• Advanced Photonics
  • Vol. 7, Issue 3, 034003 (2025)
Feng Yang1,†, Yu Tong1,2, Kun Wang2,3,*, Yali Chen1..., Ziyong Kang1 and Hongqiang Wang1,*|Show fewer author(s)
Author Affiliations
  • 1Northwestern Polytechnical University, School of Materials Science and Engineering, Center for Nano Energy Materials, State Key Laboratory of Solidification Processing, Xi’an, China
  • 2Northwestern Polytechnical University, Chongqing Innovation Center, Chongqing, China
  • 3Northwestern Polytechnical University, School of Microelectronics, Xi’an, China
  • show less
    DOI: 10.1117/1.AP.7.3.034003 Cite this Article Set citation alerts
    Feng Yang, Yu Tong, Kun Wang, Yali Chen, Ziyong Kang, Hongqiang Wang, "Recent advances in tin perovskites and their applications," Adv. Photon. 7, 034003 (2025) Copy Citation Text show less

    Abstract

    Lead halide perovskites have started a new era for solar cells. However, the toxicity of lead poses a challenge for their practical applications. Replacing lead with tin provides a feasible way to reduce the toxicity of lead perovskites and can further promote the applications of perovskites. Due to their reduced toxicity with advantageous optical and electronic properties compared to their lead counterparts, tin (II) perovskites (TiPes) have attracted significant interest in recent years, not only for pursuing high-performance solar cells but also in other application fields. We aim to provide a comprehensive overview of recent advances in TiPes, covering fundamental crystal structure, optoelectronic properties, fabrication methods, and applications. A detailed comparison with lead perovskites is included, emphasizing TiPes’ unique strengths while presenting their application challenges. Finally, potential solutions to the challenges are proposed, along with a vision for their future development and potential.

    1 Introduction

    Metal halide perovskites have emerged as a class of semiconductor materials suitable for various electronic or optoelectronic applications.14 In the field of photovoltaics, perovskite solar cells (PSCs) have witnessed a rapid increase in power conversion efficiency (PCE) from 3.8% to over 26% in the past decade.57 Owing to the easy processibility and exceptional carrier properties of perovskites, their applications have also been extended to light-emitting diodes (LEDs), photodetectors (PDs), lasers, and even transistors.811 However, concerns remain within the scientific community regarding the practical application prospects of perovskites due to the high toxicity of water-soluble lead salts in their degradation products.12,13 Although physical or chemical encapsulation provides the feasibility to decrease lead leakage from perovskites to the environment, challenges still remain in the ultimate elimination of lead release. Substituting Pb with low-toxicity elements, such as tin (Sn), germanium (Ge), copper (Cu), or manganese (Mn), could be an alternative way to solve this problem. Among the low-toxicity elements, Sn, with its similar electronic configuration and ionic radius to Pb, is considered one of the most promising substitutes for lead counterparts.1417

    Tin (II) perovskites (TiPes) possess unique properties that make them attractive candidates for a wide range of applications. For instance, bulk TiPes feature a narrower bandgap compared with their lead counterparts, thus facilitating a broader spectrum coverage of the whole ultraviolet-visible-near infrared range.18,19 Furthermore, TiPes offer potential properties, such as a long hot carrier lifetime, which provides the possibility of breaking the Shockley–Queisser limit for solar cells.2022 The rise of TiPes stems from the need to replace toxic lead in perovskite solar cells. After development in the past decade, tin PSCs have achieved PCE exceeding 15%, representing the most advanced heavy metal-free solar cells to date.23 Alongside these breakthroughs, research on TiPes has expanded into other applications traditionally dominated by lead perovskites, such as LEDs, PDs, lasers, and transistors.2427 The rapid advancement of TiPes is largely attributed to the fundamental research on lead perovskites, with many key achievements built upon the latter’s successes. However, most existing review articles predominantly highlight the potential application advantages of TiPes in specific areas, often neglecting a thorough, one-to-one comparative analysis of the structure, property, and application differences between tin and lead perovskites.2831

    This review provides a comprehensive summary of TiPes, highlighting their properties and recent advancements across various applications. It covers the crystal structure and optoelectronic properties of TiPes, methods for preparing thin films, single crystals, and nanocrystals, as well as their current applications in solar cells, LEDs, PDs, lasers, and transistors. In addition, a detailed comparison with lead perovskites is presented, clarifying the strengths and weaknesses of TiPes and offering clear guidance for material selection in specific application scenarios. Finally, this review discusses the key challenges faced by TiPes, proposes potential solutions, and envisions their future development trajectory.

    2 Structure and Properties

    Perovskites are a class of compounds with ABX3 chemical formula, where A represents monovalent cations, B primarily includes Pb or Sn ions, and X stands for halogen anions.32 The perovskite structure, as depicted in Fig. 1(a), consists of a three-dimensional (3D) inorganic network created by co-angled BX6 octahedra, with A-site cations occupying the octahedral voids. Lead perovskites derive their excellent physical properties from the 6s2 lone pair of electrons of Pb2+. Being the same group element, Sn shows similar electron arrangements and properties to Pb, and due to its low biological toxicity, it is considered a promising substitute for Pb.35 Although TiPes exhibit structural and property similarities to their lead counterparts, they also possess distinctive performance characteristics, attracting considerable interest from researchers.36 To provide a clearer comparison, we have compiled a summary table highlighting the differences between tin and lead perovskites in terms of optoelectronic properties and room-temperature (RT) crystal structures (Table 1). TiPes have garnered significant attention due to their higher light absorption coefficients and broader bandgap tunability, in addition to their low toxicity, which is a key driver for research interest.17 The table provides an overview of these fundamental properties, which will be discussed further in detail.

    Crystal structure of perovskites. (a) Crystal structure of ABX3 perovskites. (b) Tolerance factors (t) of a series of halide perovskites. (c) Schematic diagram of the temperature-induced phase transition of FASnI3. Reproduced with permission from Ref. 33. (d) XRD patterns of MASnI3 during two sequential compression−decompression cycles. Reproduced with permission from Ref. 34.

    Figure 1.Crystal structure of perovskites. (a) Crystal structure of ABX3 perovskites. (b) Tolerance factors (t) of a series of halide perovskites. (c) Schematic diagram of the temperature-induced phase transition of FASnI3. Reproduced with permission from Ref. 33. (d) XRD patterns of MASnI3 during two sequential compression−decompression cycles. Reproduced with permission from Ref. 34.

    PerovskiteBandgap (eV)Absorption coefficient (cm1)Exciton binding energy (meV)Mobility (cm2V1s1)Diffusion length (μm)RT crystal structure
    FASnI31.3510531Electron: 103 (Hall effect)Orthorhombic/cubic
    FASnBr32.495Orthorhombic
    FASnCl33.55
    MASnI31.23104 to 10529Hole: 50 (Hall effect)Electron: 0.279, hole: 0.193Tetragonal
    MASnBr32.1565Orthorhombic
    MASnCl33.69257Triclinic
    CsSnI31.310518Hole: 585 (Hall effect)0.016Orthorhombic (B-γ/Y phase)/cubic
    CsSnBr31.8Orthorhombic
    CsSnCl32.8Cubic
    FAPbI31.498 to 35Hole: 35 (SCLC)Cubic
    FAPbBr32.1522 to 60Hole: 62 (SCLC)Orthorhombic
    FAPbCl33.02110Orthorhombic
    MAPbI31.559.1×1042 to 50Electron: 66 (Hall effect)Electron: 0.129, Hole: 0.105Tetragonal
    MAPbBr32.2125 to 150Hole: 24 (SCLC)1.3 to 4.3Cubic
    MAPbCl32.8869Hole: 42 (SCLC)3.0 to 8.5Cubic
    CsPbI31.73>104201Cubic
    CsPbBr32.39.8×10437Electron: 52 (SCLC)0.08Orthorhombic
    CsPbCl33.0167

    Table 1. Main physical properties of tin and lead perovskites (the data originate from Refs. 3747" target="_self" style="display: inline;">47).

    2.1 Crystal Structure

    Crystal structure stability is the key to the performance of TiPes devices, and the well-known Goldschmidt tolerance factor (t) is usually used to roughly estimate the stability of the perovskite phase structures, as illustrated in Eq. (1).48t=RA+RX2(RB+RX),where RA, RB, and RX represent the ionic radii of the A, B, and X ions, respectively. Empirical data reveal that maintaining t values between 0.8 and 1 preserves the 3D perovskite structure, with the ideal cubic structure achieved between 0.9 and 1. Deviation from this range can result in phase transitions or lower-dimensional structural transformations. Among the currently spotlighted TiPes, the A-site of the ABX3 formula is often occupied by spacer organic cations such as methylamine ions [(CH3NH3)+, r=2.17  ] or formamidine ions [(CH(NH2)2+, r=2.53  ]. Inorganic cations Cs+ are typically introduced into octahedral voids to replace organic cations and enhance thermal stability.4951 The halogen ions that occupy the X-site anion are usually Cl, Br, and I. Furthermore, given the extensive use of SnF2 for tin compensation, it is important to examine the impact of F on the TiPes crystal structure. For instance, Pascual et al. showed that fluoride plays a crucial role in the reduction of Sn4+. F selectively forms complexes with Sn4+, resulting in SnF4, and the likelihood of introducing SnF4 complexes into perovskite lattice is lower compared to Sn4+.52 Although substituting Pb with Sn appears to increase t and should theoretically promote the formation of a stable cubic structure [Fig. 1(b)], poor stability is generally observed in the cubic TiPes. It is essential to note that the t-value alone is a necessary but not sufficient condition for determining the stability of the perovskite structure. Other non-geometric factors, such as bond valence and chemical stability, are also crucial.53,54

    As is known, perovskites could undergo structural and symmetry changes with temperature, and high temperatures typically favor an improved cubic symmetry. For TiPes, their phase transition from cubic to tetragonal, orthorhombic, or further to rhombohedral or monoclinic, usually occurs when the temperature is decreased. A few soft-lattice TiPes experience more intricate transitions. For example, MASnCl3 undergoes a continuous phase transition from cubic (Pm3m), through rhombohedral (P3m) and monoclinic (Pc), to triclinic (P1) in a narrow temperature range near 300 K.55 Temperature-induced phase transition behavior of FASnI3 in a single crystal with a size of about 0.1 mm was investigated by Kahmann et al.33 As shown in Fig. 1(c), FASnI3 has a cubic (Pm3m) crystal structure at room temperature, where FA molecules are free to rotate and in a completely disordered state. Upon cooling, FA molecules are gradually oriented in the ab plane and the crystal undergoes a phase transition to a tetragonal (P4/mbm) system at 255 K. With further temperature reduction, FA molecules become fully ordered, resulting in a transition to P4bm space group at 155 K. However, this is partially inconsistent with the earlier report on FASnI3 single crystals by Stoumpos et al., in which a transition of FASnI3 from an orthorhombic (Amm2) crystal system at room temperature to a low-symmetry Imm2 space group at 180 K was demonstrated.56 Besides single crystal, a phase transition from a tetragonal (P4/mbm) to an orthorhombic (Pnma) crystal system at 100 K has been reported.57 This difference in crystal structure may be inextricably linked to defects and the size of the crystal.33 Besides temperature, pressure also serves as an external factor that governs the crystal structure and symmetry. Figure 1(d) illustrates the XRD patterns of MASnI3 during successive compression and decompression. For MASnI3, the tilt of the SnI6 octahedra gradually increases with pressure, promoting the crystal structure shift from tetragonal to orthorhombic. As pressure exceeds 3 GPa, MASnI3 shows a disordered crystal structure and transforms into the amorphous phase. Remarkably, the crystalline properties of MASnI3 are preserved even above 30 GPa during secondary compression, showcasing a significant enhancement in structural stability.34 This stability improvement is ideal for application in optoelectronic devices, but the introduction of exogenous stresses is difficult. Fortunately, this can be alternatively realized by external doping, which introduces stresses, thus improving the stability.58,59

    In recent years, the attractive physical properties unlocked by nanoscale effects have fueled a rapid surge in research aimed at reducing the size and dimension of perovskites [Fig. 2(a)].60 Perovskites with various sizes and dimensions, including microcrystals, nanosheets, nanowires, nanocrystals, and quantum dots, have been effectively synthesized.6164 Usually, “smaller” TiPes tend to maintain the ABX3 structural framework and retain the crystal structure of the parent bulk material.65 For example, in recent work, it was verified that CsSnI3 nanocrystals with a lateral size of 10 nm possess the same stable γ-orthorhombic phase as the parent bulk.66 Based on the ABX3 framework, the introduction of long-chain ammonium cations [typical examples shown in Fig. 2(b)] leads to a reduction in the dimensionality of TiPes, e.g., down to two or quasi-two dimensions (2D or quasi-2D). Their typical structural formula is A2An1BnX3n+1 (RP phase) or AAn1BnX3n+1 (DJ phase), with A representing the introduced long-chain cation, and n indicating the number of octahedral [BX6]4 layers.67 Organic cations with different chain lengths lead to partial distortion of the inorganic octahedra, reducing the symmetry of the TiPes crystal structure and also affecting the TiPes crystal orientation, size, and growth distribution.68 Low-dimensional TiPes have garnered significant attention for their exceptional stability, primarily attributed to the excellent moisture resistance offered by the organic intercalation layer. Moreover, this intercalation layer serves as a barrier, to effectively suppress ionic migration and further enhance the stability. In addition, the quantum well structure and energy funneling effect endow them with potentially good energy transfer capability and high radiative recombination efficiency, which could be beneficial for photovoltaic and optoelectronic applications.69 Nevertheless, the out-of-plane charge transfer in these materials is constrained, necessitating precise adjustments in their composition and crystalline orientation for practical applications.

    TiPes with different dimensions. (a) Schematic diagram illustrating the reduction in size and dimensionality of TiPes. (b) Typical examples of ammonium cations used to reduce the dimensionality of TiPes.

    Figure 2.TiPes with different dimensions. (a) Schematic diagram illustrating the reduction in size and dimensionality of TiPes. (b) Typical examples of ammonium cations used to reduce the dimensionality of TiPes.

    2.2 Charge Carrier and Exciton Properties

    For perovskites, under external excitation (light, voltage, radiation, etc.), electrons in the valence band absorb energy (EEg) and are excited to the conduction band, yielding free charge-carriers or excitons.70 Some of the carriers are excited to energy levels higher than the valence band maximum (VBM) or conduction band minimum (CBM), forming hot carriers. These hot carriers rapidly lose energy and relax back to the VBM or CBM through coupling with lattice vibrations (phonons). Four stages are usually involved in this process, i.e., hot carrier generation, thermalization (reaching thermal equilibrium through interactions and energy exchange), dephasing (eliminating coherence between ground and excited states), and cooling (interaction with the lattice, releasing heat) [Fig. 3(a)].72,73 If the hot carriers can be extracted for application before they relax to the band edge, it would overcome the SQ limit and bring a revolutionary breakthrough for photovoltaic technology.20,73,74

    Charge carrier and exciton properties. Schematic diagram of (a) carrier generation (including hot carrier generation) and (b) carrier recombination. (c)–(e) Energy-dependent and time-dependent photoluminescence spectra of perovskite thin films at low and high excitation densities for FASnI3, FAPbI3, and MAPbI3, showing that the FASnI3 thin films exhibit a much stronger band-filling effect and prolonged hot carrier emission. Reproduced with permission from Ref. 21. (f) Dark conductivity spectra (real part). (g) OPTPS measurements of the charge-carrier recombination dynamics of FASnI3 thin films treated with SnF2. Reproduced with permission from Ref. 71.

    Figure 3.Charge carrier and exciton properties. Schematic diagram of (a) carrier generation (including hot carrier generation) and (b) carrier recombination. (c)–(e) Energy-dependent and time-dependent photoluminescence spectra of perovskite thin films at low and high excitation densities for FASnI3, FAPbI3, and MAPbI3, showing that the FASnI3 thin films exhibit a much stronger band-filling effect and prolonged hot carrier emission. Reproduced with permission from Ref. 21. (f) Dark conductivity spectra (real part). (g) OPTPS measurements of the charge-carrier recombination dynamics of FASnI3 thin films treated with SnF2. Reproduced with permission from Ref. 71.

    Fang et al. discovered the photoluminescence of long-lived hot carriers (a few ns) in FASnI3 and observed a large blue shift with increasing excitation energy in the time-integrated photoluminescence energy spectrum.21 They attributed this phenomenon to the slow hot carrier relaxation and strong dynamic band-filling effects. Notably, the hot carrier lifetime in FASnI3 is longer than that of lead counterparts, highlighting its potential for hot carrier applications in TiPes [Figs. 3(c)3(e)].21,75 In addition, Dai et al. reported that for small-sized FASnI3 nanocrystals with quantum confinement, the high-energy states are unaffected by the fast relaxation and exhibit a separated energy level structure. At low injected carrier densities (<1 carrier pair per nanoparticle), the relaxation of hot carriers in nanocrystals is slowed down by two orders of magnitude, and the hot carriers are preserved up to 15  ps.22 Moreover, studies have shown that doping tin into lead perovskite nanocrystals significantly retards the hot carrier relaxation process, which can be primarily attributed to several aspects as follows. First, the introduction of tin creates a new high-energy band structure that alters energy relaxation pathways and effectively suppresses rapid cooling. Second, tin doping weakens the coupling between carriers and optical phonons, reducing the efficiency of energy transfer to the lattice. Third, it decreases phonon-phonon scattering (specifically, the decay of optical phonons into acoustic phonons, known as Klemens decay), further slowing the lattice’s energy absorption. Fourth, tin doping lowers the lattice thermal conductivity, particularly the transport efficiency of acoustic phonons, thereby reducing the dissipation of thermal energy. These synergistic effects significantly slow down the hot carrier relaxation in tin-doped lead nanocrystals, further highlighting the unique potential of tin perovskites for hot carrier-related applications.76

    Bulk APbI3 perovskites are known for their relatively small exciton binding energy. Similarly, TiPes exhibit comparable or even lower exciton binding energies. For example, it has been reported that the exciton binding energies for the MASnI3, FASnI3, and CsSnI3 are 29, 31, and 18  meV, respectively. This means that excitons in TiPes are more likely to dissociate into free carriers, which is beneficial for photovoltaics and photodetection applications.71,77 Moreover, TiPes are theoretically predicted to exhibit superior electrical properties compared to their lead counterparts. For instance, MASnI3 demonstrates intrinsic electron and hole mobilities exceeding 2000 and 200  cm2V1s1, respectively, which are significantly higher than those of MAPbI3.78,79 However, the spontaneous p-type doping commonly observed in TiPes negatively impacts carrier performance, resulting in the actual mobility of TiPes being lower than that of their lead counterparts (Table 1). Theoretical and experimental studies show that in MASnI3, when the p-doping density is 1015  cm3, the carrier diffusion length can exceed 1  μm; when the p-doping density increases to 1018  cm3, the carrier diffusion length rapidly decreases to 30  nm.80

    The exciton binding energy of 3D TiPes is small, and the excitons can dissociate rapidly into free carriers at room temperature, which dominates the carrier recombination process.81 In an ideal case with other factors excluded, all carriers are expected to undergo radiative recombination by emitting photons. However, in the practical situation, there are many non-radiative recombination channels in perovskites, and most carriers usually take the priority to recombine via non-radiative processes, such as defect-assisted recombination, interface-induced recombination, Auger recombination, and band-tailed recombination [Fig. 3(b)].72 Because the radiative recombination is responsible for the luminescence performance, the photoluminescence quantum yields (PLQYs) of 3D perovskites are typically very low.37 From the lesson of lead perovskites, researchers have used dimensionality reduction or crystal size reduction to excite quantum and/or dielectric confinement effects in TiPes, which in turn increase the carrier density and exciton binding energy to induce high PLQYs.8284 Compared with other recombination processes, Auger recombination is usually weaker and has less impact on the device performance. The other three non-radiative pathways can be grouped into one category, namely single molecule relaxation. The rate constant of single-molecule complexation in TiPes is reported to be three orders of magnitude higher than that of lead counterparts, which stems from the instability and severe p-type self-doping of TiPes.36 This can be one of the reasons for the inferior performance of TiPes devices compared with the lead perovskite devices. Small amounts of SnF2 introduction typically contribute significantly to complementing Sn2+ and passivating defects.52,85,86 For instance, a decrease in dark current and a slowdown in photoconductivity decay are observed with the addition of small amounts of SnF2 to FASnI3 films, indicating a reduced doped hole concentration and an extended carrier lifetime [Figs. 3(f) and 3(g)].71 However, a particular note of caution is that the over-doping of SnF2 should be avoided because it may introduce additional non-radiative recombination channels.87 In recent years, many attempts have been made to improve the stability and reduce the doping level of TiPes, such as using reductive agents, introducing low-dimensional components, and adding passivation, which are detailed in the subsequent sections.8890

    2.3 Band Structure and Optical Properties

    In general, TiPes exhibit a similar energy band structure to lead perovskites. For example, in MASnI3, the VBM is contributed by the antibonding orbitals of the Sn 5s orbital and the I 5p orbital, dominated by I 5p; and the CBM is composed of the antibonding orbitals of the Sn 5p orbital and the I 5p orbital, dominated by Sn 5p [Fig. 4(a)].91,96 Various theoretical models have been applied to calculate the band gap of perovskites, and introducing spin-orbit coupling (SOC) effects into an efficient GW (G stands for Green’s function, and W denotes the dynamically screened Coulomb interaction) scheme achieves greater success in deriving the band gap of TiPes.97101 Angelis et al. used this approach to calculate the electronic density of states (DOS) of MASnI3 [Fig. 4(b)] and compared it with the lead perovskites. The substitution of Pb with Sn induces an upward shift of the atomic energy level, implying an upward shift of both VBM and CBM. However, the smaller splitting of the s and p states in the Sn atom makes the VBM more upshifted than the CBM, leading to a narrower band gap of MASnI3 than MAPbI3.65,92,93 The energy band structure of MASnI3 calculated by SOC-GW is shown in Fig. 4(c), indicating that it has a band gap of 1.1  eV in the Brillouin zone, which is close to the experimental value (1.2  eV). Similar to lead perovskites, the band gap of TiPes can be tuned directly by varying the A-site and X-site ions.102 Tao et al. established the band gap of the extensively studied tin and lead perovskites, as shown in Fig. 4(d).93 The effect of the halogen elements on the band gap is predominant, with the band gap increasing substantially when the X-site changes from I to Br and to Cl. The primary reason for this phenomenon is the downward shift of the p level in the halide, which results in a significant downward shift of the VBM. The octahedral tilt and distortion caused by the A-site ion is a secondary factor affecting the band gap, and this factor has a greater impact on TiPes than lead perovskites due to the smaller B-site ion radius. In TiPes with lattice structures showing large deviations from the ideal cubic one (FASnBr3, MASnCl3, and FASnCl3), the A-site can impose an even larger influence on the band gap than the B-site substitution.103 With these features, TiPes usually exhibit a narrower bandgap and broader bandgap tuning range in comparison with their lead counterparts, which makes it even more advantageous in meeting the ideal bandgap requirements for single-junction/multi-junction photovoltaic devices and tuning the emission colors of LEDs.

    Band structure and optical properties. (a) Energy band of MASnI3 and MAPbI3. Reproduced with permission from Ref. 91. Based on SOC-GW calculation: (b) electronic DOS of MASnI3 and (c) energy band structure of MASnI3 and MAPbI3. Reproduced with permission from Ref. 92. (d) Schematic energy level diagram of tin and lead perovskites. Reproduced with permission from Ref. 93. (e) Optical absorption and steady-state PL spectra of CsSnX3. Reproduced with permission from Ref. 94. (f) Optical absorption spectra and (g) PL emission spectra of TiPes [x represents the percentage of MA+ substituted by ethylenediamine ions (en)]. Reproduced with permission from Ref. 95.

    Figure 4.Band structure and optical properties. (a) Energy band of MASnI3 and MAPbI3. Reproduced with permission from Ref. 91. Based on SOC-GW calculation: (b) electronic DOS of MASnI3 and (c) energy band structure of MASnI3 and MAPbI3. Reproduced with permission from Ref. 92. (d) Schematic energy level diagram of tin and lead perovskites. Reproduced with permission from Ref. 93. (e) Optical absorption and steady-state PL spectra of CsSnX3. Reproduced with permission from Ref. 94. (f) Optical absorption spectra and (g) PL emission spectra of TiPes [x represents the percentage of MA+ substituted by ethylenediamine ions (en)]. Reproduced with permission from Ref. 95.

    As is well known in perovskites, halogen engineering is considered an effective option for tuning the band gap and optical properties, which is typically achieved by changing the precursor halide ratio or by anion exchange reactions. Figure 4(e) shows the PL emission of CsSnX3 (X=Cl, Cl0.5Br0.5, Br, Br0.5I0.5, I), covering the whole visible to the near-infrared (NIR) range.94 Compared with lead counterparts, TiPes have greater potential for applications in the NIR region (e.g., lasers and PDs).104,105 However, the PL intensity is extremely low and the stability obtained is unsatisfactory. Spanopoulos et al. substituted MA+ with ethylenediamine cations to trigger Sn and I vacancies without changing the structure to obtain 3D hollow TiPes.95 At room temperature, all hollow perovskites show good absorption of visible light with intensive PL emission, and the emission wavelength can be tuned from 991 to 718 nm [Figs. 4(f) and 4(g)]. The introduction of Sn and I vacancies leads to the formation of spatially confined domains in the perovskite structure, which in turn leads to lattice distortion and the generation of stable excitons, thus enhancing the PL.106 In addition, the introduction of the N element significantly improves the stability, and the resultant 3D hollow TiPes can stay stable in the air for at least 9 days. Apart from the doping modification of the native, reducing the dimensionality is also an effective approach to enhance the stability and optical properties.17 For example, 2D-layered perovskite (OCTAm)2SnBr4 (OCTAm for octylammonium cation) shows bright orange emission at 600 nm with a near-unit value of absolute. Notably, the PLQYs of this perovskite can remain unchanged after 240 days of storage at room temperature under ambient air and humidity conditions.107 This 2D structure provides spatial confinement for electrons and holes, leading to strong coulombic interactions, thus increasing the probability of radiative recombination. Moreover, the organic cations provide a barrier for hindering the oxygen and moisture invasion, thereby improving the stability.108 Besides, the lower-dimensional structure of TiPes also exhibits luminescence enhancement due to self-trapped exciton emission.81,82 It is also worthy to highlight the advantages of single-halogen TiPes in the field of pure red luminescence, with PL emission peak located in the range of 630 to 640 nm.83,109 Although higher external quantum efficiency (EQE) red LEDs have been reported in lead perovskites, they generally suffer from color impurity and phase separation.110,111 However, TiPes LEDs can avoid these issues and yield color-pure and stable red emission.

    2.4 Environmental and Operational Stability

    The stability of perovskites is influenced by both intrinsic and extrinsic factors. Intrinsic factors primarily involve ion migration within the perovskite, whereas extrinsic factors include the combined effects of environmental conditions such as water, oxygen, and temperature. Halogen ion migration is a critical factor impacting the stability of lead perovskites, whereas this issue is significantly suppressed in TiPes, which can be attributed to the unique defect environment of TiPes, i.e., tin vacancies.35 Theoretically, TiPes are expected to exhibit superior stability; however, extrinsic factors have a significantly detrimental effect on TiPes, especially in the presence of oxygen.112 For Pb, the insufficient shielding of the nuclear charge by the 4f electrons induces a lanthanide contraction, resulting in a smaller atomic radius and stronger binding of the 6s electrons [Fig. 5(a)]. By contrast, Sn, with a higher atomic energy level and lower electronegativity, has a smaller energy gap between the s and p orbitals and lacks the lanthanide contraction effect as Pb has.113 As a result, Sn’s outer electrons escape more easily, leading to an increased tendency for oxidative decomposition.31 As the humidity or temperature increases, the decomposition trend of perovskite is exacerbated, a phenomenon that is also observed in lead perovskites.113Figure 5(b) illustrates the degradation pathway of a typical TiPes device, which occurs in three main stages: precursor, perovskite, and device. Consequently, stabilization strategies are applied at each of these three stages.

    Environmental and operational stability. (a) Electronic structure of Sn and Pb atoms. (b) Schematic of typical TiPes device degradation. Reproduced with permission from Ref. 113. (c) Schematic diagram of vanillin reduction of TiPes. Reproduced with permission from Ref. 114. (d) Schematic diagram of 2D DJ-phase TiPes. Reproduced with permission from Ref. 115. (e) Schematic of the encapsulation of TiPes with poly(methyl methacrylate) and Al2O3. Reproduced with permission from Ref. 116.

    Figure 5.Environmental and operational stability. (a) Electronic structure of Sn and Pb atoms. (b) Schematic of typical TiPes device degradation. Reproduced with permission from Ref. 113. (c) Schematic diagram of vanillin reduction of TiPes. Reproduced with permission from Ref. 114. (d) Schematic diagram of 2D DJ-phase TiPes. Reproduced with permission from Ref. 115. (e) Schematic of the encapsulation of TiPes with poly(methyl methacrylate) and Al2O3. Reproduced with permission from Ref. 116.

    For precursors, several stabilization strategies have been developed. These include tin compensation strategies, such as adding SnF2 to offset the loss of Sn2+ due to oxidation; tin reduction strategies, such as introducing reductants to reduce Sn4+ [Fig. 5(c)]; ligand incorporation, for example, using organic ammonium salts to coordinate with Sn2+; and solvent substitution strategies, where the commonly used DMSO—known to oxidize Sn2+—is replaced with alternatives such as N,N-diethylformamide (DEF) and N,N-dimethylpropylene urea (DMPU).52,96,114,117 By contrast, lead precursors typically do not require oxidation-focused stabilization strategies. Instead, their stabilization primarily relies on ligand engineering, with Lewis bases coordinating with Pb2+ to enhance precursor stability.118 Although this approach resembles the ligand incorporation strategy used for tin precursors, important differences exist. Sn2+, as a soft Lewis acid, preferentially coordinates with soft bases, such as sulfur- or nitrogen-containing ligands. On the contrary, Pb2+, which lies between soft and hard Lewis acids, can coordinate with both soft bases and moderately hard bases, such as oxygen-containing ligands, to form stable coordination complexes.119 These differences in chemical behavior highlight the distinct requirements for stabilizing tin and lead precursors.

    Component engineering is an important tool for stabilizing perovskite. In TiPes, introducing large ammonium ions into the A-site is a commonly used and effective strategy to enhance material stability.120 As mentioned earlier, the introduction of ammonium cations with different valences results in the formation of distinct phases: the +1 valence ammonium cations lead to the RP phase, whereas the +2 valence cations form the DJ phase. Compared with RP-phase perovskites, DJ-phase generally exhibits improved stability. Park et al. incorporated diammonium organic spacers, 3-(aminomethyl)piperidinium iodide (3AMPI2) and 4-(aminomethyl)piperidinium iodide (4AMPI2), into TiPes, resulting in the formation of 2D DJ-phase perovskites, 3AMPSnI4 and 4AMPSnI4 [Fig. 5(d)].115 Specifically, 3AMPSnI4 and 4AMPSnI4 retain their crystal structures after 24 h of air exposure. By contrast, the 2D RP-phase PEA2SnI4 undergoes oxidation, as evidenced by the formation of a SnO2 phase. The DJ-phase perovskite exhibits excellent stability; however, the anchoring of diammonium cations to the perovskite slabs increases the rigidity of the structure, which may lead to significant structural distortion, thereby affecting the performance of the perovskite devices.81 In addition to ammonium cations, the incorporation of Cs+ also contributes to enhancing the thermal stability of perovskites. Particularly in the field of lead PSCs, efficient and stable perovskite formulations often rely on a combination of FA+ and Cs+.121,122 However, this strategy is more challenging to apply in TiPes because CsSnI3 is not resistant to oxidation and can be rapidly converted to the 1D orthorhombic phase in a humid environment.123 Nevertheless, studying CsSnI3 remains valuable, as it has the potential to contribute to improved thermal stability.

    Proper device structure design is also an effective means to improve stability. Unlike lead perovskites, variable-valence metal oxides are not suitable as transport layer materials in tin perovskite devices, as these may induce the oxidation of Sn2+.81 In general, organic transport materials exhibit better compatibility with TiPes. Furthermore, the interface between TiPes and transport layers requires careful modulation. For instance, introducing FACl can remove surface Sn4+ doping, whereas evaporating a thin Sn layer can reduce interfacial Sn4+.124,125 Interface engineering significantly benefits the suppression of tin oxidation and isolation from water and oxygen; however, research in this area for TiPes remains insufficient. In addition, drawing on the experience of encapsulating lead perovskites can also boost tin device stability. For example, Gahlot et al. achieved long-term stability in air for 15 days by encapsulating TiPes with a combination of the long-chain organic material poly(methyl methacrylate) and the inorganic material Al2O3 [Fig. 5(e)].116

    Beyond oxygen, moisture ingress poses another critical threat to perovskite stability. Water infiltration induces crystal lattice distortion, surface state degradation, defect proliferation, and material decomposition, severely compromising device efficiency and stability.126 Conventional encapsulation strategies (e.g., polymer or inorganic oxide coatings) could block moisture but may also hinder charge carrier transport due to insulating layers.127,128 Developing a suitable ligand engineering strategy can help solve this problem, in which Rao’s team suggests two principles for ligand design: (i) designing suitable carbon chain length to balance the need for hydrophobicity and charge transport and (ii) designing suitable terminal groups for robust binding to the perovskite surface.129131 Taking (11-mercaptoundecyl)-N,N,N-trimethylammonium bromide (MUTAB) as an example, this ligand modifies oleylamine/oleic acid-capped CsPbBr3 nanocrystals via a ligand exchange strategy.132 One end of the molecule passivates surface defects via strong Pb–S bonds, whereas the other end repels water molecules through steric hindrance, achieving dual-functional protection. CsPbBr3 nanocrystals based on this strategy can maintain stability in water for over 60 days while retaining excellent charge generation and transport efficiency. This strategy provides a paradigm for water stability research in lead-based perovskites. Although TiPes are still in the early stages of water stability research due to the easily oxidizable nature of Sn2+, the ligand engineering strategy can draw on the experience from lead-based systems, pointing the way toward the development of stable lead-free perovskite devices.

    2.5 Toxicity and Sustainability

    As research on metal halide perovskite advances, concerns have been raised regarding their potential toxicity. Babayigit et al. systematically investigated the environmental impacts of TiPes and their lead counterparts, identifying SnI2 and PbI2 as their primary degradation products, respectively.133 To assess the toxicity of these degradation products, the researchers introduced equal amounts of SnI2 and PbI2 into the aquatic environment of zebrafish Danio rerio embryos, regularly monitoring the embryos for lethality and defective phenotypes (collectively referred to as malformations). The results showed that both the lethal and teratogenic concentrations of SnI2 could show higher toxicity risk. Importantly, the study proposed a unique toxicity mechanism distinct from conventional heavy metal lead poisoning. The toxicity of SnI2 primarily arises from its degradation process: in the presence of H2O and O2, SnI2 breaks down into SnO2, accompanied by the formation of the intermediate product HI. The release of HI rapidly acidifies the aquatic environment by lowering the pH of the growth medium, thereby inducing harmful effects. The proposed reaction for SnI2 in aqueous medium is 8SnI2+4H2O+O22SnI4+6Sn(OH)I+2HI,SnI4+2H2OSnO2+4HI.

    This mechanism provides valuable insights into the environmental risks posed by TiPes and highlights the need for further investigation into their stability and safety.

    However, this experimental result is based on the assumption that SnI2 is the primary degradation product of TiPes. In reality, unlike Pb2+, Sn2+ is highly susceptible to oxidation due to the absence of the lanthanide contraction effect. In an oxygen-rich environment, TiPes preferentially undergo the following reactions.134,1352ASnI3+O2SnI4+SnO2+2AI.SnI4 can further react with AI, allowing it to be consumed: SnI4+2AIA2SnI6.

    Owing to this competitive response mechanism between Eqs. (3) and (5), the actual generation of HI is greatly reduced. Concerning HI is also produced during the degradation of lead perovskites, the induced impact difference of HI between TiPes and lead perovskites is decreased as well. Notably, the significant impact is the heavy metal issue with extreme toxicity in lead perovskite, regardless of its valence state. Moreover, Pb can persist in soil and water, accumulating over time and causing irreversible harm to biological systems.15,136 Therefore, TiPes is considered to be one of the good substitutes for lead perovskites given their heavy metal-free nature. It is worth mentioning that despite TiPes exhibiting lower toxicity compared with their lead counterparts, their environmental impact still needs to be further studied.

    Beyond toxicity, cost considerations are also crucial for practical applications. In the case of lead PSCs, for example, their high light absorption coefficient allows them to achieve high light absorption at a relatively thin film thickness, thus reducing the amount of Pb required per square meter of module to less than 1.4 g. If 1 terawatt of photovoltaic modules is deployed, it is estimated that only 7100  t of Pb will be needed. Given the abundant Pb reserves and mature recycling technologies, material costs can be significantly minimized.137 In comparison, TiPes have a higher light absorption coefficient and require a lower theoretical amount than lead.38 Although global tin reserves are lower than those of lead, they are still sufficient to meet industrial demand. Compared with its reserves, the greater concern lies in the high manufacturing cost of TiPes, primarily due to Sn’s susceptibility to oxidation. With the advancements in preparation processes and antioxidant technologies, the cost disadvantage of TiPes is expected to gradually diminish, rendering TiPes more competitive in future sustainable development.

    3 Fabrication Methods

    Generally, fabricating high-quality perovskite is a prerequisite for achieving high-performance devices, which means that the optimization of the TiPes preparation method is very important. This section provides an overview of the various types of approaches for preparing TiPes with different morphologies, including polycrystalline films, single crystals, and nanocrystals.

    3.1 Polycrystalline Thin Films

    TiPes films are typically prepared by a one-step deposition method or a sequential deposition method from their solution with precursor salts (e.g., SnX2, FAX, MAX, CsX) dissolving in polar aprotic solvents. The deposition of the solution on the substrate is generally assisted by a spin-coating process. The one-step deposition method for TiPes films usually suffers from rapid crystallization due to the higher Lewis acidity of Sn2+ than Pb2+, resulting in poor film quality with large amounts of pinholes and defects. Therefore, more factors need to be considered for TiPes film preparation. Hao et al. found that using dimethyl sulfoxide (DMSO) in the one-step deposition of MASnI3 films could create a SnI2-DMSO intermediate phase, promoting uniform nucleation and slowing down the film growth [Fig. 6(a)].138 However, DMSO’s low evaporation rate and high viscosity make it challenging to obtain high-quality perovskite films via spin coating. It is often used in combination with more volatile N,N-dimethylformamide (DMF) to improve the film quality.140,141 Besides, an anti-solvent, such as chlorobenzene (CB), toluene (TL), and diethyl ether (DE), is generally used to rapidly initiate nucleation and induce the formation of uniform, smooth, and dense perovskite films,142144 among which, CB is one of the most widely used anti-solvents in improving the film quality.140 Considering the high toxicity of CB, the use of green anti-solvent and the development of anti-solvent-free processes are widely appreciated. He et al. developed an innovative cooling procedure combining thermal convection and spontaneous nucleation to prepare inverted PSCs with a PCE of 8.9% without anti-solvent.61 Zhang et al. substituted CB with green diethyl carbonate (DEC) in a one-step preparation process to obtain FASnI3 thin films. By controlling crystal orientation and suppressing defects, the film quality was enhanced, resulting in solar cells with a champion PCE of 14.2%.145 Unlike the one-step method, sequential deposition typically involves two steps deposition of SnX2 and AX without the assistance of anti-solvents.146 However, the rapid reaction between the two precursors (e.g., SnI2 and FAI), coupled with the high solubility of SnI2 in the conventional solvent isopropanol, leads to poor film quality. Zhou et al. addressed this issue by using reducing formic acid as a solvent for the organic ammonium salts. This approach not only mitigated the erosion of SnI2 but also slowed down the crystallization rate through hydrogen bonding between formic acid and FAI, enabling the successful preparation of high-quality TiPes films. In addition, by leveraging the reducing properties of dithiothreitol, the PCE of the perovskite device was increased to 12.68%.147 Such progress in solvent and additive engineering provides promising opportunities for the broader adoption of the sequential deposition method.146,148150

    Fabrication of polycrystalline TiPes films. (a) Schematic diagram of one-step deposition method for MASnI3 film preparation. Reproduced with permission from Ref. 138. (b) Schematic diagram of MASnI3 film preparation based on ion exchange/insertion reaction. Reproduced with permission from Ref. 139.

    Figure 6.Fabrication of polycrystalline TiPes films. (a) Schematic diagram of one-step deposition method for MASnI3 film preparation. Reproduced with permission from Ref. 138. (b) Schematic diagram of MASnI3 film preparation based on ion exchange/insertion reaction. Reproduced with permission from Ref. 139.

    Vapor phase deposition complements the conventional spin coating process. Compared with the solution processing methods, the perovskite films obtained by vapor deposition usually possess smooth surfaces, high coverage, and good reproducibility independent of the operator.151,152 For example, Yokoyama et al. deposited MAI on SnI2 substrates using a low-temperature vapor-assisted solution process (LT-VASP) to obtain MASnI3 films with extremely high coverage and good uniformity; such films overcome the short-circuiting behavior seen in conventional solution methods.153 Alternatively, depositing SnI2 on the FAI substrate generates FASnI3 films, and the introduction of PEAI into this system further enhances overall film quality.152 Wang et al. deposited MAI on a substrate pre-spin-coated with SnF2/poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) and obtained MASnI3 thin films with the aid of an ion exchange/insertion method [Fig. 6(b)].139 The films were homogeneous and dense without pinholes, and a high PCE of 7.78% was obtained when applied to planar PSCs. The vapor phase deposition possesses a large degree of freedom in operation, allowing free choice between procedures such as simultaneous evaporation of precursors, spin coating followed by evaporation, and direct evaporation of perovskite. Therefore, currently, it has been widely used in the field of high-quality TiPes film preparation.154156

    Recently, scalable production processes such as blade coating, spraying coating, inkjet printing, screen printing, and electrodeposition have been found to be highly productive in controlling large-area film formation of perovskites.157,158 For example, the maximum PCE of lead PSCs obtained by spraying or scratch coating can reach 19.4% and 23.19%, respectively, demonstrating the potential for scalable production.159,160 The research progress of TiPes in related fields has been relatively slow due to the challenges in controlling the crystallization rate. Zuraw et al. pioneered the application of the blade coating technique to the preparation of TiPes thin films. By finely tuning the composition of the (BA0.5PEA0.5)2FA3Sn4I13 system, they successfully fabricated high-quality thin films and achieved a PCE of 5.7% on flexible modules with an area of up to 25  cm2.161 Similarly, the blade coating technique has also been applied to the scalable fabrication of tin-lead hybrid PSCs, further demonstrating the potential of TiPes for large-area production.162,163 However, there remains a significant performance gap in PCE between the blade-coating process and the spin-coating method. To further enhance the performance of large-area TiPes devices, greater emphasis should be placed on improving the film quality and homogeneity during the fabrication process.

    3.2 Single Crystals

    The synthesis processes of TiPes single crystals mainly include slow cooling crystallization (SCC) or inverse temperature crystallization (ITC).39,164 In SCC, the temperature is gradually decreased to trigger the supersaturation of the precursor solution, ultimately yielding perovskite single crystals. For example, as shown in Fig. 7(a), Han et al. synthesized (4-FPEA)2SnI4 single crystals in ethylene glycol solvent using a slow solution temperature lowering method. Compared with the single crystals in HI, the quality of the obtained single crystals was higher due to the ligand effect of ethylene glycol.165 To grow perovskite single crystals with larger sizes, a combination of top seeding methods and SCC can be used.167 Based on this method, Dang et al. successfully grew MASnI3 and FASnI3 single crystals with crystal sizes up to 20  mm×16  mm×10  mm and 8  mm×6  mm×5  mm, respectively, in HI-H3PO2 mixed solution.50 In ITC, millimeter-sized perovskite single crystals can be obtained by raising the temperature to supersaturate the precursor solution and allowing it to crystallize over an extended period. For instance, Bakr et al. used ITC to produce MASnBr3 single-crystal films with a surface area of up to 1  cm2, and the single-crystal thickness can be controlled from 5 to 10  μm by surface tension; such large-size single-crystal films are desirable for PD applications.168,169 Yuan et al. demonstrated the synthesis of MASnI3 single crystals with 3D dimensions in the micron range in a GBL solvent, leveraging spatial constraints. These single crystals exhibit sharp absorption edges and strong optical absorption, featuring an ideal optical band gap of 1.24 eV, which provides the opportunity to fabricate efficient single-junction tin PSCs.170 In addition, CsSnI3 single-crystal nanowires were fabricated through a slow evaporation of the solvent, resulting in perfect lattice structure, low carrier trap density, long carrier lifetime, and excellent carrier mobility. These characteristics position them as attractive candidates for applications in flexible optoelectronic devices.63

    Fabrication of single crystals, nanocrystals, and quantum dots. (a) Schematic diagram of (4-FPEA)2SnI4 single crystals prepared by slow solution temperature lowering method. Reproduced with permission from Ref. 165. (b) Schematic for the synthesis of CsSnI3 nanocrystal by the hot injection. (c) TEM images of CsSnI3 nanocrystal with different Cs:Sn ratios. Reproduced with permission from Ref. 166.

    Figure 7.Fabrication of single crystals, nanocrystals, and quantum dots. (a) Schematic diagram of (4-FPEA)2SnI4 single crystals prepared by slow solution temperature lowering method. Reproduced with permission from Ref. 165. (b) Schematic for the synthesis of CsSnI3 nanocrystal by the hot injection. (c) TEM images of CsSnI3 nanocrystal with different Cs:Sn ratios. Reproduced with permission from Ref. 166.

    3.3 Nanocrystals and Quantum Dots

    TiPes nanocrystals or quantum dots can be synthesized by hot injection and ligand-assisted reprecipitation (LARP), which are similar to lead perovskites.157 The hot injection method is a widely used approach for the preparation of TiPes nanocrystals/quantum dots.22,65,171 Long-chain organic ligands, such as aliphatic carboxylic acids and amines, are usually used, and the size and morphology can be regulated via controlling the ligand chain length and reaction temperature.157As shown in Fig. 7(b), cubic-phase CsSnI3 nanocrystals were synthesized via a hot-injection method, where a tri-n-octylphosphine (TOP) solution containing tin(II) 2-ethylhexanoate [Sn(Oct)2] and trimethylsilyl iodide (TMSI) was injected into a Cs2CO3-rich precursor solution. The nanocrystal size could be precisely tuned by adjusting the molar ratio of Sn to Cs precursors [Fig. 7(c)].166 This method not only produces regular and uniform cubic nanocrystals but also enables the formation of well-ordered nanoplate structures through further surface ligand modulation.172 In parallel, Li et al. conducted an in-depth investigation into the oxidation mechanism of Sn2+ and proposed a strategy combining TOP with metallic tin powder. This approach effectively mitigates tin oxidation driven by oxygen and solvents, enabling the preparation of monodisperse α-CsSnI3 films. As a result, the carrier lifetime is significantly extended by two orders of magnitude, reaching 278 ns.173 LARP is another effective strategy to synthesize perovskite nanocrystals, and the main principle is to use a polar solvent containing the respective ions and a nonpolar solvent for rapid mixing, and to exploit the large solubility difference of the ions in the two types of solvents to provide the crystallization driving force.157,174 LARP has been utilized for the preparation of lead counterparts of various morphologies and sizes, but there are only a few reports on the TiPes.62,64 Weidman et al. prepared BA2SnI4 and BA2[FASnI3]SnI4 (BA: butylamine) nanoplates by dropping TL into the solutions with BAI, CsI, and SnI2 in DMF.62 The thickness of such nanoplatelets can be controlled by changing the ratio of precursors. FASnI3 nanocrystals were successfully synthesized using LARP and the stability of the nanocrystals is improved by introducing SnF2.64 However, it is worth noting that, unlike the lead perovskites, the synthesis of TiPes nanocrystals and quantum dots has a higher requirement of inert-atmosphere protection to prevent oxidation.

    Similar to lead perovskite nanocrystals, TiPes nanocrystals can also be synthesized indirectly by leveraging ion exchange properties. Gahlot et al. demonstrated the successful preparation of 3D ASnX3 nanocrystals by utilizing 2D [R-NH3]2SnX4 and cationic oleates under diffusive conditions at the liquid-solid interface at room temperature. Similarly, anion exchange of [R-NH3]2SnX4 and ASnX3 can also be achieved using benzoyl halides as halogen sources.175 However, compared with lead perovskites, TiPes typically exhibit lower ion exchange efficiency. In addition, the instability of Sn2+ makes it prone to oxidation during the exchange process, which further accelerates crystal degradation. Consequently, the fabrication of high-quality TiPes nanocrystals still primarily relies on the direct synthesis methods described earlier.

    4 Applications

    Similar to lead perovskites, the extensive investigation of TiPes starts with a focus on the application in solar cells, where the TiPes were used to replace toxic lead perovskites as light absorbers and significant progress has been made in this area. With further exploration of the optoelectronic properties of TiPes, their applications have been gradually extended to LEDs, PDs, lasers, and other fields. In this section, we overview the progress of the application of TiPes in various fields, with special emphasis on the recent advancements of high-performance devices based on TiPes.

    4.1 Solar Cells

    Table 2 summarizes the device structures and photovoltaic performance parameters of representative tin and lead perovskites, providing a clear insight into the latest research advancements in tin PSCs. It is known that two device configurations are typically used for constructing PSCs, i.e., regular (n-i-p) and inverted (p-i-n) structures [Fig. 8(a)].192 In contrast to the balanced development of regular and inverted lead PSCs, high-efficiency tin PSCs are typically designed with p-i-n structures. This preference primarily arises from the challenges associated with the common n-i-p structure, where the electron transport layer (ETL) tends to oxidize the TiPes film and excessive p-type doping leads to imbalanced carrier transport.183 Li et al. attempted to replace the conventional metal-oxide electron transport layer (ETL) with the mixed chalcogenide compound Sn(S0.92Se0.08)2, achieving an improvement in the PCE of the n-i-p structure to 11.78% by suppressing the generation of Sn4+.177 However, this efficiency remains lower than the high PCE of over 15% achieved in the p-i-n structure. Currently, research on tin PSCs continues to focus primarily on the p-i-n structure.40 The PCE of p-i-n device configuration tin PSCs remains significantly lower than that of lead counterparts, particularly in terms of fill factor (FF) and open-circuit voltage (VOC). This performance gap arises from challenges such as rapid crystallization, Sn2+ oxidation, energy level mismatches in the transport layers, and inefficient carrier transport at the interfaces.195

    PerovskiteConfigurationJSC (mAcm2)VOC (V)FF (%)PCE (%)Ref.
    MASnI3ITO/PEDOT:PSS/PVK/PCBM/BCP/Ag20.680.57667.78139
    FA0.98EDA0.01SnI3ITO/EMIC-PEDOT:PSS/PVK/ICBA/C60/BCP/Ag23.860.7979.4514.98176
    PEA0.15FA0.85SnI3FTO/Sn(S0.92Se0.08)2/PVK/PTAA/Ag22.280.7372.6811.78177
    FA0.98EDA0.01SnI3ITO/NiOx/2PADBC/PVK/C60/BCP/Ag23.230.82574.0614.19178
    PEA0.15FA0.85SnI3FTO/MeO-2PACz + 6PA/PVK/ICBA/BCP/Ag17.60.82964.59.4179
    FA0.75MA0.25SnI3ITO/PEDOT:PSS/PVK/FACl/C60/BCP/Ag24.90.7776.714.7125
    (Cs0.02(FA0.9DEA0.1)0.98)0.98EDA0.01SnI3ITO/PEDOT:PSS/PVK/EDA/(PCBM + P3HT + ICBA)/C60/BCP/Ag24.720.8473.615.33180
    PEA0.15FA0.85SnI3ITO/PEDOT:PSS/perovskite/PCBM/BCP/Ag24.810.856 72.3715.38181
    FA0.85TEA0.15SnI3ITO/PEDOT:PSS/PVK/ICBA/BCP/Ag21.70.97474.115.723
    CsSnI3ITO/PEDOT:PSS/PVK/ICBA/BCP/Ag24.940.757413.68182
    FASnI3FTO/bl & mp-TiO2 /PVK/Spiro-OMeTAD:DPI-TPFB23.590.64971.2510.9183
    FAPbI3FTO/SnO2/PVK/Spiro-OMeTAD/Au25.69 1.17886.1526.08184
    MAPbI3ITO/PTAA/PVK/PCBM/BCP/Ag23.21.23580.7123.12185
    CsPbI3FTO/TiO2/PVK/Spiro-OMeTAD/Ag20.711.2683.821.86186
    CsPbI3xBrxFTO/TiO2/PVK/Spiro-OMeTAD/Au20.691.28383.6222.2187
    Cs0.05FA0.95PbI3TO/NiOx/Me-4PACz/PVK/C60/BCP/Ag26.181.19485.7626.81188
    Cs0.03FA0.945MA0.025Pb(I0.975Br0.025)3ITO/SnO2/PVK/OAI/Spiro-OMeTAD/Ag25.41.283.7925.54189
    Cs0.05MA0.05FA0.9PbI3ITO/2PACz + Me-4PACz/PVK/C60/SnO2/Cu26.51.1885.526.7122
    (3FBA)2MA3Pb4I13ITO/PVCz-ThOMeTPA /PVK/PCBM/Cr/Au22.421.228222.37190
    (BA0.75PFA0.25)2MA4Pb5I16ITO/MPA-CPA/PVK/PEACl/PCBM + PMMA/LiF/C60/BCP/Ag22.381.2675.221.15191

    Table 2. Summary of state-of-the-art tin and lead perovskite photovoltaic parameters.

    Application of TiPes in solar cells. (a) Schematic diagram of formal and inverted solar cell structure. Reproduced with permission from Ref. 192. (b) Schematic crystal structures of trans-2, trans-3, trans-4, and e. Reproduced with permission from Ref. 193. (c) Schematic illustration of the interface between F3-TMOS, NH2-TMOS anchored PEDOT:PSS and TiPes. (d) Current density voltage curves. (e) EQE spectra of tin PSCs with/without F3-TMOS. Reproduced with permission from Ref. 194. (f) Schematic diagram of FPEABr molecular dipole. (g) Ultraviolet photoelectron spectra (UPS) of FASnI3 films with and without FPEABr. (h) Schematic diagram of interface dipole and (i) energy level alignment. Reproduced with permission from Ref. 23.

    Figure 8.Application of TiPes in solar cells. (a) Schematic diagram of formal and inverted solar cell structure. Reproduced with permission from Ref. 192. (b) Schematic crystal structures of trans-2, trans-3, trans-4, and e. Reproduced with permission from Ref. 193. (c) Schematic illustration of the interface between F3-TMOS, NH2-TMOS anchored PEDOT:PSS and TiPes. (d) Current density voltage curves. (e) EQE spectra of tin PSCs with/without F3-TMOS. Reproduced with permission from Ref. 194. (f) Schematic diagram of FPEABr molecular dipole. (g) Ultraviolet photoelectron spectra (UPS) of FASnI3 films with and without FPEABr. (h) Schematic diagram of interface dipole and (i) energy level alignment. Reproduced with permission from Ref. 23.

    To address these issues, careful selection of transport layers is critical. For the hole transport layer (HTL), poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT:PSS) is commonly used due to its polar surface components, which strongly interact with perovskite precursors and promote crystallization. However, the hygroscopic nature of PEDOT:PSS can lead to the rapid degradation of TiPes. By contrast, inorganic NiOx HTLs offer better stability, but the presence of trivalent Ni can oxidize the TiPes, leading to poor PCE. Li et al. addressed this issue by introducing a self-assembled monolayer (SAM) of (4-(7H-dibenzo[c,g]carbazol-7-yl)ethyl)phosphonic acid (2PADBC) at the perovskite-NiOx interface.178 The phosphonic acid group in 2PADBC anchors to under-coordinated Ni cations, inhibiting their reaction with the perovskites and boosting device efficiency to 14.19%. SAM can be directly used as an HTL, achieving efficiencies exceeding 26% in lead PSCs.122 However, in tin PSCs, its highest reported efficiency remains below 10%, primarily due to uneven coverage of SAM on the conductive glass surface.179 For the ETL, PCBM and C60 are commonly used. However, due to their better energy level alignment with TiPes, new fullerene derivatives are increasingly attracting attention. Jiang et al. used indene-C60 bisadduct (ICBA) with higher CBM as ETL to obtain VOC up to 0.94 V and boost the PCE of tin PSCs to 12.4%.196 Sun et al. applied four regioisomers (trans-2, trans-3, trans-4, and e) of diethylmalonate-C60 bisadduct (DCBA) as ETL, among which, trans-3 exhibits the best energy level alignment and enhanced interfacial interactions with the TiPes layer, achieving a certified PCE of 14.3% [Fig. 8(b)].193

    Beyond transport layers, bulk phase engineering has been extensively explored, with Lewis bases emerging as particularly important additives in TiPes. Their significance lies in the ability to modulate the crystallization process. For example, Chen et al. demonstrated the effectiveness of the small-molecule Lewis base urea in enhancing the PCE of tin PSCs.197 DFT theoretical calculations further reveal that Lewis bases effectively passivate deep energy level traps in TiPes, which in turn suppresses non-radiative recombination and enhances carrier transport efficiency.198 Building on this, Zhou et al. doped TiPes with 2,8-dibromo-dibenzothiophene-S,S-dioxide (BrDS) molecules. By effectively modulating crystallization and passivating defects, BrDS significantly improved the FF from 73.91% to 79.45% and boosted the PCE from 11.9% to 14.98%.176 Chen et al. further advanced the role of Lewis base additives by synthesizing two pyridyl-substituted fullerene derivatives: cis-(CPPF) and trans-(TPPF).181 The spatial configurations significantly affect their electron density distribution, thereby influencing the interaction between the Lewis basic groups and TiPes. CPPF’s two nitrogen atoms can coordinate with one tin colloidal cluster, whereas TPPF can coordinate with two, leading to larger clusters in the case of TPPF. This reduces the number of nucleation sites, resulting in higher-quality TiPe films. The tin PSCs using TPPF achieve a PCE of 15.38% (certified 15.14%), representing a substantial improvement.

    Interfacial engineering is also essential for improving the performance of tin PSCs.199 In addition to the role of the buried interface in regulating crystallization and the upper interface in protecting the perovskite layer, its critical contribution to improving carrier extraction efficiency is particularly noteworthy.200,201 It is known that tin perovskite films usually suffer from lower carrier mobility and shorter carrier lifetime due to unfavorable film quality. Besides, the higher density of interfacial defects and poor energy level alignment bring more challenges.196 To address these challenges, Teng et al. developed silane coupling agents with varying molecular dipole moments by designing different terminal groups. Specifically, the electron-withdrawing −F in F3-TMOS creates a dipole moment directed toward the HTL, whereas the electron-donating NH2 group in NH2-TMOS generates an opposing dipole moment [Fig. 8(c)]. Notably, the incorporation of F3-TMOS as a bottom interface modifier significantly improves device performance. The modified HTL induces downward energy band bending, effectively optimizing the energy level alignment between the HTL and the perovskite layer. As a result, the device achieves a VOC of 0.91 V and a PCE of 14.67% [Figs. 8(d) and 8(e)].194 Similarly, interfacial engineering at the upper interface has demonstrated notable performance enhancements. For example, fluorophenethylamine hydrobromide (FPEABr) forms molecular dipoles on the TiPes surface while passivating surface defects [Figs. 8(f) and 8(h)]. This modification resulted in improved energy level alignment at the upper interface, leading to an increase in VOC to 0.974 V and achieving a PCE of 15.7% [Figs. 8(g) and 8(i)].23

    Stability is another key factor in advancing the practical applications of PSCs. Table 3 presents a summary of stability studies on tin PSCs and offers a comparative analysis with their lead counterparts. The stabilization strategies for TiPes primarily focus on preventing the rapid oxidation of Sn2+. For instance, Yang et al. developed a two-stage reduction strategy using S2O32, which not only reduces Sn4+ to Sn2+ in the precursor but also converts I2 to I during the film formation process. This approach significantly enhances device stability, enabling unencapsulated devices to retain 90% of their initial efficiency after 628 h of maximum power point (MPP) tracking in a N2 environment.204 Beyond oxygen, water-induced degradation poses another critical challenge for TiPes.

    PerovskiteConfigurationStrategyTest conditionsPCE (%)StabilityRef.
    FASnI3ITO/PEDOT:PSS/PVK/C60/BCP/AgAntioxidant, DipI + NaBH4MPP tracking, N2, unencapsulated10.611300 h, 96% PCE retained202
    CsSnI2.6Br0.4ITO/NiOx/PVK/PCBM/ZrAcac/AgAntioxidant, DMKOShelf stability, N2, 85°C, unencapsulated11.2720 h, 70% PCE retained203
    FASnI3(10%pFPEABr)FTO/PEDOT:PSS/PVSK/ICBA/BCP/AgAntioxidant, K2S2O3MPP tracking, N2, unencapsulated14.78628 h, 90% PCE retained204
    PEA0.15FA0.85SnI3ITO/PEDOT:PSS/PVK/PCBM/BCP/AgAntioxidant, TPPFMPP tracking, N2, unencapsulated15.38500 h, 93% PCE retained181
    FASnI3ITO/PEDOT:PSS/PVK/C60/BCP/AgComponent regulation, TEABrMPP tracking, N2, unencapsulated9.42000 h, 95% PCE retained205
    PEAxFA0.75MA0.25xSnI2BrITO/PEDOT:PSS/PVK/PCBM/BCP/AgComponent regulation, PEABrShelf stability, air, 25% to 30% RH, unencapsulated7.96300 h, 80% PCE retained206
    PEA0.15FA0.85SnI0.85Br0.15ITO/PEDOT:PSS/PVK/PCBM/BCP/AgCrystallization regulation, ureaShelf stability, N2, unencapsulated14.22200 h, 90% PCE retained197
    FA0.75MA0.25SnI2.75Br0.25ITO/PEDOT:PSS/PVK/CF3PEAI/PCBM/BCP/AgInterface passivation, CF3PEAIShelf stability, air, 15% RH, unencapsulated10.35150 h, 70% PCE retained207
    FA0.78GA0.2SnI3ITO/PEDOT:PSS/PVK/C60/BCP/Cu/BiElectrode covering, BiMPP tracking, air, 43°C, 30% to 70% RH, unencapsulated6.6100  h, 70% PCE retained208
    FASnI3ITO/PPr-SBT-14/PVK/C60/BCP/AgHTL regulationShelf stability, N2, unencapsulated7.66000 h, 100% PCE retained209
    FA0.85MA0.1Cs0.05PbI3FTOc-TiO2/PVK/MeO-PEAI/Spiro-OMeTAD/AuIon migration inhibition, 0.1% water in anti-solventMPP tracking, N2, 50°C, unencapsulated263500 h, 95% PCE retained210
    FAPbI3FTO/SnO2/PVK/PEAI/Spiro-OMeTAD/AgStrain regulation, Ph–Se–ClMPP tracking, day/night cycles, N2, unencapsulated26.343 day/night cycles, over 80% PCE retained211
    Cs0.05(FA0.91MA0.09)0.95Pb(I0.935Br0.05Cl0.015)3ITO/MeO-4PACz/PVK/PCBM/AgCrystallization regulation, GAI/NMPMPP tracking, N2, 40% to 50% RH, unencapsulated25.382160 h, 93.09% PCE retained121
    Cs0.05MA0.05FA0.9PbI3FTO/2PACz + Me-4PACz/PVK/C60/SnO2/CuInhibition of deprotonation, PDAI2/3MTPAIMPP tracking, air, 85°C, 50% RH, glass-encapsulated26.71130 h, 93.09% PCE retained122
    FAPbI3ITO/NiOx/MeO-2PACz/PVK/(anti,S)16,17-bis[60]PCBM/C60/SnO2/Cr/AuDefect passivation, NH4FMPP tracking, air, 40°C, 46% ± 15% RH, encapsulated24.83000 h, 97% PCE retained212
    FA0.95Cs0.05PbI3ITO/NiOx/PTAA/AlOx/PVK/PCBM@DCBP/BCP/AgAg diffusion inhibition, DCBPShelf stability, air, 85°C, 85% RH, encapsulated26.031500 h, 90% PCE retained213
    MA0.1Cs0.05FA0.85Pb(I0.95Br0.05)3ITO/(SnO2-PEIE)/(PCBM/MnSO4)/PVK/PDCBT/PTAA-BCF/AuHTL and ETL regulationMPP tracking, N2, 60°C to 65°C, unencapsulated20.91250 h, 99% PCE retained214

    Table 3. Comparison of stabilization strategies for tin and lead PSCs.

    Quasi-2D structures play a significant role in this field, as the capping layers they form can effectively block the intrusion of water and oxygen.120 To date, various cations have been employed to construct quasi-2D structures for TiPes, including PEA+, FPEA+, TEA+, DFPD+, and 3AMPY+.23,205,206,215,216 For example, a device prepared by partially substituting the A-site with PEA+ maintains 80% of its initial efficiency after 300 h in an air environment (25% to 30% RH), whereas the PCE of a conventional device drops rapidly to below 20% within just 5 h.206 Compared with TiPes, lead perovskites are less sensitive to oxygen, and their stabilization strategies primarily focus on preventing moisture intrusion, inhibiting the deprotonation of A-site ions, and addressing the widespread issue of ion migration. The first two strategies are highly relevant to TiPes; however, studies have shown that TiPes are less affected by ion migration.35 Stability testing for lead perovskites has already become increasingly standardized, but significant progress is still needed to establish comparable testing protocols for TiPes.

    In the field of conventional bandgap PSCs, lead perovskites remain dominant in terms of both efficiency and stability.211 TiPes, on the contrary, shows significant potential due to its low toxicity, making it suitable for indoor photovoltaic applications.217 Moreover, its tunable bandgap to a narrow value offers opportunities to extend the spectral response range of tandem solar cells.218 However, to fully harness the advantages of TiPes, substantial progress in efficiency is still needed.

    4.2 Light-Emitting Diodes

    The application of TiPes in LEDs began in 2016, initially targeting the NIR spectrum. Since then, it has expanded to red and blue light emission. Table 4 provides a comparison of the key performance parameters between state-of-the-art tin and lead perovskite LEDs. In terms of efficiency and luminance (or radiance), TiPes LEDs do not demonstrate significant advantages. However, TiPes LEDs show advantages in their tunable emission in the NIR-II region.83,229,230 Zhang et al. achieved high-quality TiPes films through a vapor-assisted spin-coating method. Solvent vapors were shown to lead to in situ recrystallization of TiPes during the film formation process, which significantly improves the crystal quality and reduces defects. By further adding the reducing agent phenylhydrazine hydrochloride (PH), the NIR LED devices based on ITO/PEDOT: PSS/perovskite/TmPyPB/LiF/Al structure can achieve a high EQE of 5.3% [Figs. 9(a)9(c)].231 This performance is mainly affected by the rapid aggregation of TiPes emitters during the preparation process, which would cause imperfect alignment of nanoparticles and formation of defects, thus leading to severe non-radiative recombination.

    ConfigurationLuminance (cd  m2)Radiance (Wsr1m2)FWHM (nm)EL (nm)EQE (%)StabilityRef.
    ITO/PEDOT:PSS/Poly-TPD/PEA2SnI4/TPBi/LiF/Al451236303.51T50=13.7  min219
    ITO/NiOx/TFB/(BrPMA)2SnBr4/TPBi/LiF/Al800704671.3T50=22  min220
    ITO/PEDOT:PSS/CsSnI3/B3PYMPM/LiF/Al9315.6T50=1475  min221
    ITO/m-PEDOT:PSS/CsSnI3/SPPO13/B3PYMPM/LiF/Al9457.6T50=82.6  h222
    ITO/PEDOT:PSS/CsSnI3/TPBi/LiF/Al226719482.63T50=39.5  h223
    ITO/PEDOT:PSS/PEAI-FA0.9Cs0.1SnI3/TPBi/LiF/Al128948.318
    ITO/PEAI-FA0.9Cs0.1SnI3/TPBi/LiF/Al8989811.6224
    ITO/PEDOT:PSS/TEA2SnI4/TPBi/LiF/Al24.963020.29T50=27.6  h24
    ITO/TFB/PVK/PEABr-CsPbBr3/BPBiPA/LiF/Al11,3702451632.1T50=3.56  h225
    ITO/PEDOT:PSS:PFI/PF8Cz/CsPb(Brx/Cl1x)3/TPBi/PO-T2T/LiF/Al48026.4226
    ITO/PEDOT:PSS/TFB/LiF/PEA2Csn1Pbn(Br/I)3n+1/TPBi/LiF/Al27064829.04T50=43.7  min227
    ITO/PEIE-ZnO /FAPbI3-5AVA-PyNI/TFB/MoOx/Au39080532T50=19  h228

    Table 4. Summary of key parameters of state-of-the-art tin and lead perovskite LEDs.

    Application of TiPes in light-emitting diodes. (a) Schematic of the LED device structure. (b) Cross-sectional SEM image of LED device (scale bar: 100 nm). (c) Current density and radiance-dependent EQE of TiPes LED. Reproduced with permission from Ref. 231. (d) Crystal structure image of PEA2SnI4 perovskite and electrostatic potential image of GSH. (e) Photos of air aging (40% RH, 25°C) of PEA2SnI4 perovskite with and without GSH addition. Reproduced with permission from Ref. 232. (f) EL spectra of CsSnI3 LED with moisture treatment. (g) Statistical results and (h) operation lifetime test results of EQE with and without moisture treatment for CsSnI3 LEDs. Reproduced with permission from Ref. 222.

    Figure 9.Application of TiPes in light-emitting diodes. (a) Schematic of the LED device structure. (b) Cross-sectional SEM image of LED device (scale bar: 100 nm). (c) Current density and radiance-dependent EQE of TiPes LED. Reproduced with permission from Ref. 231. (d) Crystal structure image of PEA2SnI4 perovskite and electrostatic potential image of GSH. (e) Photos of air aging (40% RH, 25°C) of PEA2SnI4 perovskite with and without GSH addition. Reproduced with permission from Ref. 232. (f) EL spectra of CsSnI3 LED with moisture treatment. (g) Statistical results and (h) operation lifetime test results of EQE with and without moisture treatment for CsSnI3 LEDs. Reproduced with permission from Ref. 222.

    The formation of quasi-2D materials has been shown to effectively passivate defects and reduce non-radiative recombination. In addition, they can self-assemble into multi-quantum well structures and provide energy transfer channels, enhancing the light emission efficiency of LEDs. Min et al. compared the in situ PL images of 3D FA0.9Cs0.1SnI3 before and after the introduction of PEAI and found that the incorporation of PEAI significantly increased the lifetime of the 3D TiPes emitter, confirming the suppression of non-radiative recombination by the quasi-2D structure.18 In addition, the introduction of vitamin B1 (VmB1) can further benefit the PL and the LEDs fabricated using PEAI-VmB1-based FA0.9Cs0.1SnI3 exhibited an emission peak at 894  nm and achieved an impressive EQE of up to 8.3%. Quasi-2D structures are widely and favorably used in TiPes LEDs. It is crucial to note that the crystallization of perovskites with low-dimensional phases requires precise tuning, as varying n-values in these phases can lead to disordered heterogeneous interfaces and hinder efficient energy transfer. Wang and colleagues discovered that the strong interaction between SnF2 and PEAI inhibits the formation of 2D perovskites. However, upon introducing tryptophan, a stronger interaction occurs between tryptophan and SnF2, leading to the partial release of PEA+ ions. This facilitates the formation of 2D/3D mixed-dimensional bilayer perovskites during sequential annealing. As a result, the corresponding TiPes LEDs achieve a high EQE of 11.6%.224

    Compared to lead perovskite LEDs, TiPes LEDs also demonstrate superior red color purity. For instance, an early study on the PEA2SnI4 system reported a red TiPes LED with chromaticity coordinates of (0.706, 0.294) on the Commission Internationale de L’Eclairage (CIE) chromaticity coordinate chart, which are remarkably close to the ideal red emitter coordinates of (0.708, 0.292). However, the device suffered from a low EQE of only 0.3% and a luminance of 70  cdm2, with this underperformance primarily attributed to the inherent instability of Sn2+.229 Bai et al. utilized L-glutathione reduced (GSH) as a surface ligand to inhibit the oxidation of Sn2+ and passivated the defects to construct high-quality PEA2SnI4 perovskite films [Fig. 9(d)]. The additive not only led to a significant increase in device EQE to 9.32% but also enhanced the stability. As shown in Fig. 9(e), the target film still emits bright red light under UV light after air aging for 90 min.232 Recent findings by Han et al. demonstrated that poly-functional cyanuric acid (CA) can establish strong chemical interactions on the surface of TEA2SnI4, inducing an electron localization effect that generates stable dimer/trimer superstructures.24 This superstructure may form strong chemical coordination with Sn2+, stabilizing Sn2+ as well as improving the crystal ordering of TEA2SnI4. Notably, the champion EQE of the target device reached 20.29%, rendering it on par with state-of-the-art lead counterparts.233,234

    Due to the tunability of the bandgap, TiPes LEDs can theoretically be applied to green and blue light emission, similar to lead perovskite LEDs. Han et al. achieved blue light emission in quasi-2D TiPes by modulating cations, which enhanced lattice rigidity and reduced the electron-phonon coupling effect.220 A blue LED based on (BrPMA)2SnBr4 demonstrated an EQE of 1.3% and a maximum brightness of 800  cdm2. However, this remains the only report in the field to date, and its EQE of 1.3% falls far short of the 26.4% EQE achieved by state-of-the-art lead perovskite LEDs.226 This performance gap is primarily attributed to the inferior crystal quality and higher defect density of Br-containing tin-based systems, which significantly reduce luminous efficiency. To address these challenges, further research is needed to optimize the crystalline quality of TiPes and minimize defect concentrations, thereby overcoming existing technological barriers and improving device performance.

    In terms of stability research, the gap between TiPes LEDs and lead perovskite LEDs is gradually diminishing, particularly in studies focused on the NIR region. Yuan et al. reported highly stable, bright inorganic TiPes LEDs by incorporating N-phenylthiourea (NPTU) and SnF2.223 They proposed that, prior to Auger recombination becoming dominant at higher doping concentrations, the presence of free background holes from p-type doping could enhance the PLQY and promote radiative recombination. By skillfully leveraging the crystallization control of NPTU and SnF2, they managed to regulate p-doping density and reduce trap density, thereby boosting both radiative efficiency and stability. The modified CsSnI3 achieved a radiance of 226  Wsr1m2 at >2800  mAcm2 and a record T50 of 39.5 h at 100  mAcm2 for NIR LEDs with emission >900  nm.235,236 Guan et al. developed an innovative moisture-triggered SnX4 hydrolysis strategy to eliminate Sn4+-induced defects in CsSnI3 thin films, effectively reducing background carrier density while providing persistent protection through the formation of Sn(OH)4.222 The moisture-treated CsSnI3 exhibited stable electroluminescence (EL) at 945 nm [Fig. 9(f)], achieving a peak EQE of 7.6% with excellent reproducibility [Fig. 9(g)]. Notably, this treatment significantly extended the operational lifetime of CsSnI3 by 3.7-fold, reaching 82.6 h under a constant current density of 20  mAcm2 [Fig. 9(h)].

    In the LED field, lead perovskites have achieved significant breakthroughs in EQE, reaching 26.4%, 32.1%, and 29.04% for blue, green, and red light emissions, respectively. In the NIR range (805 nm), lead perovskites have also recently achieved a high EQE of 32%.225228 By contrast, current research on TiPes primarily focuses on red and NIR emissions, with EQE values still lagging behind those of lead perovskites. However, TiPes hold promise for surpassing lead perovskites in the NIR region, particularly in the >900  nm range, an area where pure lead perovskites are less effective and difficult to apply.223 Furthermore, the inherent low ion migration probability of TiPes promises higher stability of LEDs compared with lead perovskites, which is a critical advantage for practical applications.

    4.3 Photodetectors

    In halide perovskites, applying a constant voltage results in the accumulation of halide ions at the interface, which can help reduce the barrier for collecting photogenerated carriers [Fig. 10(a)]. This will not only contribute to enhancing the photocurrent but, at the same time, will also lead to increased dark current due to ion shielding effects. Besides, the perovskite photosensitive layer is a polycrystalline film with carriers moving along different pathways. Enhancing vertical carrier transport by reducing grain size promotes increased charge recombination at defect sites, which suppresses dark current. However, this comes at the expense of photocurrent. Achieving optimal photodetection performance necessitates striking a balance between increasing photocurrent and decreasing dark current.27 Compared with lead perovskites, TiPes have a stronger binding affinity with halogen ions and lower ion mobility. In addition, TiPes have high carrier mobility, which is beneficial for the transport of photo-induced charge carriers. These inherent advantages of TiPes make them suitable candidates for PD applications.14,239

    Application of TiPes in photodetectors. (a) Schematic diagram of the effect of halogen ion migration on the barrier. Reproduced with permission from Ref. 27. (b) Schematic diagram of fabrication of CsSnI3 PDs. Reproduced with permission from Ref. 237. (c) Responsivity of the flexible FASnI3 PDs with and without CNI treatment. (d) Schematic diagram of photoplethysmography test. (e) Heart rate results of rigid (blue) and flexible (red) FASnI3 PDs with CNI treatment at 0 V and 200 μW cm−2 light intensity. Reproduced with permission from Ref. 238.

    Figure 10.Application of TiPes in photodetectors. (a) Schematic diagram of the effect of halogen ion migration on the barrier. Reproduced with permission from Ref. 27. (b) Schematic diagram of fabrication of CsSnI3 PDs. Reproduced with permission from Ref. 237. (c) Responsivity of the flexible FASnI3 PDs with and without CNI treatment. (d) Schematic diagram of photoplethysmography test. (e) Heart rate results of rigid (blue) and flexible (red) FASnI3 PDs with CNI treatment at 0 V and 200  μWcm2 light intensity. Reproduced with permission from Ref. 238.

    Han et al. reported the application of all-inorganic TiPes in NIR PDs.240 The CsSnI3 nanowire arrays were deposited by solid-source CVD on mica substrates with a narrow band gap of 1.34 eV. Assembly of the CsSnI3 nanowire arrays into an NIR PD achieved a responsivity (R) of 54  mAW1 and a detection rate (D*) of 3.85×105  Jones with response times of 83.8/243.4  ms (rise/decay). The formation of the oxidized phase is a crucial factor constraining PDs’ performance, and most subsequent studies focus on preventing the generation of the oxidized phase. For example, Cao et al. introduced ascorbic acid as a reductant in CsSnI3 to suppress the oxidation, with the device preparation process shown in Fig. 10(b).237 The resultant PDs exhibited enhanced responsivity and stability over a wide spectral range of 350 to 1000 nm; the optimal R was 0.257  AW1, and the response time was 0.35/1.6  ms.

    Structures in 2D and quasi-2D have exhibited enhanced moisture resistance, leading to improved stability and detection performance. Yang et al. investigated the photoelectric response of 2D BA2SnI4+5%SnF2 and quasi-2D BA2FASn2I7+10%SnF2 under 405-nm optical radiation.241 These structures achieved high on/off ratios of 458 and 1108, respectively, at an irradiation intensity of 9  mWcm2. Notably, the long-term performance of 2D and quasi-2D PDs remained stable even under extended irradiation. At lower light intensities, these 2D/quasi-2D PDs exhibited remarkable R and D* values of 2654/1557  mAW1 and 1.46×1013/6.23×1012  Jones, respectively. The ultra-high stability, detection capability, and on/off ratio of low-dimensional structures are difficult to achieve with conventional TiPes PDs. However, disordered quantum well structures may affect carrier mobility, thereby impacting the detector’s fast response. Therefore, the application of low-dimensional materials should be based on more precise control of layer numbers and crystallization.

    Despite notable progress, the sensitivity and response speed of TiPes in the conventional wavelength range (300 to 800 nm) still lag significantly behind those of lead perovskites, primarily due to their inherent instability. However, the band gap limitation of lead perovskites makes NIR detection challenging, creating opportunities for TiPes in imaging and medical applications.242 For example, self-powered PDs based on quasi-single-crystal FASnI3 thin films exhibit high detection rates in the NIR range (780 to 890 nm) exceeding 1013  Jones.243 In addition, TiPes possess low-temperature processing capabilities, offering significant advantages in the field of flexible PDs. Liu et al. demonstrated the potential of TiPes in flexible PDs. They found that FASnI3 flexible PDs treated with 2-cyanoethan-1-aminium iodide (CNI) exhibited excellent responsivity in the NIR region, achieving 370  mAW1 at 785 nm [Fig. 10(c)]. In addition, the PDs showed high detectivity (9.12×1012  Jones at 785 nm) and fast response times (3.91  μs). These photodetectors are well-suited for wearable electronic devices, enabling real-time monitoring of human heart rate under low-light and zero-bias conditions [Figs. 10(d) and 10(e)].238 Therefore, with advancements in material design, defect passivation, and device optimization, TiPes are poised to play a crucial role in NIR and flexible detection.

    4.4 Lasers

    TiPes are also utilized for fabricating tunable lasers. Xing et al. concluded that lasing originates from free electron-hole bimolecular recombination, and CsSnI3 (SnF2 treatment) has an extremely high recombination rate for lasing applications. Ultra-stable NIR coherent emission (700 to 1000 nm) can be achieved at room temperature through halogen modulation. This wavelength range, however, is challenging to achieve with lead perovskites.104 Chirvony et al. found the NIR random laser behavior in FASnI3 polycrystalline films. A high-quality factor (104) with a low amplified spontaneous emission threshold (2  μJcm2) was obtained at 20 K.244 Interestingly, the random lasing (RL) mode generated in FASnI3 polycrystalline films exhibits exceptional spectral stability; as the excitation flux increases, the spectral position of the narrow PL lines remains unchanged [Figs. 11(a) and 11(b)]. This stands in stark contrast to the disordered variations seen in the RL spectra of lead perovskites. This suggests that TiPes are promising for practical applications with narrow laser lines and high spectral stability.

    Application in lasing. (a) RL spectra and (b) normalized RL spectra of FASnI3 thin film at different excitation fluences (20 K, excitation by 532 nm pulses). Reproduced with permission from Ref. 244. (c) Temperature-dependent PL spectra of PEA2SnI4 at an excitation fluence of 2 mJ cm−2 and (d) FWHM of the ASE peak. Reproduced with permission from Ref. 245. (e) Schematic presentation of BA+ substitution by conjugated monomers. (f) Lasing spectra for (PEA)2SnI4 at 129 μJ cm−2, (2T)2SnI4 at 210 μJ cm−2, and (3T)2SnI4 at 252 μJ cm−2 [Inset is a lasing image of (3T)2SnI4]. (g) Temperature-dependent PL spectra of (3T)2MASn2I7 at different excitation fluences. Reproduced with permission from Ref. 25.

    Figure 11.Application in lasing. (a) RL spectra and (b) normalized RL spectra of FASnI3 thin film at different excitation fluences (20 K, excitation by 532 nm pulses). Reproduced with permission from Ref. 244. (c) Temperature-dependent PL spectra of PEA2SnI4 at an excitation fluence of 2  mJcm2 and (d) FWHM of the ASE peak. Reproduced with permission from Ref. 245. (e) Schematic presentation of BA+ substitution by conjugated monomers. (f) Lasing spectra for (PEA)2SnI4 at 129  μJcm2, (2T)2SnI4 at 210  μJcm2, and (3T)2SnI4 at 252  μJcm2 [Inset is a lasing image of (3T)2SnI4]. (g) Temperature-dependent PL spectra of (3T)2MASn2I7 at different excitation fluences. Reproduced with permission from Ref. 25.

    2D perovskites exhibit higher radiative recombination efficiencies, making them highly promising candidates for laser applications. A series of 2D lead perovskites (n>1), such as (BA)2MAn1PbnI3n+1, has demonstrated intrinsic laser emission in the visible region without requiring an external cavity.246 However, due to the limitations of high Auger recombination rates and strong exciton-phonon coupling in 2D TiPes, their laser design requires a more flexible and meticulous approach. Petrozza and his colleagues successfully synthesized PEA2SnI4 through precise crystal engineering, achieving amplified spontaneous emission (ASE) at an exceptionally low threshold with an optical gain exceeding 4000  cm1 at 77 K.245 They observed significant thermal sensitivity in PEA2SnI4 and found that higher temperatures resulted in reduced ASE intensity and broadened FWHM [Figs. 11(c) and 11(d)]. Even at room temperature, albeit with a weaker signal, ASE was still detectable, underscoring the potential utility of room temperature lasers based on 2D TiPes using PEA2SnI4. Li et al. employed a mixed-solvent strategy to synthesize a series of RP-phase TiPes. By skillfully tuning the quantum well thickness and organic ligand design, they achieved modulation of broad-spectrum optoelectronic properties.25 It has been shown that substituting BA+ with conjugated ligands (PEA+, 2T+, 3T+) enables optically pumped lasing in 2D TiPes [Fig. 11(e)]. By contrast, RP-phase TiPes can achieve optically pumped lasing at n=1, whereas lead perovskites require n>1. This further demonstrates the superiority of TiPes [Fig. 11(f)]. Furthermore, the temperature-dependent lasing properties of monolayer TiPe, i.e., (3T)2MASn2I7, have been studied, which clearly demonstrate that the temperature has a huge impact on the modes and threshold of lasing [Fig. 11(g)].

    TiPes offer significant advantages for achieving room-temperature NIR coherent emission, with notable breakthroughs enabled by low-dimensional and nanomaterials.25,246 Despite significant progress, TiPes continue to face challenges in achieving high phase purity and precise control over shape and size, especially when compared to lead perovskites. Furthermore, the relationship between the crystal structure, carrier dynamics, and lasing performance of TiPes requires further in-depth investigation.

    4.5 Transistors

    TiPes are promising candidates for realizing high-performance transistors. Their intrinsically high hole mobility offers a significant opportunity to drive the development of high-performance p-channel (hole) field-effect transistors (FETs) and complementary metal-oxide-semiconductors (CMOSs).247 Shao et al. first demonstrated the feasibility of employing FASnI3 as a semiconductor channel in FETs. FETs utilizing FASnI3 exhibit exceptional hole mobility (0.21  cm2V1s1), impressive on and off current ratio (Ion/Ioff, 104), and relatively low threshold voltages (VTH, 2.8 V).248 Zhu et al. successfully enhanced the quality of MASnI3 thin films and improved the device performance of thin-film transistors (TFTs) through halogen ion modulation.26 The introduction of a small quantity of Br ions could enhance coordination with Sn2+, effectively governing the nucleation and crystallization kinetics of TiPes. This led to improved crystallinity and orientation while reducing defects such as pin-holes. The introduction of Cl ions promoted the formation of TiPes, resulting in a smoother film surface. The effective synergistic combination of these two halogens led to a significantly enhanced performance of TFTs, achieving a high hole mobility of 20  cm2V1s1, an Ion/Ioff ratio as high as 107, and the elimination of a bias voltage to deactivate the transistor. It is important to clarify that the hysteresis phenomenon, primarily caused by ion migration, is a major concern in lead perovskites but is dramatically suppressed in TiPes. The hysteresis of TiPes TFTs, on the other hand, is mainly related to minority carrier traps, i.e., I vacancies (VI). However, in composite halide TiPes, the presence of Br/Cl has a passivating effect on VI, resulting in negligible hysteresis in TFTs [Figs. 12(a) and 12(b)].

    Application of TiPes in transistors. (a) TFTs structure. (b) Transfer characteristics of TFTs with different halogen-substituted TiPes. Reproduced with permission from Ref. 26. (c) Inverter structure and optical image. (d) Butterfly voltage transfer characteristic (VDD=12.5 V); noise margin (NM) in blue color. (e) Gain [=(d Vout)/(d Vin)]. Reproduced with permission from Ref. 249.

    Figure 12.Application of TiPes in transistors. (a) TFTs structure. (b) Transfer characteristics of TFTs with different halogen-substituted TiPes. Reproduced with permission from Ref. 26. (c) Inverter structure and optical image. (d) Butterfly voltage transfer characteristic (VDD=12.5  V); noise margin (NM) in blue color. (e) Gain [=(dVout)/(dVin)]. Reproduced with permission from Ref. 249.

    The stability of perovskite can also be improved by incorporating a 2D structure for better application in transistors, for example, the introduction of the diammonium cation 4AMP+ and 3AMP+ dramatically improves the stability of TiPes in air.115 In addition, the 2D layer is capable of passivating defects, and rational crystallization modulation can enhance the charge carrier mobility of TiPes. Zhu et al. developed p-type TFTs using TiPes with a hybrid cation engineering approach involving Cs, FA, and PEA. The gradual introduction of PEA and Cs ions had a beneficial effect on the crystallization process of the FA film, resulting in reduced surface roughness and improved film quality.249 The ternary (CsxFA1x)PEA2Sn8I25 film exhibited a low density of charged defects, achieving a remarkable field-effect mobility up to 70  cm2V1s1, and demonstrated an Ion/Ioff ratio of 108. These achievements represent leading performance in perovskite TFTs. By employing cation engineering, the presence of iodide interstitials/vacancies in TiPes was reduced, leading to diminished dual-sweep hysteresis. In addition, the combination of (CsxFA1x)PEA2Sn8I25 TFTs with n-type In2O3 enabled the creation of complementary semiconductor inverters [as depicted in Fig. 12(c)]. The inverter displays typical rail-to-rail voltage transition characteristics, featuring a sharp transition when the input voltage reaches VDD/2, signifying a logic state shift from “1” to “0” [Fig. 12(d)]. With a VDD of 12.5 V, the inverter exhibits an impressive gain of 370, surpassing the performance of the lead-based integrated inverter [Fig. 12(e)]. In the realm of fully inorganic TiPes, exceptional field-effect hole mobility exceeding 50  cm2V1s1 and an Ion/Ioff ratio surpassing 108 have been realized.250

    Due to p-type doping in TiPes, it is well-suited for applications in p-channel field-effect transistors. In addition, when combined with n-channel metal-oxide transistors, perovskite can be used to create CMOS electronics, offering enhanced noise resistance and low power consumption.247 However, their performance remains difficult to match that of n-type FETs due to the challenges of high p-type doping, whereas n-type transistors can be readily achieved using lead perovskite.251

    4.6 Other Applications

    Apart from the abovementioned applications, TiPes, known for their exceptional semiconductor properties, have also seen a gradual expansion of application in other areas. In the study conducted by Qian et al., an ITO/MASnBr3/Au device structure was assembled to evaluate its memory capacitance characteristics, as depicted in Fig. 13(a).252 When subjected to periodic bias voltage sweeping, the device exhibits significant hysteresis loops in the capacitance–voltage (CV) curve, whereas the charge–voltage (QV) curve displays two distinctive pinched loops [illustrated in Fig. 13(b)]. These observations validate the device’s suitability as an ideal memory capacitor. Besides, the oxygen sensitivity of TiPes offers potential for oxygen sensing applications. Cai et al. leveraged this trait and created an innovative oxygen sensor by pairing 2D PEA2SnI4 with a tilted fiber Bragg grating. This sensor exhibits high sensitivity and rapid response [Figs. 13(c) and 13(d)], showcasing TiPes’ potential for other sensor applications.253 Moreover, TiPes, benefiting from their excellent Seebeck coefficient and ultra-low thermal conductivity, have achieved an optimal thermoelectric power factor of 342  μWm1K2 in CsSnI3 thin films, highlighting their promising potential in thermoelectric applications.254

    Applications of TiPes in memory capacitor and sensor. (a) Schematic diagram of ITO/MASnBr3/Au device structure. (b) Typical C–V and Q–V loops detected at 1 MHz. Reproduced with permission from Ref. 252. (c) Sensitivity of modified PEA2SnI4 to various gases (gas concentration of 0.5%). (d) Response of the TiPes sensor to oxygen in a wet or dry atmosphere. Reproduced with permission from Ref. 253.

    Figure 13.Applications of TiPes in memory capacitor and sensor. (a) Schematic diagram of ITO/MASnBr3/Au device structure. (b) Typical CV and QV loops detected at 1 MHz. Reproduced with permission from Ref. 252. (c) Sensitivity of modified PEA2SnI4 to various gases (gas concentration of 0.5%). (d) Response of the TiPes sensor to oxygen in a wet or dry atmosphere. Reproduced with permission from Ref. 253.

    5 Conclusion and Outlook

    With their outstanding optical and electrical properties, TiPes have become very promising lead-free semiconductors and significant progress has been achieved in their applications such as solar cells, LEDs, PDs, and lasers.23,243,255 However, the instability of TiPes remains the greatest obstacle to their replacement of lead perovskites. This instability is primarily caused by external factors such as oxygen, moisture, and temperature, which accelerate degradation. Various strategies have been proposed to enhance the stability of TiPes, spanning from precursor protection to perovskite optimization and device fabrication. Drawing on the insights into the stability of lead perovskites, we believe the most promising development strategies to improve the operational stability of TiPes can be identified. (i) Component regulation: efficient and stable lead perovskites are commonly achieved through the combination of FA+ and Cs+ ions.121,122 However, this approach is limited in the case of TiPes due to the sensitivity of CsSnI3 to oxygen and moisture.123 We believe that adjusting the ratio of these two ions with 2D-phase organic ammonium ions is desirable, as it can effectively isolate moisture and oxygen while simultaneously improving the thermal stability of the devices. (ii) Interfacial engineering: the interface of TiPes is typically the region where defects concentrate and degradation initiates. Therefore, the formation of a surface passivation layer is crucial to enhancing the stability of TiPes. In this regard, lead perovskites provide valuable insights, particularly in terms of passivating uncoordinated divalent metal ions and forming protective layers to block water and oxygen. Both systems share these common characteristics. However, the implementation of this approach must take into account the unique properties of TiPes, such as the coordination activity of Sn2+ and the corrosive effect of the commonly used solvent isopropanol (IPA). Feasible strategies include utilizing the soft Lewis acid characteristics of TiPes to select S- or N-containing groups for coordination. In addition, it may be beneficial to introduce groups that can coordinate with the active sites (−OH) of IPA or use mixed solvents to reduce the polarity-induced corrosion of IPA on the TiPes.

    Although some issues remain unresolved, considering the unique features of TiPes, it still shows great potential for making breakthroughs in the following areas in the future.

    1. 1.Extraction and utilization of hot carriers

    In TiPes, longer hot carrier lifetimes than those of lead counterparts have been reported. These hot carriers are present under both pulsed excitation and continuous wave excitation, confirming the possibility that hot carriers can be extracted for solar cell applications.21,76 Currently, hot carrier-based extraction applications are facing a dual challenge: the standard solar irradiation energy makes it difficult to meet the output demand of hot carriers, and in addition, the relaxation time of hot carriers needs to be further slowed down to enhance the probability of their extraction. To address these challenges, several potential solutions have been proposed: (i) achieving an appropriate perovskite grain size for slow relaxation of hot carriers at low injected carrier densities, (ii) designing device structures that maximize the use of sunlight, and (iii) selecting effective passivation agents to eliminate deep energy level defects that can directly trap hot carriers.22,256

    1. 2.Applications in narrow bandgap field

    The application of TiPes in narrow band gap areas provides an important opportunity to break the monopoly of lead perovskites. These areas include (i) solar cells: narrow bandgap TiPes can broaden the spectral absorption range of conventional solar cells and thus have the potential to be used in all-perovskite tandem solar cells. (ii) NIR emission and detection: due to the limitations of the bandgap, lead perovskites are unable to achieve effective light emission or detection beyond 850 nm.28 The addition of narrow-bandgap TiPes fills this gap. In the field of light emission, CsSnI3-based LED devices can extend the emission wavelength to over 940 nm. In addition, compared with lead perovskites, TiPes exhibit lower Auger recombination, meaning they can maintain ultra-low efficiency roll-off at high current densities.223 In the field of light detection, TiPes can extend the detection range to 1000 nm, achieving a detectivity of over 1012  Jones and microsecond-level fast response. Although TiPes show potential for application in these fields, their relatively low PCE, as well as light emission and detection efficiencies, still limits their practical use. One feasible solution is to use tin-lead mixed perovskites.242,257,258 From the perspective of TiPes, adding a small amount of lead can improve device efficiency while maintaining a low bandgap. From the perspective of lead perovskites, this approach can reduce the use of toxic lead and effectively narrow the bandgap of lead perovskites.

    1. 3.Applications in low toxicity field

    Due to the low toxicity of TiPes, they present promising opportunities and breakthroughs in various fields, particularly in applications that are integrated into daily life, where concerns about toxicity are more significant. Specifically, (i) indoor photovoltaics: common indoor light sources emit light in the 400 to 700 nm range, and the optimal bandgap for indoor photovoltaics is between 1.7 and 1.9 eV. TiPes can achieve this bandgap range through halide composition adjustments. Moreover, the halide ion migration in TiPes under light is minimal, which suggests that their application in indoor photovoltaics could offer improved stability. Devices based on FASnI2Br perovskites have already achieved a PCE of 20.12% under indoor lighting (3000 K, 1000 lux).217 (ii) Wearable flexible electronics: TiPes offer a distinct advantage over lead perovskites in low-temperature film formation, making them more suitable for fabrication on flexible substrates. Studies have been reported on flexible TiPes optoelectronic devices, such as flexible solar cells and photodetection.161,238 These are expected to boost the application of TiPes in wearable self-powered health detection devices. It is important to note that the above application scenarios are all based on low-light conditions, making it essential to achieve high light utilization efficiency in such environments. Combining light-capturing technologies, such as metamaterial nanostructures and gratings, to further enhance light absorption efficiency could be crucial to fully harness the excellent optical properties of TiPes.

    Feng Yang received his MS degree from Nanjing University of Aeronautics and Astronautics (NUAA) and is now advancing his research as a PhD candidate under the supervision of Prof. Yu Tong at Northwestern Polytechnical University. His research focuses on tin perovskite solar cells.

    Yu Tong is currently a professor at School of Materials Science and Engineering, Northwestern Polytechnical University. He received his PhD from Ludwig-Maximilians-University Munich in 2018 under the supervision of Prof. Jochen Feldmann. His recent research interest is novel semiconductor optoelectronic functional materials and devices.

    Kun Wang is an associate professor at School of Microelectronics, Northwestern Polytechnical University. She received her PhD from Technical University of Munich in 2019 under the supervision of Prof. Peter Müller-Buschbaum. Her research primarily focuses on optoelectronic conversion materials and devices.

    Hongqiang Wang is a professor at the School of Materials Science and Engineering, Northwestern Polytechnical University. Before then, he sequentially worked in National Institute of Advanced Industrial Science and Technology, Japan (postdoc), Max Planck Institute of Colloids and Interfaces (Alexander von Humboldt Fellow), and University of Liverpool (Marie Curie Fellow). His research focuses on advanced energy and catalytic materials, encompassing both fundamental and applied studies.

    Biographies of the other authors are not available.

    References

    [1] B. R. Sutherland, E. H. Sargent. Perovskite photonic sources. Nat. Photonics, 10, 295-302(2016). https://doi.org/10.1038/nphoton.2016.62

    [2] M. Ahmadi, T. Wu, B. Hu. A review on organic-inorganic halide perovskite photodetectors: device engineering and fundamental physics. Adv. Mater., 29, 1605242(2017). https://doi.org/10.1002/adma.201605242

    [3] T. Li et al. Ambient air processed inverted inorganic perovskite solar cells with over 21% efficiency enabled by multifunctional ethacridine lactate. Angew. Chem. Int. Edit., 63, e202407508(2024). https://doi.org/10.1002/anie.202407508

    [4] T. Li et al. In situ dehydration condensation of self-assembled molecules enables stabilization of CsPbI3 perovskites for efficient photovoltaics. Adv. Funct. Mater., 34, 2409621(2024). https://doi.org/10.1002/adfm.202409621

    [5] Y. Zhao et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. Science, 377, 531-534(2022). https://doi.org/10.1126/science.abp8873

    [6] S. Zhang et al. Minimizing buried interfacial defects for efficient inverted perovskite solar cells. Science, 380, 404-409(2023). https://doi.org/10.1126/science.adg3755

    [7] S. Liu et al. Buried interface molecular hybrid for inverted perovskite solar cells. Nature, 632, 536-542(2024). https://doi.org/10.1038/s41586-024-07723-3

    [8] J. Peng et al. Crystallization of CsPbBr3 single crystals in water for X-ray detection. Nat. Commun., 12, 1531(2021). https://doi.org/10.1038/s41467-021-21805-0

    [9] C. Wang et al. Low-threshold blue quasi-2D perovskite laser through domain distribution control. Nano Lett., 22, 1338-1344(2022). https://doi.org/10.1021/acs.nanolett.1c04666

    [10] H. Guo et al. Embedding laser generated GaAs nanocrystals in perovskite wires for enhanced charge transport and photodetection. Sci. China Phys. Mech., 65, 274204(2022). https://doi.org/10.1007/s11433-021-1827-x

    [11] H. Qi et al. Homogenizing energy landscape for efficient and spectrally stable blue perovskite light-emitting diodes. Adv. Mater., 36, 2409319(2024). https://doi.org/10.1002/adma.202409319

    [12] G. Nazir et al. Stabilization of perovskite solar cells: recent developments and future perspectives. Adv. Mater., 34, e2204380(2022). https://doi.org/10.1002/adma.202204380

    [13] M. M. Byranvand et al. Tin-based halide perovskite materials: properties and applications. Chem. Sci., 13, 6766-6781(2022). https://doi.org/10.1039/D2SC01914K

    [14] J. Li et al. Review on recent progress of lead-free halide perovskites in optoelectronic applications. Nano Energy, 80, 105526(2021). https://doi.org/10.1016/j.nanoen.2020.105526

    [15] Y. S. Gao et al. Recent advances in lead-safe perovskite solar cells. Adv. Funct. Mater., 32, 2208225(2022). https://doi.org/10.1002/adfm.202208225

    [16] E. W.-G. Diau, E. Jokar, M. Rameez. Strategies to improve performance and stability for tin-based perovskite solar cells. ACS Energy Lett., 4, 1930-1937(2019). https://doi.org/10.1021/acsenergylett.9b01179

    [17] X. Li et al. Lead-free halide perovskites for light emission: recent advances and perspectives. Adv. Sci., 8, 2003334(2021). https://doi.org/10.1002/advs.202003334

    [18] H. Min et al. Additive treatment yields high-performance lead-free perovskite light-emitting diodes. Nat. Photonics, 17, 755-760(2023). https://doi.org/10.1038/s41566-023-01231-y

    [19] M. Qin et al. Regulating the crystallization kinetics and lattice strain of lead-free perovskites with perovskite quantum dots. ACS Energy Lett., 7, 3251-3259(2022). https://doi.org/10.1021/acsenergylett.2c01692

    [20] S. Kahmann, S. Shao, M. A. Loi. Cooling, scattering, and recombination—the role of the material quality for the physics of tin halide perovskites. Adv. Funct. Mater., 29, 1902963(2019). https://doi.org/10.1002/adfm.201902963

    [21] H. H. Fang et al. Long-lived hot-carrier light emission and large blue shift in formamidinium tin triiodide perovskites. Nat. Commun., 9, 243(2018). https://doi.org/10.1038/s41467-017-02684-w

    [22] L. Dai et al. Slow carrier relaxation in tin-based perovskite nanocrystals. Nat. Photonics, 15, 696-702(2021). https://doi.org/10.1038/s41566-021-00847-2

    [23] Y. Shi et al. Interfacial dipoles boost open-circuit voltage of tin halide perovskite solar cells. ACS Energy Lett., 9, 1895-1897(2024). https://doi.org/10.1021/acsenergylett.4c00529

    [24] D. Han et al. Tautomeric mixture coordination enables efficient lead-free perovskite LEDs. Nature, 622, 493-498(2023). https://doi.org/10.1038/s41586-023-06514-6

    [25] Y. Li et al. Phase-pure 2D tin halide perovskite thin flakes for stable lasing. Sci. Adv., 9, eadh0517(2023). https://doi.org/10.1126/sciadv.adh0517

    [26] H. Zhu et al. High-performance hysteresis-free perovskite transistors through anion engineering. Nat. Commun., 13, 1741(2022). https://doi.org/10.1038/s41467-022-29434-x

    [27] W. Jang et al. Prevention of noise current generation in tin‐based lead‐free perovskites for highly sensitive photodetection. Adv. Funct. Mater., 32, 2207713(2022). https://doi.org/10.1002/adfm.202207713

    [28] H. Wang et al. Tin-halide perovskites for near-infrared light-emitting diodes. ACS Energy Lett., 9, 2500-2507(2024). https://doi.org/10.1021/acsenergylett.4c00470

    [29] L. Chen et al. On the durability of tin-containing perovskite solar cells. Adv. Sci., 11, 2304811(2024). https://doi.org/10.1002/advs.202304811

    [30] Z. K. Le et al. Ion migration in tin-halide perovskites. ACS Energy Lett., 9, 1639-1644(2024). https://doi.org/10.1021/acsenergylett.4c00198

    [31] Z. Zhang et al. Mechanistic understanding of oxidation of tin-based perovskite solar cells and mitigation strategies. Angew. Chem. Int. Ed., 62, e202308093(2023). https://doi.org/10.1002/anie.202308093

    [32] J. Wang et al. Methods for passivating defects of perovskite for inverted perovskite solar cells and modules. Adv. Energy Mater., 14, 2401414(2024). https://doi.org/10.1002/aenm.202401414

    [33] S. Kahmann et al. Negative thermal quenching in FASnI3 perovskite single crystals and thin films. ACS Energy Lett., 5, 2512-2519(2020). https://doi.org/10.1021/acsenergylett.0c01166

    [34] X. Lu et al. Enhanced structural stability and photo responsiveness of CH3NH3SnI3 perovskite via pressure-induced amorphization and recrystallization. Adv. Mater., 28, 8663-8668(2016). https://doi.org/10.1002/adma.201600771

    [35] K. Dey et al. Substitution of lead with tin suppresses ionic transport in halide perovskite optoelectronics. Energy Environ. Sci., 17, 760-769(2024). https://doi.org/10.1039/D3EE03772J

    [36] T. J. Macdonald et al. Engineering stable lead-free tin halide perovskite solar cells: lessons from materials chemistry. Adv. Mater., 35, e2206684(2023). https://doi.org/10.1002/adma.202206684

    [37] Y. Jiang, X. Wang, A. Pan. Properties of excitons and photogenerated charge carriers in metal halide perovskites. Adv. Mater., 31, e1806671(2019). https://doi.org/10.1002/adma.201806671

    [38] M. Konstantakou, T. Stergiopoulos. A critical review on tin halide perovskite solar cells. J. Mater. Chem. A, 5, 11518-11549(2017). https://doi.org/10.1039/C7TA00929A

    [39] Y. Cho, H. R. Jung, W. Jo. Halide perovskite single crystals: growth, characterization, and stability for optoelectronic applications. Nanoscale, 14, 9248-9277(2022). https://doi.org/10.1039/D2NR00513A

    [40] T. Wu et al. Lead-free tin perovskite solar cells. Joule, 5, 863-886(2021). https://doi.org/10.1016/j.joule.2021.03.001

    [41] M. Singh, J. P. Akash, J. P. Tiwari. Solar cells based on Pb-free and perovskite-inspired materials: futuristic of perovskite solar cells. ACS Appl. Energy Mater., 7, 10212-10229(2024). https://doi.org/10.1021/acsaem.4c02047

    [42] B. A. Al-Asbahi et al. Effect of deposition method on the structural and optical properties of CH3NH3PbI3 perovskite thin films. Opt. Mater., 103, 109836(2020). https://doi.org/10.1016/j.optmat.2020.109836

    [43] L. M. Herz. Charge-carrier mobilities in metal halide perovskites: fundamental mechanisms and limits. ACS Energy Lett., 2, 1539-1548(2017). https://doi.org/10.1021/acsenergylett.7b00276

    [44] S. D. Stranks et al. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science, 342, 341-344(2013). https://doi.org/10.1126/science.1243982

    [45] S. Sen et al. Tin-based eco-friendly perovskites for sustainable future. Adv. Energy Sust. Res., 4, 2300110(2023). https://doi.org/10.1002/aesr.202300110

    [46] M. Morana et al. Cubic or not cubic? Combined experimental and computational investigation of the short-range order of tin halide perovskites. J. Phys. Chem. Lett., 14, 2178-2186(2023). https://doi.org/10.1021/acs.jpclett.3c00105

    [47] F. H. Naqvi, J.-H. Ko. Lattice dynamics and phase transitions in FAPbBr3 single crystals: temperature- and pressure-dependent Raman spectroscopy. J. Raman Spectrosc., 54, 1138-1149(2023). https://doi.org/10.1002/jrs.6589

    [48] Y. Fu et al. Metal halide perovskite nanostructures for optoelectronic applications and the study of physical properties. Nat. Rev. Mater., 4, 169-188(2019). https://doi.org/10.1038/s41578-019-0080-9

    [49] Y. Zhou et al. Room temperature “one-pot” solution synthesis of nanoscale CsSnI3 orthorhombic perovskite thin films and particles. Mater. Lett., 110, 127-129(2013). https://doi.org/10.1016/j.matlet.2013.08.011

    [50] Y. Dang et al. Formation of hybrid perovskite tin iodide single crystals by top-seeded solution growth. Angew. Chem. Int. Ed., 55, 3447-3450(2016). https://doi.org/10.1002/anie.201511792

    [51] Z. Zhu et al. Metal halide perovskites: stability and sensing-ability. J. Mater. Chem. C, 6, 10121-10137(2018). https://doi.org/10.1039/C8TC03164A

    [52] J. Pascual et al. Fluoride chemistry in tin halide perovskites. Angew. Chem. Int. Ed., 60, 21583-21591(2023). https://doi.org/10.1002/anie.202107599

    [53] Z. Fan, K. Sun, J. Wang. Perovskites for photovoltaics: a combined review of organic–inorganic halide perovskites and ferroelectric oxide perovskites. J. Mater. Chem. A, 3, 18809-18828(2015). https://doi.org/10.1039/C5TA04235F

    [54] D. Ricciarelli et al. Instability of tin iodide perovskites: bulk p-doping versus surface tin oxidation. ACS Energy Lett., 5, 2787-2795(2020). https://doi.org/10.1021/acsenergylett.0c01174

    [55] K. Yamada et al. Phase transition and electric conductivity of ASnCl3 (A = Cs and CH3NH3). Bull. Chem. Soc. Jpn., 71, 127-134(1998). https://doi.org/10.1246/bcsj.71.127

    [56] C. C. Stoumpos, C. D. Malliakas, M. G. Kanatzidis. Semiconducting tin and lead iodide perovskites with organic cations: phase transitions, high mobilities, and near-infrared photoluminescent properties. Inorg. Chem., 52, 9019-9038(2013). https://doi.org/10.1021/ic401215x

    [57] E. C. Schueller et al. Crystal structure evolution and notable thermal expansion in hybrid perovskites formamidinium tin iodide and formamidinium lead bromide. Inorg. Chem., 57, 695-701(2018). https://doi.org/10.1021/acs.inorgchem.7b02576

    [58] L. Wang et al. Pressure-induced structural evolution, optical and electronic transitions of nontoxic organometal halide perovskite-based methylammonium tin chloride. Appl. Phys. Lett., 111, 233901(2017). https://doi.org/10.1063/1.5004186

    [59] F. Zheng et al. Material innovation in advancing organometal halide perovskite functionality. J. Phys. Chem. Lett., 6, 4862-4872(2015). https://doi.org/10.1021/acs.jpclett.5b01830

    [60] X. K. Liu et al. Metal halide perovskites for light-emitting diodes. Nat. Mater., 20, 10-21(2021). https://doi.org/10.1038/s41563-020-0784-7

    [61] L. He et al. Efficient anti-solvent-free spin-coated and printed Sn-perovskite solar cells with crystal-based precursor solutions. Matter, 2, 167-180(2020). https://doi.org/10.1016/j.matt.2019.10.006

    [62] M. C. Weidman et al. Highly tunable colloidal perovskite nanoplatelets through variable cation, metal, and halide composition. ACS Nano, 10, 7830-7839(2016). https://doi.org/10.1021/acsnano.6b03496

    [63] L. Dai et al. Single-crystal nanowire cesium tin triiodide perovskite solar cell. Small, 19, e2208062(2023). https://doi.org/10.1002/smll.202208062

    [64] Z. Chen, T. P. Dhakal. Room temperature synthesis of lead-free FASnI3 perovskite nanocrystals with improved stability by SnF2 additive. Appl. Phys. Rev., 10, 011404(2023). https://doi.org/10.1063/5.0125100

    [65] Q. Fan et al. Lead-free halide perovskite nanocrystals: crystal structures, synthesis, stabilities, and optical properties. Angew. Chem. Int. Ed., 59, 1030-1046(2020). https://doi.org/10.1002/anie.201904862

    [66] K. Gahlot et al. Structural dynamics and tunability for colloidal tin halide perovskite nanostructures. Adv. Mater., 34, e2201353(2022). https://doi.org/10.1002/adma.202201353

    [67] A. G. Ricciardulli et al. Emerging perovskite monolayers. Nat. Mater., 20, 1325-1336(2021). https://doi.org/10.1038/s41563-021-01029-9

    [68] F. Li et al. Effects of alkyl chain length on crystal growth and oxidation process of two-dimensional tin halide perovskites. ACS Energy Lett., 5, 1422-1429(2020). https://doi.org/10.1021/acsenergylett.0c00286

    [69] Y. Cheng et al. Reduced-dimensional perovskites: quantum well thickness distribution and optoelectronic properties. Adv. Mater., 2410633(2024). https://doi.org/10.1002/adma.202410633

    [70] J. Shi et al. From ultrafast to ultraslow: charge-carrier dynamics of perovskite solar cells. Joule, 2, 879-901(2018). https://doi.org/10.1016/j.joule.2018.04.010

    [71] R. L. Milot et al. The effects of doping density and temperature on the optoelectronic properties of formamidinium tin triiodide thin films. Adv. Mater., 30, e1804506(2018). https://doi.org/10.1002/adma.201804506

    [72] D. Luo et al. Minimizing non-radiative recombination losses in perovskite solar cells. Nat. Rev. Mater., 5, 44-60(2019). https://doi.org/10.1038/s41578-019-0151-y

    [73] J. Fu et al. Hot carrier cooling mechanisms in halide perovskites. Nat. Commun., 8, 1300(2017). https://doi.org/10.1038/s41467-017-01360-3

    [74] H. Zhang et al. Highly mobile hot holes in Cs2AgBiBr6 double perovskite. Sci. Adv., 7, eabj9066(2021). https://doi.org/10.1126/sciadv.abj9066

    [75] M. B. Price et al. Hot-carrier cooling and photoinduced refractive index changes in organic-inorganic lead halide perovskites. Nat. Commun., 6, 8420(2015). https://doi.org/10.1038/ncomms9420

    [76] L. Dai, J. Ye, N. C. Greenham. Thermalization and relaxation mediated by phonon management in tin-lead perovskites. Light Sci. Appl., 12, 208(2023). https://doi.org/10.1038/s41377-023-01236-w

    [77] M. Bokdam et al. Role of polar phonons in the photo excited state of metal halide perovskites. Sci. Rep., 6, 28618(2016). https://doi.org/10.1038/srep28618

    [78] L. Ma et al. Carrier diffusion lengths of over 500 nm in lead-free perovskite CH3NH3SnI3 films. J. Am. Chem. Soc., 138, 14750-14755(2016). https://doi.org/10.1021/jacs.6b09257

    [79] A. Abate. Perovskite solar cells go lead free. Joule, 1, 659-664(2017). https://doi.org/10.1016/j.joule.2017.09.007

    [80] N. K. Noel et al. Lead-free organic–inorganic tin halide perovskites for photovoltaic applications. Energy Environ. Sci., 7, 3061-3068(2014). https://doi.org/10.1039/c4ee01076k

    [81] H. Zhu et al. Low-dimensional Sn-based perovskites: evolution and future prospects of solar cells. Chem, 8, 2939-2960(2022). https://doi.org/10.1016/j.chempr.2022.07.027

    [82] K. Xu et al. The 3D-structure-mediated growth of zero-dimensional Cs4SnX6 nanocrystals. Nanoscale, 14, 2248-2255(2022). https://doi.org/10.1039/d1nr06497e

    [83] Y. Liao et al. Two-dimensional tin perovskite nanoplate for pure red light-emitting diodes. J. Phys. D: Appl. Phys., 53, 414005(2020). https://doi.org/10.1088/1361-6463/ab9673

    [84] J. T. Lin et al. Harnessing dielectric confinement on tin perovskites to achieve emission quantum yield up to 21. J. Am. Chem. Soc., 141, 10324-10330(2019). https://doi.org/10.1021/jacs.9b03148

    [85] S. Gupta, D. Cahen, G. Hodes. How SnF2 impacts the material properties of lead-free tin perovskites. J. Phys. Chem. C, 122, 13926-13936(2018). https://doi.org/10.1021/acs.jpcc.8b01045

    [86] J. Zillner et al. The role of SnF2 additive on interface formation in all lead‐free FASnI3 perovskite solar cells. Adv. Funct. Mater., 32, 2109649(2022). https://doi.org/10.1002/adfm.202109649

    [87] A. Treglia et al. Effect of electronic doping and traps on carrier dynamics in tin halide perovskites. Mater. Horiz., 9, 1763-1773(2022). https://doi.org/10.1039/d2mh00008c

    [88] H. Xu et al. Orientation regulation of tin‐based reduced‐dimensional perovskites for highly efficient and stable photovoltaics. Adv. Funct. Mater., 29, 1807696(2019). https://doi.org/10.1002/adfm.201807696

    [89] F. Li et al. A cation-exchange approach for the fabrication of efficient methylammonium tin iodide perovskite solar cells. Angew. Chem. Int. Ed., 58, 6688-6692(2019). https://doi.org/10.1002/anie.201902418

    [90] C.-M. Tsai et al. Control of crystal structures and optical properties with hybrid formamidinium and 2-hydroxyethylammonium cations for mesoscopic carbon-electrode tin-based perovskite solar cells. ACS Energy Lett., 3, 2077-2085(2018). https://doi.org/10.1021/acsenergylett.8b01046

    [91] T. H. Chowdhury et al. Sn-based perovskite halides for electronic devices. Adv. Sci., 9, e2203749(2022). https://doi.org/10.1002/advs.202203749

    [92] P. Umari, E. Mosconi, F. De Angelis. Relativistic GW calculations on CH3NH3PbI3 and CH3NH3SnI3 perovskites for solar cell applications. Sci. Rep., 4, 4467(2014). https://doi.org/10.1038/srep04467

    [93] S. Tao et al. Absolute energy level positions in tin- and lead-based halide perovskites. Nat. Commun., 10, 2560(2019). https://doi.org/10.1038/s41467-019-10468-7

    [94] T. C. Jellicoe et al. Synthesis and optical properties of lead-free cesium tin halide perovskite nanocrystals. J. Am. Chem. Soc., 138, 2941-2944(2016). https://doi.org/10.1021/jacs.5b13470

    [95] I. Spanopoulos et al. Unraveling the chemical nature of the 3D ‘hollow’ hybrid halide perovskites. J. Am. Chem. Soc., 140, 5728-5742(2018). https://doi.org/10.1021/jacs.8b01034

    [96] H. Yao et al. Strategies for improving the stability of tin-based perovskite (ASnX3) solar cells. Adv. Sci., 7, 1903540(2020). https://doi.org/10.1002/advs.201903540

    [97] A. Fraccarollo et al. First principles study of 2D layered organohalide tin perovskites. J. Chem. Phys., 146, 234703(2017). https://doi.org/10.1063/1.4985054

    [98] U. G. Jong et al. First-principles study on the material properties of the inorganic perovskite Rb1xCsxPbI3 for solar cell applications. Phys. Rev. B, 98, 125116(2018). https://doi.org/10.1103/PhysRevB.98.125116

    [99] L. Leppert, T. Range, J. B. Neaton. Towards predictive band gaps for halide perovskites: lessons from one-shot and eigenvalue self-consistent GW. Phys. Rev. Mater., 3, 103803(2019). https://doi.org/10.1103/PhysRevMaterials.3.103803

    [100] D. Umadevi, G. W. Watson. Quasiparticle GW calculations on lead-free hybrid germanium iodide perovskite CH3NH3GeI3 for photovoltaic applications. ACS Omega, 4, 5661-5669(2019). https://doi.org/10.1021/acsomega.8b03291

    [101] S. X. Tao, X. Cao, P. A. Bobbert. Accurate and efficient band gap predictions of metal halide perovskites using the DFT-1/2 method: GW accuracy with DFT expense. Sci. Rep., 7, 14386(2017). https://doi.org/10.1038/s41598-017-14435-4

    [102] A. K. Jena, A. Kulkarni, T. Miyasaka. Halide perovskite photovoltaics: background, status, and future prospects. Chem. Rev., 119, 3036-3103(2019). https://doi.org/10.1021/acs.chemrev.8b00539

    [103] R. Prasanna et al. Band gap tuning via lattice contraction and octahedral tilting in perovskite materials for photovoltaics. J. Am. Chem. Soc., 139, 11117-11124(2017). https://doi.org/10.1021/jacs.7b04981

    [104] G. Xing et al. Solution-processed tin-based perovskite for near-infrared lasing. Adv. Mater., 28, 8191-8196(2016). https://doi.org/10.1002/adma.201601418

    [105] H. Liu et al. Realizing high-detectivity near-infrared photodetectors in tin–lead perovskites by double-sided surface-preferred distribution of multifunctional tin thiocyanate additive. ACS Energy Lett., 8, 577-589(2022). https://doi.org/10.1021/acsenergylett.2c02055

    [106] A. M. Smith, S. M. Nie. Semiconductor nanocrystals: structure, properties, and band gap engineering. Acc. Chem. Res., 43, 190-200(2010). https://doi.org/10.1021/ar9001069

    [107] A. Wang et al. Aqueous acid-based synthesis of lead-free tin halide perovskites with near-unity photoluminescence quantum efficiency. Chem. Sci., 10, 4573-4579(2019). https://doi.org/10.1039/c9sc00453j

    [108] Z. Fang et al. Dual passivation of perovskite defects for light-emitting diodes with external quantum efficiency exceeding 20%. Adv. Funct. Mater., 30, 1909754(2020). https://doi.org/10.1002/adfm.201909754

    [109] F. Yuan et al. Color-pure red light-emitting diodes based on two-dimensional lead-free perovskites. Sci. Adv., 6, eabb0253(2020). https://doi.org/10.1126/sciadv.abb0253

    [110] Y. Ke et al. Efficient and bright deep-red light-emitting diodes based on a lateral 0D/3D perovskite heterostructure. Adv. Mater., 36, e2207301(2022). https://doi.org/10.1002/adma.202207301

    [111] N. Li et al. Diammonium-mediated perovskite film formation for high-luminescence red perovskite light-emitting diodes. Adv. Mater., 34, e2202042(2022). https://doi.org/10.1002/adma.202202042

    [112] A. Abate. Stable tin-based perovskite solar cells. ACS Energy Lett., 8, 1896-1899(2023). https://doi.org/10.1021/acsenergylett.3c00282

    [113] J. Liu et al. Origins and suppression of Sn(II)/Sn(IV) oxidation in tin halide perovskite solar cells. Adv. Energy Mater., 13, 2300696(2023). https://doi.org/10.1002/aenm.202300696

    [114] A. Balilonda et al. High performance and efficiency recovery in Sn-based flexible perovskite solar cells by antioxidant vanillin and microwave treatment. Adv. Funct. Mater., 34, 2313833(2024). https://doi.org/10.1002/adfm.202313833

    [115] W. Park et al. Two-dimensional Dion-Jacobson tin perovskite transistors with enhanced ambient stability. ACS Energy Lett., 9, 2436-2445(2024). https://doi.org/10.1021/acsenergylett.4c00620

    [116] K. Gahlot et al. Air-stable thin films of tin halide perovskite nanocrystals by polymers and Al2O3 encapsulation. Chem. Mater., 36, 11227-11235(2024). https://doi.org/10.1021/acs.chemmater.4c02261

    [117] E. Aktas et al. One-step solution deposition of tin-perovskite onto a self-assembled monolayer with a DMSO-free solvent system. ACS Energy Lett., 8, 5170-5174(2023). https://doi.org/10.1021/acsenergylett.3c02098

    [118] Z. Zhang et al. Coordination engineering with crown ethers for perovskite precursor stabilization and defect passivation. Energy Environ. Sci., 17, 7182-7192(2024). https://doi.org/10.1039/d4ee02124j

    [119] K.-W. Yeom, D.-K. Lee, N.-G. Park. Hard and soft acid and base (HSAB) engineering for efficient and stable Sn-Pb perovskite solar cells. Adv. Energy Mater., 12, 2202496(2022). https://doi.org/10.1002/aenm.202202496

    [120] T. Hossain et al. Effects of A-site cation structure on the stability of 2D tin halide perovskites. Chem. Mater., 36, 11004-11014(2024). https://doi.org/10.1021/acs.chemmater.4c01251

    [121] H. Dong et al. Regulating the perovskite crystallization dynamics via dual modulation strategy for performance enhancement of perovskite solar cells. Adv. Energy Mater., 2403965(2024). https://doi.org/10.1002/aenm.202403965

    [122] Y. Yang et al. Amidination of ligands for chemical and field-effect passivation stabilizes perovskite solar cells. Science, 386, 898-902(2024). https://doi.org/10.1126/science.adr2091

    [123] Y. Liu et al. Near-infrared light emitting metal halides: materials, mechanisms, and applications. Adv. Mater., 36, 2312482(2024). https://doi.org/10.1002/adma.202312482

    [124] L. Wang et al. 14.31% power conversion efficiency of Sn-based perovskite solar cells via efficient reduction of Sn4+. Angew. Chem. Int. Ed., 62, e202307228(2023). https://doi.org/10.1002/anie.202307228

    [125] J. Zhou et al. Chemo-thermal surface dedoping for high-performance tin perovskite solar cells. Matter, 5, 683-693(2022). https://doi.org/10.1016/j.matt.2021.12.013

    [126] J. Cao et al. Interactions between H2O and lead halide perovskites: recent progress and applications. Matter, 7, 3728-3755(2024). https://doi.org/10.1016/j.matt.2024.08.001

    [127] C. Sun et al. Efficient and stable white LEDs with silica-coated inorganic perovskite quantum dots. Adv. Mater., 28, 10088-10094(2016). https://doi.org/10.1002/adma.201603081

    [128] H. Zhang et al. Embedding perovskite nanocrystals into a polymer matrix for tunable luminescence probes in cell imaging. Adv. Funct. Mater., 27, 1604382(2017). https://doi.org/10.1002/adfm.201604382

    [129] M. Ahlawat et al. Design principle of a water-dispersed photocatalytic perovskite through ligand deconstruction. ACS Energy Lett., 8, 2159-2168(2023). https://doi.org/10.1021/acsenergylett.3c00532

    [130] P. Aggarwal et al. Energy funneling from water-dispersed perovskites to chromophores. ACS Energy Lett., 8, 1520-1528(2023). https://doi.org/10.1021/acsenergylett.3c00197

    [131] P. Aggarwal et al. Enhanced stability and energy transfer in perovskite nanocrystals: paving the way for moisture-resistant light-harvesting devices. ACS Appl. Nano Mater., 6, 21616-21625(2023). https://doi.org/10.1021/acsanm.3c03600

    [132] M. Ahlawat et al. Harnessing Pb–S interactions for long-term water stability in cesium lead halide perovskite nanocrystals. Small, 20, 2401326(2024). https://doi.org/10.1002/smll.202401326

    [133] A. Babayigit et al. Assessing the toxicity of Pb- and Sn-based perovskite solar cells in model organism Danio rerio. Sci. Rep., 6, 18721(2016). https://doi.org/10.1038/srep18721

    [134] L. Lanzetta et al. Degradation mechanism of hybrid tin-based perovskite solar cells and the critical role of tin(IV) iodide. Nat. Commun., 12, 2853(2021). https://doi.org/10.1038/s41467-021-22864-z

    [135] J. Cao, F. Yan. Recent progress in tin-based perovskite solar cells. Energy Environ. Sci., 14, 1286-1325(2021). https://doi.org/10.1039/d0ee04007j

    [136] J. Li et al. Biological impact of lead from halide perovskites reveals the risk of introducing a safe threshold. Nat. Commun., 11, 310(2020). https://doi.org/10.1038/s41467-019-13910-y

    [137] K. J. Prince et al. Sustainability pathways for perovskite photovoltaics. Nat. Mater., 24, 22-33(2024). https://doi.org/10.1038/s41563-024-01945-6

    [138] F. Hao et al. Solvent-mediated crystallization of CH3NH3SnI3 films for heterojunction depleted perovskite solar cells. J. Am. Chem. Soc., 137, 11445-11452(2015). https://doi.org/10.1021/jacs.5b06658

    [139] P. Wang et al. Ion exchange/insertion reactions for fabrication of efficient methylammonium tin iodide perovskite solar cells. Adv. Sci., 7, 1903047(2020). https://doi.org/10.1002/advs.201903047

    [140] X. Liu et al. Solvent engineering improves efficiency of lead-free tin-based hybrid perovskite solar cells beyond 9%. ACS Energy Lett., 3, 2701-2707(2018). https://doi.org/10.1021/acsenergylett.8b01588

    [141] B. P. Nguyen et al. Phase formation and local charge transport of lead-free CH3NH3Sn(I1xBrx)3 (0x1) perovskite solar cells fabricated by solvent optimization. Sol. Energy, 186, 136-144(2019). https://doi.org/10.1016/j.solener.2019.05.007

    [142] X. Liu et al. Efficient and stable tin perovskite solar cells enabled by amorphous-polycrystalline structure. Nat. Commun., 11, 2678(2020). https://doi.org/10.1038/s41467-020-16561-6

    [143] Z. Kang et al. Homogenizing the low-dimensional phases for stable 2D–3D tin perovskite solar cells. Small, 20, 2402028(2024). https://doi.org/10.1002/smll.202402028

    [144] W. Liao et al. Lead-free inverted planar formamidinium tin triiodide perovskite solar cells achieving power conversion efficiencies up to 6.22%. Adv. Mater., 28, 9333-9340(2016). https://doi.org/10.1002/adma.201602992

    [145] Z. Zhang et al. Green-antisolvent-regulated distribution of p-type self-doping enables tin perovskite solar cells with efficiency over 14%. Energy Environ. Sci., 16, 3430-3440(2023). https://doi.org/10.1039/D3EE00601H

    [146] S. Shahbazi et al. Realizing a cosolvent system for stable tin-based perovskite solar cells using a two-step deposition approach. ACS Energy Lett., 5, 2508-2511(2020). https://doi.org/10.1021/acsenergylett.0c01190

    [147] H. Zhou et al. Reaction-diffusion and crystallization kinetics modulation of two-step deposited tin-based perovskite film via reducing atmosphere. Angew. Chem. Int. Ed., 64, e202422217(2024). https://doi.org/10.1002/anie.202422217

    [148] Z. Zhu et al. Realizing efficient lead-free formamidinium tin triiodide perovskite solar cells via a sequential deposition route. Adv. Mater., 30, 1703800(2018). https://doi.org/10.1002/adma.201703800

    [149] D. Song et al. Interfacial engineering with a hole-selective self-assembled monolayer for tin perovskite solar cells via a two-step fabrication. ACS Energy Lett., 6, 4179-4186(2021). https://doi.org/10.1021/acsenergylett.1c02124

    [150] X. Liu et al. Lead-free perovskite solar cells with over 10% efficiency and size 1 cm² enabled by solvent–crystallization regulation in a two-step deposition method. ACS Energy Lett., 7, 425-431(2021). https://doi.org/10.1021/acsenergylett.1c02651

    [151] M. Zhu, G. Cao, Z. Zhou. Recent progress toward highly efficient tin‐based perovskite (ASnX3) solar cells. Nano Select, 2, 1023-1054(2021). https://doi.org/10.1002/nano.202000249

    [152] C. Ran et al. Bilateral interface engineering toward efficient 2D–3D bulk heterojunction tin halide lead-free perovskite solar cells. ACS Energy Lett., 3, 713-721(2018). https://doi.org/10.1021/acsenergylett.8b00085

    [153] T. Yokoyama et al. Overcoming short-circuit in lead-free CH3NH3SnI3 perovskite solar cells via kinetically controlled gas–solid reaction film fabrication process. J. Phys. Chem. Lett., 7, 776-782(2016). https://doi.org/10.1021/acs.jpclett.6b00118

    [154] H. Wang et al. Fabrication of flexible resistive switching devices based on lead‐free all‐inorganic CsSnBr3 perovskite using a one‐step chemical vapor deposition method. Adv. Electron. Mater., 6, 2000799(2020). https://doi.org/10.1002/aelm.202000799

    [155] J. Xi et al. Multichannel interdiffusion driven FASnI3 film formation using aqueous hybrid salt/polymer solutions toward flexible lead-free perovskite solar cells. Adv. Mater., 29, 1606964(2017). https://doi.org/10.1002/adma.201606964

    [156] C. Mortan et al. Preparation of methylammonium tin iodide (CH3NH3SnI3) perovskite thin films via flash evaporation. Phys. Status Solidi A, 216, 1900209(2019). https://doi.org/10.1002/pssa.201900209

    [157] L. Chouhan et al. Synthesis, optoelectronic properties and applications of halide perovskites. Chem. Soc. Rev., 49, 2869-2885(2020). https://doi.org/10.1039/c9cs00848a

    [158] H. Hu et al. Nucleation and crystal growth control for scalable solution-processed organic–inorganic hybrid perovskite solar cells. J. Mater. Chem. A, 8, 1578-1603(2020). https://doi.org/10.1039/c9ta11245f

    [159] J. E. Bishop et al. Fully spray-coated triple-cation perovskite solar cells. Sci. Rep., 10, 6610(2020). https://doi.org/10.1038/s41598-020-63674-5

    [160] Q. Liang et al. Manipulating crystallization kinetics in high-performance blade-coated perovskite solar cells via cosolvent-assisted phase transition. Adv. Mater., 34, e2200276(2022). https://doi.org/10.1002/adma.202200276

    [161] W. Zuraw et al. Large-area, flexible, lead-free Sn-perovskite solar modules. ACS Energy Lett., 8, 4885-4887(2023). https://doi.org/10.1021/acsenergylett.3c02066

    [162] J. Xi et al. Scalable, template-driven formation of highly crystalline lead-tin halide perovskite films. Adv. Funct. Mater., 31, 2105734(2021). https://doi.org/10.1002/adfm.202105734

    [163] L. Chen et al. In situ SnSe deposition as passivation for scalable and stable quasi-2D lead-tin perovskite solar cells. Energy Environ. Sci., 16, 5315-5324(2023). https://doi.org/10.1039/d3ee02507a

    [164] F. Yao et al. Room-temperature liquid diffused separation induced crystallization for high-quality perovskite single crystals. Nat. Commun., 11, 1194(2020). https://doi.org/10.1038/s41467-020-15037-x

    [165] C. Han et al. Large-size and polarization-sensitive two-dimensional Sn perovskite single crystals. ACS Mater. Lett., 4, 987-994(2022). https://doi.org/10.1021/acsmaterialslett.2c00163

    [166] Q. Liu et al. Theory-guided synthesis of highly luminescent colloidal cesium tin halide perovskite nanocrystals. J. Am. Chem. Soc., 143, 5470-5480(2021). https://doi.org/10.1021/jacs.1c01049

    [167] Q. Dong et al. Electron-hole diffusion lengths >175  μm in solution-grown CH3NH3PbI3 single crystals. Science, 347, 967-970(2015). https://doi.org/10.1126/science.aaa5760

    [168] A. A. Zhumekenov et al. The role of surface tension in the crystallization of metal halide perovskites. ACS Energy Lett., 2, 1782-1788(2017). https://doi.org/10.1021/acsenergylett.7b00468

    [169] Y. Liu et al. Inch-sized high-quality perovskite single crystals by suppressing phase segregation for light-powered integrated circuits. Sci. Adv., 7, eabc8844(2021). https://doi.org/10.1126/sciadv.abc8844

    [170] Z. Yuan et al. Growing MASnI3 perovskite single-crystal films by inverse temperature crystallization. J. Phys. Condens. Matter, 34, 144009(2022). https://doi.org/10.1088/1361-648X/ac4c64

    [171] M. S. Ali et al. Lead-free CsSnCl3 perovskite nanocrystals: rapid synthesis, experimental characterization and DFT simulations. Phys. Chem. Chem. Phys., 23, 22184-22198(2021). https://doi.org/10.1039/d1cp02666f

    [172] A. B. Wong et al. Strongly quantum confined colloidal cesium tin iodide perovskite nanoplates: lessons for reducing defect density and improving stability. Nano Lett., 18, 2060-2066(2018). https://doi.org/10.1021/acs.nanolett.8b00077

    [173] Y. Li et al. Stable inorganic colloidal tin and tin-lead perovskite nanocrystals with ultralong carrier lifetime via Sn(IV) control. J. Am. Chem. Soc., 146, 3094-3101(2024). https://doi.org/10.1021/jacs.3c10060

    [174] J.-K. Chen et al. Advances and challenges in tin halide perovskite nanocrystals. ACS Mater. Lett., 3, 1541-1557(2021). https://doi.org/10.1021/acsmaterialslett.1c00444

    [175] K. Gahlot, J. Meijer, L. Protesescu. Structural and optical control through anion and cation exchange processes for Sn-halide perovskite nanostructures. Nanoscale, 16, 5177-5187(2024). https://doi.org/10.1039/d3nr06075f

    [176] X. Y. Zhou et al. Additive engineering with 2,8-dibromo-dibenzothiophene-S,S-dioxide enabled tin-based perovskite solar cells with 14.98% power conversion efficiency. Energy Environ. Sci., 17, 2837-2844(2024). https://doi.org/10.1039/d3ee03359g

    [177] T. Li et al. Metal chalcogenide electron extraction layers for nip-type tin-based perovskite solar cells. Nat. Commun., 15, 9435(2024). https://doi.org/10.1038/s41467-024-53713-4

    [178] B. Li et al. Suppressing oxidation at perovskite-NiOx interface for efficient and stable tin perovskite solar cells. Adv. Mater., 36, 2309768(2024). https://doi.org/10.1002/adma.202309768

    [179] D. Song et al. Mixed self-assembled monolayers for high-photovoltage tin perovskite solar cells. ACS Energy Lett., 9, 1466-1472(2024). https://doi.org/10.1021/acsenergylett.4c00155

    [180] L. Wang et al. Exceeding 15% performance with energy level tuning in tin-based perovskite solar cells. ACS Energy Lett., 9, 6238-6244(2024). https://doi.org/10.1021/acsenergylett.4c03172

    [181] J. Chen et al. Efficient tin-based perovskite solar cells with trans-isomeric fulleropyrrolidine additives. Nat. Photonics, 18, 464-470(2024). https://doi.org/10.1038/s41566-024-01381-7

    [182] H. Yu et al. Inorganic CsSnI3 perovskite solar cells with an efficiency above 13.6%. ACS Energy Lett., 9, 5870-5878(2024). https://doi.org/10.1021/acsenergylett.4c02742

    [183] M. Hu et al. Highly stable n–i–p structured formamidinium tin triiodide solar cells through the stabilization of surface Sn2+ + cations. Adv. Funct. Mater., 33, 2300693(2023). https://doi.org/10.1002/adfm.202300693

    [184] J. Park et al. Controlled growth of perovskite layers with volatile alkylammonium chlorides. Nature, 616, 724-730(2023). https://doi.org/10.1038/s41586-023-05825-y

    [185] W. Ali et al. Tuning lattice structure of ferroelastic twin-domains achieving efficient perovskite solar cells. ACS Energy Lett., 8, 5070-5078(2023). https://doi.org/10.1021/acsenergylett.3c01936

    [186] J. Qiu et al. Dipolar chemical bridge induced CsPbI3 perovskite solar cells with 21.86% efficiency. Angew. Chem. Int. Ed., 63, e202401751(2024). https://doi.org/10.1002/anie.202401751

    [187] Z. Wang et al. Light-driven dynamic defect-passivation for efficient inorganic perovskite solar cells. Adv. Funct. Mater., 35, 2416118(2024). https://doi.org/10.1002/adfm.202416118

    [188] Z. Xie et al. High-efficiency perovskite solar cells enabled by guanylation reaction for removing MACl residual and in-situ forming 2D perovskite. Angew. Chem. Int. Ed., 64, e202419070(2024). https://doi.org/10.1002/anie.202419070

    [189] Y. Ding et al. Stress regulation via surface micro-etching and reconstruction for enhancing triple-cation perovskite solar cells with an efficiency of 25.54%. Energy Environ. Sci., 17, 9268-9277(2024). https://doi.org/10.1039/d4ee04248d

    [190] X. Zhao et al. Tailored polymer hole-transporting materials with multisite passivation functions for effective buried-interface engineering of inverted quasi-2D perovskite solar cells. Adv. Sci., 11, 2410807(2024). https://doi.org/10.1002/advs.202410807

    [191] Z. Li et al. Uniform phase permutation of efficient Ruddlesden-Popper perovskite solar cells via binary spacers and single crystal coordination. Adv. Mater., 36, 2410408(2024). https://doi.org/10.1002/adma.202410408

    [192] Y. Rong et al. Challenges for commercializing perovskite solar cells. Science, 361, eaat8235(2018). https://doi.org/10.1126/science.aat8235

    [193] C. Sun et al. Well-defined fullerene bisadducts enable high-performance tin-based perovskite solar cells. Adv. Mater., 35, e2205603(2023). https://doi.org/10.1002/adma.202205603

    [194] T.-Y. Teng et al. Electronically manipulated molecular strategy enabling highly efficient tin perovskite photovoltaics. Angew. Chem. Int. Edit., 63, e202318133(2024). https://doi.org/10.1002/anie.202318133

    [195] Y. Chen et al. Mitigating Voc loss in tin perovskite solar cells via simultaneous suppression of bulk and interface nonradiative recombination. ACS Appl. Mater. Interfaces, 14, 41086-41094(2022). https://doi.org/10.1021/acsami.2c12129

    [196] X. Jiang et al. Ultra-high open-circuit voltage of tin perovskite solar cells via an electron transporting layer design. Nat. Commun., 11, 1245(2020). https://doi.org/10.1038/s41467-020-15078-2

    [197] Y. Chen et al. Modulating nucleation and crystal growth of tin perovskite films for efficient solar cells. Nano Lett., 24, 5460-5466(2024). https://doi.org/10.1021/acs.nanolett.4c00474

    [198] H. Yan et al. Efficient passivation of surface defects by Lewis base in lead-free tin-based perovskite solar cells. Mater. Today Energy, 27, 101038(2022). https://doi.org/10.1016/j.mtener.2022.101038

    [199] Y. Chen et al. Multifunctional buried interface modification enables efficient tin perovskite solar cells. Small Methods, 8, 2300029(2024). https://doi.org/10.1002/smtd.202300029

    [200] J.-J. Cao et al. Multifunctional potassium thiocyanate interlayer for eco-friendly tin perovskite indoor and outdoor photovoltaics. Chem. Eng. J., 433, 133832(2022). https://doi.org/10.1016/j.cej.2021.133832

    [201] H. Li et al. Surface reconstruction for tin-based perovskite solar cells. ACS Energy Lett., 7, 3889-3899(2022). https://doi.org/10.1021/acsenergylett.2c01624

    [202] J. Sanchez-Diaz et al. Tin perovskite solar cells with >1,300 h of operational stability in N2 through a synergistic chemical engineering approach. Joule, 6, 861-883(2022). https://doi.org/10.1016/j.joule.2022.02.014

    [203] W. Zhang et al. Organic-free and lead-free perovskite solar cells with efficiency over 11%. Adv. Energy Mater., 12, 2202491(2022). https://doi.org/10.1002/aenm.202202491

    [204] Y.-T. Yang et al. Dual-stage reduction strategy of tin perovskite enables high performance photovoltaic. Angew. Chem. Int. Edit., 64, e202415681(2024). https://doi.org/10.1002/anie.202415681

    [205] O. E. Solis et al. Adjusting the crystallization of tin perovskites through thiophene additives for improved photovoltaic stability. ACS Energy Lett., 9, 5288-5295(2024). https://doi.org/10.1021/acsenergylett.4c01875

    [206] S. Cho et al. Phenylethylammonium-formamidinium-methylammonium quasi-2D/3D tin wide-bandgap perovskite solar cell with improved efficiency and stability. Chem. Eng. J., 446, 137388(2022). https://doi.org/10.1016/j.cej.2022.137388

    [207] B. Chen et al. Reducing the interfacial voltage loss in tin halides perovskite solar cells. Chem. Eng. J., 445, 136769(2022). https://doi.org/10.1016/j.cej.2022.136769

    [208] A. Wijesekara et al. Highly air stable tin halide perovskite photovoltaics using a bismuth capped copper top electrode. Adv. Sci., 10, 2301497(2023). https://doi.org/10.1002/advs.202301497

    [209] C.-H. Kuan et al. Dopant-free pyrrolopyrrole-based (PPr) polymeric hole-transporting materials for efficient tin-based perovskite solar cells with stability over 6000 h. Adv. Mater., 35, 2300681(2023). https://doi.org/10.1002/adma.202300681

    [210] S. Qu et al. Revealing and inhibiting the facet-related ion migration for efficient and stable perovskite solar cells. Angew. Chem. Int. Edit., 64, e202415949(2024). https://doi.org/10.1002/anie.202415949

    [211] Y. Shen et al. Strain regulation retards natural operation decay of perovskite solar cells. Nature, 635, 882-889(2024). https://doi.org/10.1038/s41586-024-08161-x

    [212] X. Zhao et al. Operationally stable perovskite solar modules enabled by vapor-phase fluoride treatment. Science, 385, 433-438(2024). https://doi.org/10.1126/science.adn9453

    [213] C. Gong et al. Silver coordination-induced n-doping of PCBM for stable and efficient inverted perovskite solar cells. Nat. Commun., 15, 4922(2024). https://doi.org/10.1038/s41467-024-49395-7

    [214] Y. Zhao et al. A bilayer conducting polymer structure for planar perovskite solar cells with over 1,400 hours operational stability at elevated temperatures. Nat. Energy, 7, 144-152(2022). https://doi.org/10.1038/s41560-021-00953-z

    [215] B. Chang et al. Phase-pure 2D/3D tin-based perovskite films for solar cells. ACS Energy Lett., 9, 363-372(2024). https://doi.org/10.1021/acsenergylett.3c02616

    [216] F. Du et al. Managing crystallization and phase distribution via 2D perovskite seed crystals for 2D-3D tin-based perovskite solar cells. Adv. Funct. Mater., 35, 2413281(2024). https://doi.org/10.1002/adfm.202413281

    [217] Z. Gao et al. Adhesion-controlled heterogeneous nucleation of tin halide perovskites for eco-friendly indoor photovoltaics. Adv. Mater., 36, 2403413(2024). https://doi.org/10.1002/adma.202403413

    [218] S. F. Hu et al. Narrow bandgap metal halide perovskites for all-perovskite tandem photovoltaics. Chem. Rev., 124, 4079-4123(2024). https://doi.org/10.1021/acs.chemrev.3c00667

    [219] C. Wang et al. Color-stable two-dimensional tin-based perovskite light-emitting diodes: passivation effects of diphenylphosphine oxide derivatives. Adv. Funct. Mater., 33, 2301304(2023). https://doi.org/10.1002/adfm.202301304

    [220] Y. Han et al. Manipulating electron-phonon coupling for efficient tin halide perovskite blue LEDs. Adv. Mater., 2413895(2024). https://doi.org/10.1002/adma.202413895

    [221] Y. Li et al. Boosting CsSnI3-based near-infrared perovskite light-emitting diodes performance via solvent coordination engineering. Infomat, 6, e12537(2024). https://doi.org/10.1002/inf2.12537

    [222] X. Guan et al. Targeted elimination of tetravalent-Sn-induced defects for enhanced efficiency and stability in lead-free NIR-II perovskite LEDs. Nat. Commun., 15, 9913(2024). https://doi.org/10.1038/s41467-024-54160-x

    [223] F. L. Yuan et al. Bright and stable near-infrared lead-free perovskite light-emitting diodes. Nat. Photonics, 18, 170-176(2024). https://doi.org/10.1038/s41566-023-01351-5

    [224] H. Min et al. Spin coating epitaxial heterodimensional tin perovskites for light-emitting diodes. Nat. Nanotechnol., 19, 632-637(2024). https://doi.org/10.1038/s41565-023-01588-9

    [225] S.-Q. Sun et al. Enhancing light outcoupling efficiency via anisotropic low refractive index electron transporting materials for efficient perovskite light-emitting diodes. Adv. Mater., 36, 2400421(2024). https://doi.org/10.1002/adma.202400421

    [226] Y. Gao et al. Highly efficient blue light-emitting diodes based on mixed-halide perovskites with reduced chlorine defects. Sci. Adv., 10, eado5645(2024). https://doi.org/10.1126/sciadv.ado5645

    [227] S.-C. Feng et al. Efficient and stable red perovskite light-emitting diodes via thermodynamic crystallization control. Adv. Mater., 36, 2410255(2024). https://doi.org/10.1002/adma.202410255

    [228] M. Li et al. Acceleration of radiative recombination for efficient perovskite LEDs. Nature, 630, 631-635(2024). https://doi.org/10.1038/s41586-024-07460-7

    [229] H. Liang et al. High color purity lead-free perovskite light-emitting diodes via Sn stabilization. Adv. Sci., 7, 1903213(2020). https://doi.org/10.1002/advs.201903213

    [230] S.-Z. Wang et al. Environment-friendly perovskite light-emitting diodes: progress and perspective. Adv. Mater. Interfaces, 9, 2200772(2022). https://doi.org/10.1002/admi.202200772

    [231] F. Zhang et al. Vapor-assisted in situ recrystallization for efficient tin-based perovskite light-emitting diodes. Adv. Mater., 34, 2203180(2022). https://doi.org/10.1002/adma.202203180

    [232] W. Bai et al. Ligand engineering enables efficient pure red tin-based perovskite light-emitting diodes. Angew. Chem. Int. Ed., 135, e202312728(2023). https://doi.org/10.1002/ange.202312728

    [233] L. M. Kong et al. Fabrication of red-emitting perovskite LEDs by stabilizing their octahedral structure. Nature, 631, 73-79(2024). https://doi.org/10.1038/s41586-024-07531-9

    [234] Z. W. Qi et al. Ligand-pinning induced size modulation of CsPbI3 perovskite quantum dots for red light-emitting diodes. Adv. Funct. Mater., 34, 2405679(2024). https://doi.org/10.1002/adfm.202405679

    [235] Y. C. Wei et al. Overcoming the energy gap law in near-infrared OLEDs by exciton-vibration decoupling. Nat. Photonics, 14, 570-577(2020). https://doi.org/10.1038/s41566-020-0653-6

    [236] M. Vasilopoulou et al. Efficient colloidal quantum dot light-emitting diodes operating in the second near-infrared biological window. Nat. Photonics, 14, 50-56(2020). https://doi.org/10.1038/s41566-019-0526-z

    [237] F. Cao et al. Stability enhancement of lead-free CsSnI3 perovskite photodetector with reductive ascorbic acid additive. InfoMat, 2, 577-584(2020). https://doi.org/10.1002/inf2.12074

    [238] T. Liu et al. Cyano-coordinated tin halide perovskites for wearable health monitoring and weak light imaging. Adv. Mater., 36, 2400090(2024). https://doi.org/10.1002/adma.202400090

    [239] F. Wang et al. Recent progress on electrical and optical manipulations of perovskite photodetectors. Adv. Sci., 8, e2100569(2021). https://doi.org/10.1002/advs.202100569

    [240] M. Han et al. Controllable growth of lead-free all-inorganic perovskite nanowire array with fast and stable near-infrared photodetection. J. Phys. Chem. C, 123, 17566-17573(2019). https://doi.org/10.1021/acs.jpcc.9b03289

    [241] Y. Yang et al. Sn-based quasi-two-dimensional organic–inorganic hybrid halide perovskite for high-performance photodetectors. Appl. Phys. Lett., 119, 161106(2021). https://doi.org/10.1063/5.0068273

    [242] L. He et al. Highly sensitive tin-lead perovskite photodetectors with over 450 days stability enabled by synergistic engineering for pulse oximetry system. Adv. Mater., 35, 2210016(2023). https://doi.org/10.1002/adma.202210016

    [243] T. Liu et al. Quasi-single-crystal tin halide perovskite films with high structural integrity for near-infrared imaging array enabling hidden object recognition. Angew. Chem. Int. Ed., 64, e202418470(2024). https://doi.org/10.1002/anie.202418470

    [244] V. S. Chirvony et al. Unusual spectrally reproducible and high Q-factor random lasing in polycrystalline tin perovskite films. Adv. Mater., 35, e2208293(2023). https://doi.org/10.1002/adma.202208293

    [245] A. L. Alvarado-Leaños et al. Lasing in two-dimensional tin perovskites. ACS Nano, 16, 20671-20679(2022). https://doi.org/10.1021/acsnano.2c07705

    [246] Y. Liang et al. Lasing from mechanically exfoliated 2D homologous Ruddlesden-Popper perovskite engineered by inorganic layer thickness. Adv. Mater., 31, 1903030(2019). https://doi.org/10.1002/adma.201903030

    [247] A. Liu et al. High-performance metal halide perovskite transistors. Nat. Electron., 6, 559-571(2023). https://doi.org/10.1038/s41928-023-01001-2

    [248] S. Shao et al. Field-effect transistors based on formamidinium tin triiodide perovskite. Adv. Funct. Mater., 31, 2008478(2021). https://doi.org/10.1002/adfm.202008478

    [249] H. Zhu et al. Tin perovskite transistors and complementary circuits based on A-site cation engineering. Nat. Electron., 6, 650-657(2023). https://doi.org/10.1038/s41928-023-01019-6

    [250] A. Liu et al. High-performance inorganic metal halide perovskite transistors. Nat. Electron., 5, 78-83(2022). https://doi.org/10.1038/s41928-022-00712-2

    [251] R. N. Bukke et al. Strain relaxation and multidentate anchoring in n-type perovskite transistors and logic circuits. Nat. Electron., 7, 444-453(2024). https://doi.org/10.1038/s41928-024-01165-5

    [252] W. H. Qian et al. Independent memcapacitive switching triggered by bromide ion migration for quaternary information storage. Adv. Mater., 31, e1806424(2019). https://doi.org/10.1002/adma.201806424

    [253] S. Cai et al. Fast-response oxygen optical fiber sensor based on PEA2SnI4 perovskite with extremely low limit of detection. Adv. Sci., 9, 2104708(2022). https://doi.org/10.1002/advs.202104708

    [254] R. Chen et al. Directional regulation of photogenerated carriers in tin halide perovskites for enhanced thermoelectrics. ACS Energy Lett., 9, 5721-5727(2024). https://doi.org/10.1021/acsenergylett.4c02759

    [255] J. Li et al. Bright and efficient CsSnBr3 light-emitting diodes enabled by interfacial reaction-assisted crystallization. Adv. Mater., 2414841(2024). https://doi.org/10.1002/adma.202414841

    [256] J. Ye et al. Extending the defect tolerance of halide perovskite nanocrystals to hot carrier cooling dynamics. Nat. Commun., 15, 8120(2024). https://doi.org/10.1038/s41467-024-52377-4

    [257] J. Chen et al. 1.4% external quantum efficiency 988 nm light emitting diode based on tin-lead halide perovskite. Adv. Mater., 2415958(2024). https://doi.org/10.1002/adma.202415958

    [258] J. Cheng et al. Enhanced electric field minimizing quasi-Fermi level splitting deficit for high-performance tin-lead perovskite solar cells. Adv. Mater., 36, 2410298(2024). https://doi.org/10.1002/adma.202410298

    Feng Yang, Yu Tong, Kun Wang, Yali Chen, Ziyong Kang, Hongqiang Wang, "Recent advances in tin perovskites and their applications," Adv. Photon. 7, 034003 (2025)
    Download Citation