyright 2022 Wiley-VCH GmbH).126 (c) Bulk defect passivation via additive engineering with 2,6-pyridinedicarboxamide (PC) (reproduced with permission from Advanced Energy Materials, 2024, 14, 2400151. Copyright 2024 Wiley-VCH GmbH).94 (d) Inhibiting defect formation by releasing lattice stress with a 5-maleimidovaleric acid (5-MVA) additive (reproduced with permission from the Journal of Energy Chemistry, 2025, 100, 87–93. Copyright 2024 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences).96 (e) Regulating crystal growth kinetics and transformation of the phase transition pathway with a cesium formate (CsFa) additive (reproduced with permission from ACS Energy Letters, 2024, 9, 5870–5878. Copyright 2024 American Chemical Society).161 (f) Preventing Sn2+ oxidation with a dicyandiamide (DCD) antioxidant additive (reproduced with permission from the Chemical Engineering Journal, 2023, 452, 139697. Copyright 2022 Elsevier B.V.).82 |
The bottom interface (or the buried interface) of perovskite thin films is also a well-known hotspot for detrimental defects, and their passivation must consider interactions of the passivating agent with both the perovskite defects and the CTL defects. Typically, a proper passivating agent would consist of functional groups such as amine groups or ammonium halides, for the passivation of the perovskite layer, and acids such as carboxylates and sulfonates for the passivation of the bottom layer, which is usually a metal oxide.122,126,170 In 2022, Xu et al. treated a TiO2 ETL with 3-sulphonatopropyl acrylate potassium (SPA) prior to coating the CsPbI3 inorganic perovskite layer.126 As shown in Fig. 16b, both the sulfonate and carboxylate groups can interact with either the oxygen vacancies of the TiO2 layer or the halide vacancies of the perovskite layer, while the potassium cation would bind with the halides of the perovskite layer in order to inhibit formation of additional halide vacancies. Moreover, the increased hydrophilicity of the TiO2 layer due to SPA treatment increased the perovskite crystallinity, leading to overall fewer defects, and hence an increased VOC of 1.221 V, increased FF of 82.8%, and increased PCE of 20.72%. Meanwhile, Xu et al. treated the TiO2/CsPbI3 interface with a 4-amino-2,3,5,6-tetrafluorobenzoate cesium (ATFC) passivator, where the TiO2 defects were passivated by the amine group of ATFC and the perovskite defects were passivated by the carboxylate group of ATFC.123 In addition, the adjacent fluorine atoms on the benzene ring could provide additional interaction with both the TiO2 defects and the perovskite defects, ensuring a stronger degree of passivation and a high PCE of 21.11% for the final device. Similarly, Qiu et al. used a dipolar passivator, named SA, for simultaneous passivation of TiO2 and perovskite layers.122 Compared to the commonly used carboxylate groups, the sulfonate group could form a tridentate chelation with the TiO2 defects for stronger passivation, while the hydrogen bonds between the amine group of SA and halides of perovskite passivated the CsPbI3 layer. Overall, a high PCE of 21.86% was obtained. Similar strategies can also be applied to inverted inorganic perovskite solar cell devices as well. Zhang et al. applied 4-aminobenzenesulfonic acid (4-ABSA) to passivate the bottom NiOx HTL surface and the buried perovskite layer interface simultaneously, achieving a PCE of 17.4% for a CsPb0.6Sn0.4I3−xBrx solar cell device.170
While the majority of defects are usually congregated at interfaces, the impact of perovskite bulk defects such as point defects and grain boundaries must also be addressed. However, despite their similarity in nature to interfacial defects, since the physical location of such defects makes it difficult to passivate them through pre- or post-treatment methods, researchers have turned to applying additives to the perovskite precursor solution itself so that harmful defects are passivated within the bulk film during and after the thin-film annealing process.93–95,97,103,162,166,177 For example, a common approach is to use a formate additive such as a dimethylamine formate (DMAFa) or CsFa additive, since formate anions can passivate the Pb2+ defects that occur due to halide vacancies within the perovskite bulk film.162,166 Similarly, in 2021, Du et al. introduced a low concentration of 1-ethyl-3-methylimidazolium hydrogen sulfate (EMIMHSAO4) ionic liquid additive to the perovskite precursor solution.97 The imidazolium (EMIM+) cations and HSO4− anions were shown to passivate electrophilic defects and under-coordinated Pb2+ defects, respectively, within the perovskite bulk film, resulting in a high performing device with 20.01% PCE. Meanwhile, Wang et al. used a single molecule DED or 2,2-dithienylketone (DTK) additive with the intent to passivate bulk Pb2+ defects with the collaborative chelating effect of the carbonyl and thienyl groups.93 Here, they found that the bifacial ‘dione’ structure of DED allowed for more efficient passivation of the perovskite bulk defects, where the optimized devices yielded a PCE of 21.15%. A notable aspect of additive engineering is that the additives more often than not fulfill multiple niches, as shown in the studies of Wang et al. using a 2,6-pyridinedicarboxamide (PC) additive.94 The extra stability to the perovskite lattice offered by multiple balancing interactions between the amide groups or the N-pyridine group of the PC additive and the Pb2+ defects, otherwise noted as a “hinge modulation” by the authors, ensured that the perovskite film with PC can withstand longer annealing times in air, thus resulting in a more crystalline perovskite film as depicted in Fig. 16c. As a result, not only did the optimized device display a high VOC of 1.342 V, but the resulting PCE is also among the recently highest reported for inorganic perovskite solar cells at 22.07%.
The formation process of solution processed perovskite films consists of nucleation and crystal growth, where it is generally accepted that fast homogeneous nucleation and slow crystal growth are beneficial for achieving high crystallinity.178,179 As such, Zhang et al. used a sacrificial dye Rhodamine B isothiocyanate (RBITC) to facilitate nucleation while deterring crystallization in order to increase the crystallinity of the perovskite film.95 The authors showed that photo-thermal decomposition of RBITC yields benzoic acid molecules that act as anti-solvent to promote initial nucleation, while the SCN− anions interact with the Pb sites to deter crystal growth, eventually resulting in a highly crystalline inorganic perovskite solar cell with a high PCE of 20.95%. In another study by Zai et al., DMAFa was added to the perovskite precursor, where the formate anion would interact with the exposed Pb2+ sites during the annealing process.166 This interaction provided sufficient time for grain growth as the HCOO− is gradually replaced by I−, and the final solar cell device with reduced defects exhibited increased VOC, an increased FF and increased PCE at 20.40% compared to the pristine device with 18.87% PCE. Meanwhile, another risk of inorganic perovskites is the formation of undesired photo-inactive intermediate phases such as Cs4PbI6 that can degrade the overall photovoltaic performance, and thus it is crucial to regulate the crystallization process to prevent this. A common strategy is to incorporate the dimethylammonium (DMA+) additive into the precursor solution, which has been reported to aid in the formation of a stable photo-active black phase of CsPbX3, while inhibiting the formation of Cs4PbX6.163,167 A study by Cui et al. well illustrates this strategy where a dimethylamine acetate (DMAAc) molten salt additive supplies the precursor solution with extra Ac− anions that can stabilize the DMAPbI3 lattice structure, which reacts with the CsI precursors to produce a stable CsPbI3 structure with volatile DMAI byproducts.163 In particular, the authors emphasize the importance of the DMAAc additive that inhibits Cs4PbI6 formation, whereas without DMAAc, the reaction of DMAPbI3 with CsI could produce Cs4PbI6. The optimized CsPbI3 solar cell device shows a high PCE of 21.14%. Meanwhile, Sun et al. employed dimethylammonium iodide (DMAI) and lead acetate (PbAc2) to substitute the original PbI2 precursor.167 The authors showed that the addition of DMAI increases the solubility of CsI within the precursor solution, after which the PbAc2 introduces Ac− ligands to the nuclei and slows down crystal growth as the anion is slowly exchanged with the iodide ion from DMAI. The formation of Cs4PbI6 is also shown to be inhibited in the optimized solar cell device, which showed a PCE of 20.17%.
Another inherent source of defects and trap-induced recombination loss in all-inorganic perovskite solar cells is the presence of lattice stress, which could potentially cause the formation of harmful defects within the perovskite layer. On one hand, the ion size mismatch of inorganic perovskites may lead to intrinsic tensile strain throughout the bulk perovskite film.87,96,180 In 2024, Zhang et al. introduced graphdiyne (GDY), an sp-hybridized carbon framework, to be incorporated within the CsPbI3 bulk film.87 The authors found that GDY not only delays the crystal growth process for the formation of a highly crystalline perovskite film, but it also helps release lattice stress within the bulk film during the annealing process, as noted in depth-dependent crystallographic analysis. The optimized device showed impressive VOC and FF values at 1.191 V and 83.96%, respectively, and a high PCE of 20.49%. Very recently, Sun et al. also noted the significance of releasing internal perovskite lattice stress and added a flexible molecule, 5-maleimidovaleric acid (5-MVA), as an additive within the inorganic perovskite precursor solution.96 It was shown that, since both functional groups at either end of the 5-MVA molecule are effective electron donors, they could strongly interact with Pb2+ defect sites within the CsPbI2.85Br0.15 perovskite lattice and act as physical buffers to release tensile strain of the perovskite film, as seen in Fig. 16d. Its immediate effectiveness was shown by the depth-dependent analysis and the greatly improved photovoltaic characteristics resulting in a high PCE of 20.82%. On the other hand, the lattice mismatch and discrepancies in thermal expansion coefficients between the perovskite lattice and substrates such as metal oxides can cause greater lattice distortion at the substrate/perovskite interface, thus also harming the interfacial contact.122,130 In the studies by Qiu et al., a flexible SA molecule based on a propyl chain that was treated on the TiO2 surface helped release the lattice stress at the TiO2/CsPbI3 interface.122 The inverted p–i–n structured devices adopt a similar strategy through the widely accepted SAM treatment on the NiOx substrate. Xu et al. realized that a poor SAM monolayer coverage on the NiOx substrate could cause unnecessary lattice distortions at the perovskite buried interface and treated the SAM layer with niobium pentachloride (NCL) to prevent SAM agglomeration and improve its morphology.130 As a result, the authors observed a decrease in interfacial lattice stress that also led to an increase in optimized solar cell PCE to 21.24%, which is among the highest PCEs reported to date for inverted inorganic perovskite solar cells.
The problem of rapid crystallization could be resolved by adapting similar strategies to regulating crystallization kinetics of Pb-based inorganic perovskite thin films discussed in the previous section.75,81,82,152,161,169,173 For example, Chen et al. first demonstrated the poor crystallinity of a pristine CsPb0.55Sn0.45I2Br perovskite film and then showcased a CsPb0.55Sn0.45I2Br–CsCl perovskite film with much improved crystallinity and large, micro-sized grains.173 It was shown that the CsCl additive provided small halogen ions that could simply infiltrate the bulk film at the grain boundaries during fabrication and strongly bond with the Sn2+, effectively reconfiguring the grain boundaries and inhibiting the formation of halide vacancies. With the addition of a hydrophobic PbSO4 passivation barrier at the perovskite surface, the optimized device displayed an impressive PCE of 10.39% with a highly improved VOC and FF of 0.70 V and 71.82%, respectively. As such, due to the lower solubility and faster ion release of the Sn-based precursors compared to the Pb-based precursors, it is important to alleviate this discrepancy in order to form a homogeneously distributed perovskite film. Shang et al. did a good job in demonstrating this idea as they partially substituted PbI2 with PbAc2 in the perovskite precursor solution.169 As shown in this work, the release rate of Sn2+ within the precursor solution is matched by the release of Pb2+ by PbAc2, while an additional MAI additive supplies extra iodide anions for the fabrication of a homogeneous phase-pure Pb–Sn perovskite solar cell with a high PCE of 13.98%. Recently, Yu et al. fabricated CsSnI3 perovskite solar cells with a CsFa additive in order to control the crystallization kinetics, yielding a high certified PCE of 13.68%, which is one of the highest reported efficiencies for CsSnI3 devices to date.161 Here, the authors proposed that the introduction of a pseudohalide anion that could strongly bond with the Sn2+ cations could dramatically slow down crystallization by transforming the phase transition path as shown in Fig. 16e.
Meanwhile, the oxidation of Sn2+ to Sn4+ provided new challenges to the Sn-based inorganic perovskite solar cell community. To address this challenge, many studies have actively sought antioxidants or reducing agents in order to inhibit or reduce the Sn2+ oxidation.81,92,99,100,102,156,173,174 For example, some cases report the effectiveness of Zn-based additives to prevent Sn2+ oxidation based on the more suitable reduction potential of Zn2+ compared to Sn2+.173 For example, zinc oxalate additives within the CsPb0.7Sn0.3PbIBr2 perovskite precursor solution would not only reduce any present Sn4+ to Sn2+, but also the oxalate anions could strongly interact with the metal ions to regulate crystal growth. The resulting Pb–Sn-based solar cell device showed a champion PCE of 14.1%. However, such strategies would also have to consider the possible incorporation of Zn2+ within the inorganic perovskite lattice, which could alter the intrinsic properties of the perovskite thin film. A study by Wen et al. in 2023 used a dicyandiamide (DCD) additive to prevent oxidation of Sn2+.99 The authors noted that the presence of –CN and –C
NH functional groups would provide a strong interaction with the Sn2+ ions within the perovskite film, actively preventing oxidization and reducing deep level defects, for a high PCE of 14.17% in the optimized solar cell device as shown in Fig. 16f. Similarly, many researchers have incorporated antioxidative additives such as carbazide (CBZ),81 dimethyl ketoxime (DMKO),156 phthalimide (PTM),100 1-ethyl-3-methylimidazolium acetate (EMIMAc),92 thiosemicarbazide (TSC),174 and 1-(4-carboxyphenyl)-2-thiourea (CPT)102 within CsSnI3-based precursor solutions and have shown promising results.
No. | Strategy | Structure | Type | JSC [mA cm−2] | VOC [V] | FF [%] | PCE [%] | Ref. |
---|---|---|---|---|---|---|---|---|
1 | Energy level alignment | FTO/NiOx/MeO-2PACz/CsPbI3−xBrx/MPTS/PCBM/BCP/Ag | PIN | 20.4 | 1.22 | 84.2 | 21 | 119 |
2 | FTO/TiO2/CsPbI3:MACl/PEACl/spiro-OMeTAD/Ag | NIP | 20.7 | 1.2 | 83.1 | 20.6 | 182 | |
3 | ITO/PTAA/CsPbI3/OMXene/CPTA/BCP/Ag | PIN | 19.9 | 1.21 | 82 | 19.7 | 120 | |
4 | FTO/P3CT/CsPbI3/PACl/PCBM/BCP/Ag | NIP | 21.36 | 1.13 | 84 | 20.17 | 183 | |
5 | ITO/P3CT-N/CsPbI3/DPhTA/PCBM/C60/TPBi/Ag | PIN | 20.84 | 1.18 | 81.9 | 20.18 | 181 | |
6 | Dipole interlayer | FTO/TiO2/DCB/CsPbI3/spiro-OMeTAD/Ag | NIP | 20.7 | 1.26 | 83.8 | 21.9 | 122 |
7 | FTO/TiO2/CsPbI3/F3EAI/spiro-OMeTAD/Au | NIP | 20.6 | 1.2 | 82.9 | 20.5 | 114 | |
8 | FTO/TiO2/m-Al2O3/CsPbI3/CF3-PAI/carbon | NIP | 20.5 | 1.14 | 78.2 | 18.3 | 136 | |
9 | FTO/c-TiO2/CsPbI2.75Br0.25/NPA 2D-RP/P3HT/Ag | NIP | 20.6 | 1.15 | 83.5 | 19.8 | 135 | |
10 | FTO/TiO2/4A1N/CsPbI3/spiro-OMeTAD/Ag | NIP | 20.9 | 1.2 | 80 | 20 | 138 | |
11 | ITO/PEDOT:PSS/CsPb0.5Sn0.5I2Br/F-TBA/PCBM/BCP/Ag | PIN | 23.12 | 0.83 | 73 | 14.01 | 112 | |
12 | Heterojunctions (3D/low-D) | FTO/TiO2/CsPbI3/BDABr/spiro-OMeTAD/Au | NIP | 20.8 | 1.18 | 83.6 | 20.6 | 142 |
13 | ITO/PTAA/CsPbI3/Cs4PbBr6/CPTA/BCP/Ag | PIN | 20.2 | 1.23 | 84.8 | 21 | 134 | |
14 | Heterojunctions (compositional) | FTO/TiO2/CsPbI2Br/CsPbI3−xBrx/PTAA/Au | NIP | 18.1 | 1.15 | 80.9 | 16.8 | 168 |
15 | ITO/SnO2/CsPbI3−xBrx/CsF/spiro-OMeTAD/Au | NIP | 19.4 | 1.27 | 85.3 | 21 | 116 | |
16 | Heterojunctions (beta/gamma) | ITO/MeO-2PACz/β-CsPbI3/γ-CsPbI3/PCBM/BCP/Ag | PIN | 20.6 | 1.16 | 84.2 | 20.2 | 154 |
17 | FTO/TiO2/SnO2/β-CsPbI3:Zn(C6F5)2/γ-CsPbI3:GAI/P3HT + SMeTAPyr/Au | NIP | 21.7 | 1.22 | 81.5 | 21.6 | 159 |
![]() | ||
Fig. 17 The review of efficiency enhancement via modifying interface charge carrier dynamics. (a) Improving energy level alignment at the perovskite/ETL interface via a p- to n-type transition at the perovskite surface with propylamine hydrochloride (PACl) treatment (reproduced with permission from Energy & Environmental Science, 2023, 16, 2572–2578. Copyright The Royal Society of Chemistry 2023).183 (b) Improving hole transfer at the perovskite/HTL interface via introducing a dipole layer with highly polar 2,2,2-trifluoroethylammonium iodide (F3EAI) treatment at the perovskite surface (reproduced with permission from Advanced Materials, 2022, 34, 2202735. Copyright 2022 Wiley-VCH GmbH).114 (c–e) Improved charge transfer induced by internal p–n heterojunctions at the inorganic perovskite layer. (c) 3D-CsPbI3/0D-Cs4PbI6 heterojunctions via surface reconstruction with benzyldodecyl-dimethylammonium bromide (BDABr) treatment at the perovskite surface (reproduced with permission from Science Bulletin, 2023, 68, 706–712. Copyright 2023 Science China Press).142 (d) Composition graded CsPbI3−xBrx heterojunction via spray deposition of CsPbI3 on top of a CsPbI2Br layer (reproduced with permission from Joule, 2021, 5, 481–494. Copyright 2020 Elsevier Inc.).168 (e) Phase heterojunction formed by a two-step process of spin coating β-CsPbI3 followed by vacuum evaporation of γ-CsPbI3 (reproduced with permission from Nature Energy, 2022, 7, 1170–1179. Copyright R. Ji, Y. Vaynzof et al. 2022).184 |
Other studies of interface modification have focused on treating the interface with molecules that exhibit strong dipoles such as when a dipole monolayer is formed, which could either facilitate charge extraction or hinder charge back-recombination.112,114,122,135,136,138 Zhang et al. treated the CsPbI3 perovskite surface with highly polar 2,2,2-trifluoroethylammonium iodide (F3EAI) molecules, which they expected to provide a favorable path for hole extraction at the perovskite/HTL interface and yield a high PCE of 20.50%, as described in Fig. 17b.114 It was shown that polar passivators would orient the positive dipole towards the perovskite surface, effectively attracting and passivating negatively charged defects that would otherwise trap holes and hinder extraction. Zhang et al. also treated the CsPbI3/carbon–electrode interface with 4-trifluoromethyl-phenylammonium iodide (CF3-PAI), which would align itself so that the positive dipole of the CF3 group interacts with the perovskite surface.136 Here, the authors showed that, while a direct contact between the perovskite and the carbon electrode layers would lead to back recombination since electrons could also cross the interface, the presence of the dipole layer would then deter electron transfer and ensure that only holes could selectively cross the interface. Meanwhile, some studies have demonstrated that dipole layer treatment at the buried interface could also be a viable strategy. In 2023, Wang et al. used polar molecules of 4-amino-1-naphthalene sulfonate (4A1N) to functionalize the TiO2 substrate and subsequently interact with the CsPbI3 perovskite layer via the amino group.138 The resulting ion–dipole interaction yielded a large shunt resistance across the interface, even under low light conditions, thus preventing back-recombination across the interface and enabling a high PCE of 20.03% and a high indoor photovoltaic PCE of 41.21% with an output power of 137.66 μW cm−2.
In 2023, Chen et al. treated a CsPbI3 perovskite film with a benzyldodecyl-dimethylammonium bromide (BDABr) salt to demonstrate a surface reconstruction process where in situ 3-dimension-to-0-dimension phase transformation was observed.142 While the 0D Cs4PbI6 phase is generally detrimental to the perovskite film due to its photo-inactivity, in this case, the 0D phase would actually accumulate on the grain boundaries to passivate pre-existing defects and also slightly shift the surface energy level such that the abundant 3D/0D heterojunctions at the surface instead aided in efficient hole transfer, as shown in Fig. 17c. Due to the reduced defects and increased charge transfer, the optimized device showed a high PCE of 20.63%. A similar concept was later demonstrated by Heo et al. where 0D Cs4PbBr6 perovskite NCs were synthesized and spray-deposited on the surface of CsPbI3 films.134 The authors showed that the 3D/0D heterojunctions could effectively reduce lattice strain, while also providing an electron-selective junction for efficient electron extraction, and the resulting p–i–n solar cell devices yielded a high PCE of 18.78%.
Besides dimension-engineered heterojunctions, the formation of heterojunctions via compositional control was also studied. Heo et al. suggested a creative approach towards creating composition-graded heterojunctions, where a thin film of CsPbI2Br was first spray-deposited on a TiO2 substrate, followed by additional spray-deposition of CsPbI3 in order to create a CsPbI3−xBrx film with an internal graded heterojunction where the halide ratios (and the effective energy levels) changed with depth, as shown in Fig. 17d.168 The fully spray-coated process not only provided an ideal band structure alignment for efficient charge extraction through an enlarged depletion region, where the unit cell device exhibited the best PCE of 16.81%, but also allowed for the realization of a stable process for efficient fully scalable inorganic perovskite photovoltaic devices. In 2023, Chu et al. reported another method of achieving composition-graded heterojunctions for CsPbIxBr3−x solar cells with a high PCE of 21.02% and high VOC of 1.27 V.116 Here, the authors showed that CsF salt treatment on CsPbIxBr3−x films, followed by annealing, could trigger a solid-state reaction between the perovskite surface and CsF salt, proven by the wider bandgap measured at the surface. Through optimized parameters, the authors reported a high PCE of 21.02%.
A unique merit of CsPbI3 inorganic perovskites is that the perovskite material can exhibit multiple photo-active “black” phases, which can be aligned with each other for efficient charge carrier extraction. In 2022, Ji et al. reported PHJ solar cells by creating a γ-CsPbI3/β-CsPbI3 PHJ.184 The PHJ was fabricated through a two-step process, where the β-CsPbI3 layer was first solution-processed through a spin deposition procedure with a DMAI additive and then the subsequent γ-CsPbI3 layer was vacuum evaporated for deposition on the β-CsPbI3 layer, as seen in Fig. 17e. Through careful optimization of the thickness of each perovskite layer, the authors achieved a high PCE of 19.1%. Mali et al. also demonstrated γ-CsPbI3/β-CsPbI3 PHJ solar cells in 2023 and increased the device PCE to an impressive 21.59%.159 In order to further stabilize both layers and optimize energy level alignment, the authors incorporated a bis(pentafluorophenyl)zinc (Zn(C6F5)2) p-type additive within the bottom β-CsPbI3 layer, while also co-evaporating a small amount of guanidinium iodide (GAI) alongside the other precursors for the fabrication of a GAI-incorporated γ-CsPbI3 layer.
No. | Strategy (narrow/wide) | Perovskite top cell | Type | Perovskite unit cell | Tandem cell PCE [%] | Ref. | |||
---|---|---|---|---|---|---|---|---|---|
JSC [mA cm−2] | VOC [V] | FF [%] | PCE [%] | ||||||
1 | Si/Perovskite tandem (2T) | ITO/NiOx/CsPbI2.85Br0.15/ABA/PCBM/BCP/Ag | PIN | 19.92 | 1.222 | 83.67 | 20.38 | 25.31 | 185 |
2 | ITO/SnO2/CsPbI3−xBrx/NiI2/spiro-OMeTAD/Au | NIP | 17.88 | 1.36 | 80.54 | 19.53 | 22.95 | 105 | |
3 | ITO/SnO2/CsPbI2.85Br0.15:PC/ODADI/spiro-OMeTAD/Au | NIP | 20.40 | 1.342 | 80.59 | 22.07 | 27.27 | 94 | |
4 | ITO/NiOx/P3CT-N/CsPbI2.85Br0.15:ABF/PCBM/BCP/Ag | PIN | 19.62 | 1.247 | 84.97 | 20.80 | 26.26 | 187 | |
5 | CIGS/Perovskite tandem (4T) | FTO/P3CT/CsPbI3/PCBM/BCP/Ag | PIN | 18.88 | 1.10 | 84.09 | 17.46 | 24.75 | 186 |
6 | OPV/Perovskite tandem (2T) | ITO/NiOx/CbzNaph/CsPb(IxBr(1−x))3:(AQS:FPEA)/C60/BCP/Ag | PIN | 18.07 | 1.29 | 79.74 | 18.59 | 23.24 | 89 |
No. | Strategy (narrow/wide) | Top cell (wide Eg) | Type | Top unit cell | Tandem cell PCE [%] | Ref. | |||
---|---|---|---|---|---|---|---|---|---|
Bottom cell (narrow Eg) | Bottom unit cell (filtered by the top cell) | ||||||||
Bottom unit cell (unfiltered) | |||||||||
JSC [mA cm−2] | VOC [V] | FF [%] | PCE [%] | ||||||
7 | Perovskite/perovskite tandem (4T) | ITO/NiOx/CsPbI2Br/Ti0.9Sn0.1O2/IZO/MgF2 | PIN | 15.29 | 1.25 | 79.27 | 15.14 | 19.61 | 99 |
7.43 | 0.85 | 70.74 | 4.47 | ||||||
ITO/NiOx/CsPb0.6Sn0.4I2Br:DCD/ZnO/PCBM/Ag | 22.67 | 0.87 | 71.73 | 14.17 | |||||
8 | ITO/SnO2/CsPbI1.5Br1.5/spiro-OMeTAD/MoOx/ITO | NIP | 11.67 | 1.29 | 74.8 | 11.26 | 18.07 | 170 | |
11.61 | 0.85 | 72.5 | 7.15 | ||||||
ITO/PEDOT:PSS/CsPb0.5Sn0.5I2.7Br0.3/PCBM/BCP/Ag | 22.11 | 0.86 | 72.7 | 13.82 |
For example, Wang et al. demonstrated 2-terminal (2T) silicon/perovskite (Si/P) tandem cells, where they used a CsPbI2.85Br0.15 perovskite with a bandgap of 1.71 eV, which is considered ideal for the Si/P architecture, as shown in Fig. 18a.185 Here, the wide-bandgap inorganic perovskite was processed as the top cell in conjunction with the narrow-bandgap silicon bottom cell, and the resulting tandem device exhibited a high PCE of 25.31%. The same group further improved their design by etching away the defect-rich surface layer of the inorganic perovskite film by spin dripping with 3-amino-5-bromopyridine-2-formamide (ABF) in methanol solution, which, in tandem with the bottom silicon layer, now showed a higher PCE of 26.26%.187 Soon afterwards, by regulating crystallization of the inorganic perovskite layer, the same group could suppress more defects within the perovskite film, which was again applied to a Si/P tandem device with a high PCE of 27.27% and an impressive VOC of 2.024 V.94 Recently, Zhang et al. used CsPbI3 as the top cell and Cu(In,Ga)Se2 (CIGS) as the bottom cell and reported fabricating 4-terminal (4T) CIGS/Perovskite (CIGS/P) tandem cells, which is shown in Fig. 18b.186 This was the first report on 4T CIGS/P tandem cells, and the authors reported a total PCE of 24.75%. Meanwhile, Li et al. fabricated 2T Organic/Perovskite (O/P) tandem cells in 2024, where they used the CsPbI1.8Br1.2 inorganic perovskite with a bandgap of 1.78 eV.89 The authors used a 9,10-anthraquinone-2-sulfonic acid (AQS) additive to suppress voltage loss within the perovskite cell by regulating the crystallization process with the sulfonic group and reported a PCE of 23.24% for the O/P tandem cell, as shown in Fig. 18c.
![]() | ||
Fig. 18 The review of efficiency enhancement via multi-junction inorganic perovskite solar cells. (a) Silicon/CsPbI2.85Br0.15 2T tandem solar cell and its performance (reproduced with permission from Advanced Materials, 2023, 35, 2300581. Copyright 2023 Wiley-VCH GmbH).185 (b) CIGS/CsPbI3 4T tandem solar cell and its performance (reproduced with permission from the Journal of Energy Chemistry, 2024, 97, 622–629. Copyright 2024 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences).186 (c) Organic/CsPbI1.8Br1.2 2T tandem solar cell and its performance (reproduced with permission from Angewandte Chemie, 2024, 63, e202412515. Copyright 2024 Y. Li, J. Yin, S. Wu, A. K.-Y. Jen et al.).89 (d) CsPb0.5Sn0.5I2.7Br0.3/CsPbI1.5Br1.5 4T all-inorganic perovskite-only tandem solar cell and its performance (reproduced with permission from ACS Energy Letters, 2022, 7, 4215–4223. Copyright 2022 American Chemical Society).170 (e) CsPb0.6Sn0.4I2Br/CsPbI2Br 4T all-inorganic perovskite-only tandem solar cell and its performance (reproduced with permission from the Chemical Engineering Journal, 2023, 452, 139697. Copyright 2022 Elsevier B.V.).99 |
In 2022, Sun et al. demonstrated an all-perovskite 4T tandem structure (P/P) based only on inorganic Pb-based and Pb–Sn-based perovskites for the first time, where they used a wide-bandgap (1.98 eV) CsPbI1.5Br1.5 perovskite for the top cell and a narrow-bandgap (1.39 eV) CsPb0.5Sn0.5I2.7Br0.3 perovskite for the bottom cell.170 As shown in Fig. 18d, the bottom Pb–Sn perovskite cell exhibited a PCE of 7.15% from the light filtered by the top cell, and the 4T P/P tandem structure yielded a total PCE of 18.07%. Soon after, in 2023, Wen et al. also reported a 4T P/P tandem structure using all-inorganic perovskites only.99 Here, CsPbI2Br was chosen as the top wide-bandgap cell (1.92 eV), and by optimizing the Pb–Sn ratio, the authors found that CsPb0.6Sn0.4I2Br with a bandgap of 1.54 eV resulted in the highest performing bottom narrow-bandgap cell, as shown in Fig. 18e. The total PCE of the optimized 4T all-inorganic P/P tandem device was reported at 19.61%, where the PCE of the top cell was 15.14% and the PCE of the filtered bottom cell was 4.47%.
Recent research efforts have demonstrated progress in enhancing the stability and efficiency of inorganic perovskite solar cells through various strategies, including compositional engineering, additive incorporation, surface post-treatment, and interface optimization. Compositional engineering strategies effectively mitigate lattice strain and enhance thermodynamic stability by introducing foreign ions into perovskite structures. Additive engineering, employing multifunctional molecules, has proven successful in modulating crystallization kinetics, suppressing Sn2+ oxidation, and passivating defects. Surface post-treatment, achieved using multifunctional organic molecules and ionic salts, effectively suppresses surface defects and enhances perovskite lattice stability and energy-level alignment through surface reconstruction. Additionally, environmental protection layers formed through post-treatment prevent moisture ingress, thereby preserving perovskite phase stability under challenging environmental conditions. Interfacial optimization at the ETL and HTL interfaces and low-dimension heterostructure interfaces significantly reduce interfacial defects, alleviate lattice strain, improve energy-level alignment, enhance charge carrier dynamics, and suppress detrimental ion migration, thereby enhancing overall device stability and efficiency (Fig. 19a and b).
Furthermore, inorganic perovskites, owing to their suitable bandgap for tandem solar cells and thermal stability, have shown great promise as top and bottom cells in multi-junction tandem solar cells. Recent studies demonstrate their successful integration into 2T and 4T tandem configurations with silicon, CIGS, organic photovoltaics, and even other inorganic perovskites.
Despite recent progress, inorganic perovskite solar cells still exhibit lower efficiencies than hybrid perovskite solar cells. Therefore, continued research is required to further improve the efficiency of inorganic perovskite solar cells, along with the development of tandem solar cells to overcome the theoretical efficiency limit of single-junction devices. In particular, research on Sn- and Sn–Pb-based inorganic perovskites, with their theoretically ideal bandgaps for single-junction operation and compatibility as bottom cells in tandem architectures, is regarded as essential to advancing the field.
To further advance the development of inorganic perovskite solar cells, it is essential to investigate detailed degradation mechanisms associated with phase instability, interfaces, and grain boundaries. Precise defect control using novel additives and multifunctional passivation agents is required, while strategies such as using internal graded heterojunctions, low-dimensional heterostructure integrations, and interface dipole modifications hold great promise for substantially improving both efficiency and stability.
For the commercialization of inorganic perovskite solar cells, further research is needed on using inorganic CTLs for the development of all-inorganic devices, since inorganic CTLs are mechanically and chemically more robust compared to organic CTLs and do not require the use of hygroscopic additives, as well as HTM-free carbon electrode architectures, which offer improved stability and reduced fabrication costs. In addition, the development of scalable strategies to maintain high efficiency and stability in large-area devices and modules and advanced encapsulation technologies to ensure robust operational stability under extreme environmental conditions is essential.
Finally, the extended application of inorganic perovskite solar cells can be envisioned in the field of indoor photovoltaics (IPVs), as they have similar device structures and operational principles. The relatively wide bandgaps and high open-circuit voltages of inorganic perovskites make them well suited for photoelectric conversion under low-intensity, narrow-spectrum indoor lighting, positioning them as strong candidates for indoor photovoltaic applications. However, achieving high IPV performance requires addressing unique challenges such as optimizing charge transport under low light, spectral matching to artificial light sources, minimizing non-radiative recombination losses, and developing lead-free devices by replacing toxic Pb-based compositions. The strategies discussed in this review—including compositional engineering, defect passivation, and interfacial engineering—can be effectively adapted to meet these demands, offering a valuable foundation for extending the application of inorganic perovskites to diverse photovoltaic environments, including indoor environments.
This journal is © The Royal Society of Chemistry 2025 |