Next Article in Journal
Analysis of Load-Bearing Capacity and Lateral-Torsional Buckling (LTB) Stability of OSB Laminated and I-Beams Made of Wood and Wood-Based Materials
Previous Article in Journal
Nanomechanical and Adhesive Behavior of Electrophoretically Deposited Hydroxyapatite- and Chitosan-Based Coatings on Ti13Zr13Nb Alloy
Previous Article in Special Issue
Ligand-Assisted Purification of Mixed-Halide Perovskite Nanocrystals with Near-Unity PLQY for High-Color-Purity Display Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Robust Self-Trapped Exciton Emission in Sb3+-Engineered Lead-Free Cs4SnBr6 Zero-Dimensional Perovskites

School of Physics and Electronic Engineering, Hanshan Normal University, Chaozhou 521041, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(23), 5324; https://doi.org/10.3390/ma18235324
Submission received: 29 October 2025 / Revised: 18 November 2025 / Accepted: 19 November 2025 / Published: 26 November 2025

Abstract

Zero-dimensional (0D) tin halide perovskites have emerged as promising luminescent materials owing to their broadband emission, high quantum yield, and negligible self-absorption. Yet, their luminescence efficiency and stability remain insufficient for practical optoelectronic applications. Here, Sb3+ dopants are introduced into Cs4SnBr6 through a water-assisted wet ball milling strategy, resulting in bright and thermally robust emission. The doped materials exhibit pronounced self-trapped exciton (STE) luminescence centered at 525 nm with a broad full width at half maximum of 110 nm, a large Stokes shift of approximately ~1.3 eV, and a photoluminescence lifetime of ~0.8 µs. Remarkably, Sb3+ incorporation boosts the photoluminescence quantum yield (PLQY) up to 64% at room temperature while simultaneously improving thermal stability. Correlated spectroscopic analyses reveal that the Sb3+-induced lattice distortion of the [SnBr6]4− octahedra strengthens electron–phonon interactions and elevates the STE binding energy, thereby stabilizing the excited states and suppressing nonradiative losses.

Graphical Abstract

1. Introduction

Metal halide perovskites have emerged as a highly promising class of semiconductors for next-generation optoelectronic devices due to their remarkable intrinsic properties, including strong light absorption, high carrier mobilities, and exceptional photoluminescence quantum yields (PLQY) [1]. However, conventional three-dimensional (3D) lead halide perovskites face two critical challenges: poor intrinsic stability and the toxicity of Pb2⁺, which pose significant environmental and health concerns and severely impede their commercialization. This has spurred increasing interest in the development of lead-free alternatives with improved stability and low toxicity [2,3,4]. Among the emerging alternatives, zero-dimensional (0D) metal halide perovskites have attracted growing attention owing to their unique quantum confinement effect, broadband emission, high PLQY, and tunable photophysical properties [5,6,7,8]. These materials consist of isolated [MX6]4− octahedral units spatially separated by A-site cations, endowing them with stronger exciton binding energies, improved environmental stability, and enhanced PL compared to their 3D counterparts [5,6,9,10,11,12,13,14]. In particular, 0D tin halide perovskites (Cs4SnBr6, X = Br or I) have emerged as environmentally benign emitters with high PLQY, due to their electronic structure similarity to lead-based analogues [12,14]. Nevertheless, Sn2+-based perovskites are inherently unstable—the high-energy 5s2 lone pair of Sn2+ is prone to oxidation to Sn4⁺, leading to structural degradation and PL quenching [14,15,16]. To mitigate this, Zhang et al. [14] used SnF2 as a stable tin precursor to prepare efficient Cs4SnBr6, while rapid thermal treatment (RTT) has also been shown to enhance self-trapped exciton (STE) emission by strengthening electron–phonon coupling and increasing exciton binding energy [17]. Despite these advances, achieving both high emission efficiency and robust thermal stability in 0D tin halide perovskites remains a critical challenge.
Ionic doping has proven to be an effective strategy to tune the optoelectronic properties of halide perovskites without fundamentally altering their host structure [18,19,20,21,22]. By introducing suitable impurity ions, it is possible to engineer band structures, modify the lattice, control defect density, and even generate new emissive centers. For instance, our previous work demonstrated that Mn2⁺ doping not only broadened the emission band of Cs4SnBr6 but also increased the PLQY to ~75% while enhancing thermal stability [23]. Among various dopants, Sb3⁺ ions have garnered particular interest owing to their unique ns2 electronic configuration [22,24,25,26,27]. Sb3⁺ can act as both an emissive center and a lattice modulator, enabling energy level tuning and nonradiative defect passivation [24,28]. Liu et al. [24] synthesized Sb3+-doped Cs3ZnCl5 phosphors with full-visible emission, while Zhou et al. [29] introduced Sb3+ into (C8H22N2Cl)2SnCl6 to passivate trap states and induce the formation of (SbCl5)2− pyramidal units, which stabilized STE states and enhanced stability even under harsh conditions. Similarly, Jin et al. [28] demonstrated that Sb3⁺ doping in nonemissive Rb4CdCl6 softened the lattice and produced intense green emission at 525 nm with a PLQY of 70.2%. These studies highlight the great potential of Sb3+ doping in achieving efficient and stable emission in 0D perovskites. However, a systematic understanding of the effect of Sb3+ incorporation on the structural evolution, photophysics, and thermal stability of Cs4SnBr6 remains elusive.
In this work, we developed Sb3+-doped zero-dimensional Cs4SnBr6–SnF2 perovskites via a water-assisted wet ball milling method and systematically investigated the impact of Sb3+ incorporation on their crystal structure, PL, and thermal stability. The results reveal that appropriate Sb3⁺ doping leads to a dual enhancement of material properties, substantially boosting both the PLQY and thermal robustness of Cs4SnBr6. Comprehensive spectroscopic analyses further elucidate the underlying mechanism, showing that Sb3+ incorporation induces lattice distortion of [SnBr6]4− octahedra, enhances exciton–phonon coupling, and stabilizes STE states, thereby enabling more efficient and thermally stable emission.

2. Experimental Section

2.1. Materials

Cesium bromide (CsBr, 99.9%, Aladdin, Shanghai, China), ammonium bromide (NH4Br, 99.99%, Aladdin, Shanghai, China), tin(II) fluoride (SnF2, 99.9%, Macklin, Shanghai, China), and antimony(III) bromide (SbBr3, 99.99%, Macklin, Shanghai, China) were employed as received without further purification. Deionized water served as the solvent during the milling process.

2.2. Synthesis of Sb3+-Doped Cs4SnBr6–SnF2 Powders

Sb3+-doped Cs4SnBr6–SnF2 perovskite powders were synthesized via a water-assisted wet ball milling route using an MSK-SFM-3-11 planetary ball mill (Hefei Kejing Material Technology Co., Ltd., Hefei, China). CsBr, SnF2, and NH4Br were used as the primary precursors, maintaining a fixed molar ratio of 4 mmol: 1 mmol: 2 mmol, respectively. The amount of Sb3+ was precisely adjusted to 0, 0.05, 0.1, 0.3, and 0.5 mmol to obtain a series of samples denoted as S–x (x = 0, 0.05, 0.1, 0.3, 0.5). In a typical synthesis, the precursor mixture was placed into a ball-milling jar along with 60 μL of deionized water and milled at 600 rpm for 30 min to ensure uniform mixing and reaction. The resulting slurry was subsequently dried at 60 °C under vacuum for 240 min. After cooling to ambient temperature, the dried solid was reintroduced into the jar and subjected to a second milling cycle at 600 rpm for another 30 min, yielding finely mixed Sb3+-doped Cs4SnBr6–SnF2 powders. The samples were not subjected to post-synthesis purification.

2.3. Characterization

The crystal structures of the synthesized powders were examined by X-ray diffraction (XRD) on a MiniFlex 600 diffractometer (Rigaku, Kyoto, Japan). Morphology and elemental distributions were investigated using field-emission scanning electron microscopy (FESEM, Hitachi SU5000, Tokyo, Japan) equipped with an energy-dispersive X-ray spectrometer (EDS, Bruker X-Flash Detector 630M, Karlsruhe, Germany). PL measurements, including steady-state spectra, temperature-dependent PL, and time-resolved PL decay, were carried out using an FLS1000 spectrofluorometer (Edinburgh Instruments, Livingstone, UK) coupled with a closed-cycle helium cryostat and temperature control system. The PLQYs were directly measured by using the same PL spectrometer (Edinburgh Instruments, Livingstone, UK) with an integrating sphere and excitation light at 340 nm. Power-dependent PL measurements were performed on a LabRAM HR Evolution high-resolution Raman spectrometer (HORIBA, Kyoto, Japan). To assess thermal stability, Raman measurements were combined with a Linkam THMS600 heating–cooling stage (Linkam Scientific Instruments Ltd., Redhill, UK), enabling in situ monitoring under controlled thermal environments.

3. Results and Discussion

Figure 1 presents the XRD patterns of Cs4SnBr6-SnF2 powders containing different Sb3+ doping levels. The diffraction features reveal the coexistence of Cs4SnBr6 and CsBr phases, suggesting that a small portion of CsBr precursor remains unreacted during the solid-state milling synthesis. A distinct reflection near 29.7° corresponds to the CsBr phase (PDF #89-3628), whereas the characteristic peaks located at approximately 16.3°, 22.8°, 25.3°, 30.4°, and 33.8° are indexed to the (110), (300), (131), (223), and (330) planes of Cs4SnBr6, respectively. These reflections are in good agreement with previously reported patterns of SnF2-derived Cs4SnBr6 compounds [14]. As the Sb3+ concentration increases, the diffraction peak associated with the (223) plane gradually becomes narrower and more intense, indicating enhanced crystallinity and improved structural ordering within the perovskite lattice. Importantly, no additional peaks or shifts in the main reflections are detected, signifying that the introduction of Sb3+ does not alter the original phase composition. To further clarify the incorporation mechanism of Sb3+ in Cs4SnBr6, we considered the defect chemistry associated with heterovalent substitution. Because Sb3+ replaces Sn2+ on the B-site, each substitution event introduces one additional positive charge into the lattice. Even at the low doping levels employed here, charge neutrality must be maintained and is most plausibly accommodated by a very small concentration of charge-compensating point defects. For clarity, the doped compositions are therefore denoted as Cs4(Sn1−xSbx)Br6, with the understanding that the lattice hosts a minor population of defect species required to balance the heterovalent substitution. This interpretation is consistent with our compositional analysis, which shows a gradual decrease in Sn content with increasing Sb loading and no detectable Cs deficiency. Furthermore, the pronounced PL quenching observed at higher Sb3+ concentrations (Figure 3) aligns with this scenario, as a larger density of charge-compensating defects would introduce additional nonradiative recombination pathways. Taken together, these results support a substitutional Sb3+-for-Sn2+ incorporation mechanism accompanied by a small defect population that becomes increasingly influential at higher doping levels.
Figure 2a shows the SEM micrograph of the S-0.1 sample along with the corresponding elemental mapping images for Cs, Br, Sn, Sb, and F. The uniform and continuous elemental signals across the entire particle region reveal that all constituent elements are evenly dispersed within the microstructure. Such homogeneous distribution also implies that introducing Sb3+ into the lattice does not lead to elemental segregation or local enrichment. The compositional evolution with increasing Sb3+ content is summarized in Figure 2b. A gradual decline in the relative atomic percentage of Sn is observed as the Sb3+ doping level increases. This inverse trend between Sb and Sn contents provides strong evidence that Sb3+ ions are incorporated into the host Cs4SnBr6 framework through substitution of Sn2+ lattice sites. Based on the EDS-derived elemental compositions shown in Figure 2b, the calculated B-site fractions Sb/(Sn + Sb) are 0.0388, 0.0837, 0.2449, and 0.3417 for the 0.05, 0.1, 0.3, and 0.5 mmol samples, respectively. These values confirm that the effective Sb incorporation remains well below the nominal precursor ratios, consistent with the limited solubility of Sb3+ in the Cs4SnBr6 lattice. Such low substitution levels account for the absence of measurable lattice-parameter changes in XRD across the doping series, indicating that Sb3+ induces only subtle local structural perturbations rather than large-scale lattice modification.
The room-temperature PL emission spectra of Sb3+-doped powders with various doping levels are depicted in Figure 3a. Regardless of Sb3+ concentration, all samples display a characteristic broadband green emission centered near 525 nm with an FWHM of roughly 110 nm, a signature typically associated with STE recombination in 0D tin halide perovskites. As shown in the inset, when the emission wavelength is fixed at 525 nm, the optimal excitation wavelength consistently appears around 340 nm. This consistent excitation behavior across all compositions points to a common luminescent origin. The observed Stokes shift of ~1.3 eV aligns well with reported values for Cs4SnBr6 [9,12,14]. The effect of Sb3+ concentration on emission intensity is evident: the PL signal strengthens with increasing dopant content up to 0.1 mmol and then gradually weakens at higher concentrations. As shown in Figure 3b, the emission band shape remains unchanged when the excitation wavelength is varied from 280 to 360 nm. At the optimal level of 0.1 mmol, the PLQY is enhanced from 54.3% for the undoped sample to 64% (Figure 4). Moreover, the excitation and emission intensities follow similar variation trends, providing evidence that moderate Sb3+ incorporation can enrich the population of radiative excited states. Excessive Sb3+, however, may create additional nonradiative recombination pathways, thereby reducing the overall emission efficiency. It should also be noted that the minor CsBr phase detected in the samples is optically inactive in the visible region and thus does not contribute to the measured PL. The emission peak position in Figure 3 and the decay lifetimes in Figure 5 remain essentially unchanged across all doping concentrations, further confirming that the residual CsBr has no measurable influence on the intrinsic luminescence behavior of the Cs4SnBr6 host lattice.
To further elucidate the radiative recombination dynamics of Sb3+-doped Cs4SnBr6-SnF2 powders, TRPL measurements were carried out with the emission monitored at 525 nm. The resulting decay profiles for different Sb3+ concentrations are presented in Figure 5. Each curve can be well described by a biexponential function (Equation (1)), which captures both fast and slow recombination components [30]:
I ( t ) = I 0 + I 1 e x p ( t / τ 1 ) + I 2 e x p ( t / τ 2 )
where I(t) is the emission intensity at time t, I0 is a background constant, Ii (i = 1, 2) are amplitude factors, and τi denote the characteristic lifetimes. The decay traces reveal microsecond-scale dynamics with average lifetimes (τave) of roughly 0.8 μs for all samples. Such long-lived emission is a hallmark of STE-mediated recombination in zero-dimensional tin halide perovskites [11,12,31,32], corroborating the interpretation of the PL spectra. A clear concentration-dependent trend is observed: the average lifetime increases gradually at low Sb3+ doping levels, reaches a maximum at 0.1 mmol, and subsequently decreases with further doping. This behavior parallels the evolution of the PL intensity shown in Figure 3. The lifetime extension at moderate doping levels is likely associated with an increase in radiative STE population, whereas higher dopant concentrations promote nonradiative decay channels, indicative of concentration quenching.
The power-dependent PL response of the S-0.1 powder reveals key information about its radiative recombination behavior. As illustrated in Figure 6a, when the excitation power is gradually increased from 160 to 2820 nW, the emission intensity rises proportionally, while the peak energy remains fixed. The absence of peak shifts suggests that the emission originates from an intrinsic transition rather than defect-assisted states. Over the entire excitation range, the integrated PL intensity follows a power-law dependence, described as follows:
I = η I 0 k
where I is the integrated emission intensity, P represents the excitation power, and k is a characteristic exponent related to the dominant recombination channel. A best fit gives k = 1.02, consistent with exciton-mediated radiative recombination rather than defect- or band-edge-dominated processes [33,34]. When considered together with the large Stokes shift (~1.30 eV), broad emission (FWHM ≈ 110 nm), and microsecond-scale decay lifetime, these results support that the green emission arises from radiative relaxation of STEs within Jahn–Teller distorted [SnBr6]4− units [10,12,32]. To probe the thermal response of this emission, variable-temperature PL spectra were recorded between 80 and 300 K (Figure 6b). The emission intensity becomes stronger at lower temperatures, which can be attributed to the suppression of thermally activated nonradiative recombination channels. Since the luminescence of STEs strongly depends on exciton binding strength, the temperature-dependent integrated PL intensity was analyzed using an Arrhenius model:
I P L ( T ) = I P L ( T 0 ) 1 + β e x p ( E b / k B T )
where IPL(T0) is the PL intensity at 80 K, β is a prefactor related to nonradiative recombination, kB is Boltzmann’s constant, and Eb denotes the activation energy for thermal quenching. The fitting yields Eb ≈ 610 meV (Figure 6d), substantially exceeding the value of the pristine sample [17], implying strong exciton localization and effective suppression of thermal detrapping in the doped system. The narrowing of the emission linewidth with decreasing temperature provides further insight into the coupling between electronic states and phonons. According to Stadler’s model, the temperature dependence of the FWHM is expressed as follows:
F W H M ( T ) = 2.36 S ω p h o n o n c o t h 2 k B T ω p h o n o n
where S is the Huang–Rhys factor, ℏω represents the phonon energy, and kB is Boltzmann’s constant. Fitting the data yields S ≈ 38.9 and ℏω ≈ 18.9 meV (Figure 6c), indicating pronounced electron–phonon interaction. This phonon energy (~152 cm−1) corresponds to the higher-frequency component of the Sn–Br vibrational manifold (inset of Figure 6c) [23,35], demonstrating that this vibrational mode is involved in STE formation. The enhanced STE emission observed at 0.1 mmol Sb3+ doping can therefore be attributed to the combined effect of stronger electron-phonon coupling and an increased exciton binding energy, which together promote efficient radiative recombination. At higher Sb3+ concentrations, however, excessive lattice distortion and defect formation introduce additional nonradiative channels, explaining the reduced PL intensity beyond the optimal doping level. It is also noteworthy that Sb3+ does not introduce dopant-centered emission; instead, it serves as an ns2-type lattice modulator that perturbs the [SnBr6]4− octahedra and stabilizes intrinsic STE emission without altering its spectral characteristics. This behavior differs fundamentally from that of Mn2+-doped Cs4SnBr6, where Mn2+ not only modifies the lattice and enhances electron–phonon coupling but also acts as an activator center with its own radiative transition (Mn2+ 4T16A1), resulting in dual-band emission [23].
To assess the thermal resilience of the Sb3+-doped Cs4SnBr6 system, the integrated PL intensity was tracked during a full heating–cooling cycle. As illustrated in Figure 7, increasing the temperature from ambient conditions to 398 K leads to a continuous decrease in PL intensity, reflecting thermally activated quenching of the emissive states. Following a single heating–cooling cycle, approximately 40% of the original emission intensity is recovered. For comparison, undoped Cs4SnBr6 undergoes a far more pronounced degradation, retaining less than 10% of its initial PL intensity after similar thermal cycling [23]. This contrasting behavior points to a stabilization effect introduced by Sb3+ incorporation, which likely strengthens the self-trapped exciton states and suppresses competing nonradiative recombination pathways, thereby enhancing the thermal durability of the emission.

4. Conclusion

In this work, Sb3+ incorporation into Cs4SnBr6 zero-dimensional perovskites was achieved using a water-assisted wet ball-milling approach, enabling precise modulation of their photoluminescence behavior. A clear enhancement in both emission efficiency and thermal tolerance was observed, with the PLQY reaching 64% at the optimal doping level. Detailed investigations combining time-resolved PL, power-dependent PL, temperature-dependent measurements, and Raman analysis uncovered the underlying mechanism. The Sb3+ dopants induce lattice distortion of the [SnBr6]4− framework, intensifying electron–phonon interactions and reinforcing the exciton binding energy of self-trapped excitons. These effects collectively promote radiative recombination while suppressing nonradiative loss pathways, resulting in more efficient and stable light emission. This strategy not only offers an effective means to manipulate the optical response of lead-free 0D tin halide perovskites but also sheds light on new design principles for robust luminescent materials. The strong and thermally robust green STE emission positions these phosphors as promising candidates for solid-state lighting and wide-gamut display technologies, particularly as environmentally benign, lead-free alternatives to traditional perovskite emitters. Moreover, the lead-free composition and low-temperature, scalable wet-milling synthesis route underscore their potential for sustainable photonic applications and large-scale industrial production.

Author Contributions

Investigation: H.W., W.Z., J.S., Z.L., Y.Z. and T.Q.; formal analysis: H.W., W.Z., R.H., J.S., Z.L., Y.Z. and H.L.; writing—original draft: H.W.; writing—review and editing: R.H. and H.L.; funding acquisition: R.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the National Natural Science Foundation of China (National Natural Science Foundation of China, 62504066), Research Projects of the Department of Education of Guangdong Province (Department of Education of Guangdong Province, 2024ZDZX1026), Program of Hanshan Normal University (Hanshan Normal University, XTP202402), and Project of Guangdong Provincial Education Science Planning (Department of Education of Guangdong Province, 2024GXJK390).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dong, H.; Ran, C.; Gao, W.; Zhang, Z.; Li, N.; Xia, Y.; Huang, B.; Liu, Z.; Chen, Y.; Li, X. Metal Halide Perovskite for Next-Generation Optoelectronics: Progresses and Prospects. eLight 2023, 3, 19. [Google Scholar] [CrossRef]
  2. Kumar, A.; Singh, S.; Satapathi, S. Impact of Indium Doping in Lead-Free (CH3NH3)3Bi2−xInxI9 Perovskite Photovoltaics for Indoor and Outdoor Light Harvesting. ACS Appl. Electron. Mater. 2024, 6, 8360–8368. [Google Scholar] [CrossRef]
  3. Li, C.; Ma, J.; Li, P.; Song, Y.; Huang, W. Sensing and Ionovoltaic Power Generation of Two-Dimensional Cs3Sb2X9 (X = Cl/Br/I) Perovskite Microcrystals. ACS Appl. Electron. Mater. 2024, 6, 5255–5268. [Google Scholar]
  4. Dávid, A.; Morát, J.; Chen, M.; Gao, F.; Fahlman, M.; Liu, X. Mapping Uncharted Lead-Free Halide Perovskites and Related Low-Dimensional Structures. Materials 2024, 17, 491. [Google Scholar] [CrossRef]
  5. Li, M.; Xia, Z. Recent Progress of Zero-Dimensional Luminescent Metal Halides. Chem. Soc. Rev. 2021, 50, 2626–2662. [Google Scholar] [CrossRef]
  6. Sun, S.; Lu, M.; Gao, X.; Shi, Z.; Bai, X.; Yu, W.W.; Zhang, Y. 0D Perovskites: Unique Properties, Synthesis, and Their Applications. Adv. Sci. 2021, 8, 2102689. [Google Scholar] [CrossRef]
  7. Cheng, S.; Nikl, M.; Beitlerova, A.; Kucerkova, R.; Du, X.; Niu, G.; Jia, Y.; Tang, J.; Ren, G.; Wu, Y. Ultrabright and Highly Efficient All-Inorganic Zero-Dimensional Perovskite Scintillators. Adv. Opt. Mater. 2021, 9, 2100460. [Google Scholar] [CrossRef]
  8. Yakunin, S.; Benin, B.M.; Shynkarenko, Y.; Nazarenko, O.; Bodnarchuk, M.I.; Dirin, D.N.; Hofer, C.; Cattaneo, S.; Kovalenko, M.V. High-Resolution Remote Thermometry and Thermography Using Luminescent Low-Dimensional Tin-Halide Perovskites. Nat. Mater. 2019, 18, 846–852. [Google Scholar] [CrossRef]
  9. Benin, B.M.; Dirin, D.N.; Morad, V.; Wörle, M.; Yakunin, S.; Rainò, G.; Nazarenko, O.; Fischer, M.; Infante, I.; Kovalenko, M.V. Highly Emissive Self-Trapped Excitons in Fully Inorganic Zero-Dimensional Tin Halides. Angew. Chem. Int. Ed. 2018, 57, 11329–11333. [Google Scholar] [CrossRef]
  10. Tan, L.; Wang, W.; Li, Q.; Huang, Y.; Liu, J.; Sun, Z. Colloidal Syntheses of Zero-Dimensional Cs4SnX6 (X = Br, I) Nanocrystals with High Emission Efficiencies. Chem. Commun. 2020, 56, 387–390. [Google Scholar] [CrossRef]
  11. Lin, H.; Zhou, C.; Tian, Y.; Siegrist, T.; Ma, B. Low-Dimensional Organometal Halide Perovskites. ACS Energy Lett. 2017, 3, 54–62. [Google Scholar] [CrossRef]
  12. Zhang, X.; Wang, H.; Wang, S.; Wang, Y.; Yang, Q.; Li, X.; Wang, H.; Gu, L.; Yu, Y.; Huang, K.; et al. Room Temperature Synthesis of All Inorganic Lead-Free Zero-Dimensional Cs4SnBr6 and Cs3KSnBr6 Perovskites. Inorg. Chem. 2019, 59, 533–538. [Google Scholar] [CrossRef]
  13. Chen, D.; Wan, Z.; Chen, X.; Yuan, Y.; Zhong, J. Large-Scale Room-Temperature Synthesis and Optical Properties of Perovskite-Related Cs4PbBr6 Fluorophores. J. Mater. Chem. C 2016, 4, 10646–10653. [Google Scholar] [CrossRef]
  14. Zhang, Q.; Liu, S.; He, M.; Zheng, W.; Wan, Q.; Liu, M.; Liao, X.; Zhan, W.; Yuan, C.; Liu, J.; et al. Stable Lead-Free Tin Halide Perovskite with Operational Stability >1200 h by Suppressing Tin (II) Oxidation. Angew. Chem. Int. Ed. 2022, 61, e202205463. [Google Scholar] [CrossRef] [PubMed]
  15. Jellicoe, T.C.; Richter, J.M.; Glass, H.F.J.; Tabachnyk, M.; Brady, R.; Dutton, S.E.; Rao, A.; Friend, R.H.; Credgington, D.; Greenham, N.C.; et al. Synthesis and Optical Properties of Lead-Free Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2016, 138, 2941–2944. [Google Scholar] [CrossRef] [PubMed]
  16. Kang, C.; Rao, H.; Fang, Y.; Zeng, J.; Pan, Z.; Zhong, X. Antioxidative Stannous Oxalate Derived Lead-Free Stable CsSnX3 (X = Cl, Br, and I) Perovskite Nanocrystals. Angew. Chem. Int. Ed. 2021, 60, 660–665. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, H.; Lin, Z.; Song, J.; Huang, Y.; Lu, X.; Xu, J.; Zhong, Y.; Li, Y.; Chen, J. Boosting the Self-Trapped Exciton Emission in Cs4SnBr6 Zero-Dimensional Perovskite via Rapid Heat Treatment. Nanomaterials 2023, 13, 2259. [Google Scholar] [CrossRef]
  18. He, C.; Qiu, J.; Mu, Z.; Zhang, Z.; Xia, Z.; Guo, X. Unlocking the Potential of Halide Perovskites Through Doping. CCS Chem. 2023, 5, 1961–1972. [Google Scholar] [CrossRef]
  19. Stroyuk, O.; Raievska, O.; Hauch, J.; Brabec, C.J. Doping/Alloying Pathways to Lead-Free Halide Perovskites with Ultimate Photoluminescence Quantum Yields. Angew. Chem. Int. Ed. 2023, 62, e202212668. [Google Scholar] [CrossRef]
  20. Chang, T.; Wei, Q.; Zeng, R.; Wang, S.; Chui, Y.; Zhao, Z.; Tang, Z.; Lu, C.; Huang, X. Efficient Energy Transfer in Te4+-Doped Cs2ZrCl6 Vacancy-Ordered Perovskites and Ultrahigh Moisture Stability via A-Site Rb-Alloying Strategy. J. Phys. Chem. Lett. 2021, 12, 1829–1837. [Google Scholar] [CrossRef]
  21. Luo, J.; Wang, X.; Li, S.; Liu, J.; Guo, Y.; Niu, G.; Yao, L.; Fu, Y.; Gao, L.; Dong, Q.; et al. Efficient and Stable Emission of Warm-White Light from Lead-Free Halide Double Perovskites. Nature 2018, 563, 541–545. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, X.; Shen, Q.; Chen, Y.; Zhang, Y.; Ding, J.; Bai, Y.; Gélvez-Rueda, M.C.; Grozema, F.C.; Gao, X. Bright Luminescence of Sb Doping in All-Inorganic Zinc Halide Perovskite Variant. J. Alloys Compd. 2022, 895, 162610. [Google Scholar] [CrossRef]
  23. Lin, Z.; Wang, A.; Huang, R.; Wu, H.; Huang, Y.; Lu, X.; Xu, J.; Zhong, Y.; Li, Y.; Chen, J. Manipulating the Sublattice Distortion Induced by Mn2+ Doping for Boosting the Emission Characteristics of Self-Trapped Excitons in Cs4SnBr6. J. Mater. Chem. C 2023, 11, 5680–5687. [Google Scholar] [CrossRef]
  24. Liu, X.; Zhang, W.; Xu, R.; Yang, Y.; Yang, Z.; Ma, Z.; Zhao, J.; Zhao, C. Bright Tunable Luminescence of Sb3+ Doping in Zero-Dimensional Lead-Free Halide Cs3ZnCl5 Perovskite Crystals. Dalton Trans. 2022, 51, 10029–10035. [Google Scholar] [CrossRef]
  25. Jing, Y.; Liu, Y.; Jiang, X.; Li, M.; Mao, X.; Li, M.; Li, Y.; Wang, S.; Zhao, J.; Li, X. Sb3+ Dopant and Halogen Substitution Triggered Highly Efficient and Tunable Emission in Lead-Free Metal Halide Single Crystals. Chem. Mater. 2020, 32, 5327–5334. [Google Scholar] [CrossRef]
  26. Han, P.; Luo, C.; Yang, S.; Yang, Y.; Deng, W.; Han, K. All-Inorganic Lead-Free 0D Perovskites by a Doping Strategy to Achieve a PLQY Boost from <2% to 90%. Angew. Chem. Int. Ed. 2020, 59, 12709–12713. [Google Scholar]
  27. Su, B.; Li, M.; Song, E.; Wang, Z.; Zhang, Q.; Moon, C.; Liu, C.; Zhou, G.; Wang, G.; Xia, Z. Sb3+-Doping in Cesium Zinc Halides Single Crystals Enabling High-Efficiency Near-Infrared Emission. Adv. Funct. Mater. 2021, 31, 2105316. [Google Scholar] [CrossRef]
  28. Jin, J.; Peng, Y.; Xu, Y.; Li, T.; Li, L.; Wang, S.; Li, S.; Lu, X.; Zhang, J.; Wang, Y.; et al. Bright Green Emission from Self-Trapped Excitons Triggered by Sb3+ Doping in Rb4CdCl6. Chem. Mater. 2022, 34, 5717–5725. [Google Scholar] [CrossRef]
  29. Zhou, L.; Zhang, L.; Li, H.; Chen, R.; Wang, J.; Hu, J.; Li, X.J.; Jiao, H.; Zhang, Z.; Zhai, T.; et al. Defect Passivation in Air-Stable Tin(IV)-Halide Single Crystal for Emissive Self-Trapped Excitons. Adv. Funct. Mater. 2021, 31, 2108561. [Google Scholar] [CrossRef]
  30. Lin, K.H.; Liou, S.C.; Chen, W.L.; Cheng, H.C.; Chen, C.C.; Dong, C.L.; Chen, J.F.; Chen, K.H.; Chen, L.C. Tunable and Stable UV-NIR Photoluminescence from Annealed SiOx with Si Nanoparticles. Opt. Express 2013, 21, 23416–23424. [Google Scholar] [CrossRef]
  31. Chiara, R.; Ciftci, Y.O.; Queloz, V.I.E.; Törmä, P.; Scopelliti, R.; El Kazzi, M.; Ongaro, N.; Sessolo, M.; Bolink, H.J.; Palazon, F.; et al. Green-Emitting Lead-Free Cs4SnBr6 Zero-Dimensional Perovskite Nanocrystals with Improved Air Stability. J. Phys. Chem. Lett. 2020, 11, 618–623. [Google Scholar] [CrossRef]
  32. Li, S.; Luo, J.; Liu, J.; Tang, J. Self-Trapped Excitons in All-Inorganic Halide Perovskites: Fundamentals, Status, and Potential Applications. J. Phys. Chem. Lett. 2019, 10, 1999–2007. [Google Scholar] [CrossRef]
  33. Ulloa, J.M.; Llorens, J.M.; Alén, B.; Reyes, D.F.; Sales, D.L.; González, D.; Hierro, A. High Efficient Luminescence in Type-II GaAsSb-Capped InAs Quantum Dots upon Annealing. Appl. Phys. Lett. 2012, 101, 253112. [Google Scholar] [CrossRef]
  34. Bergman, L.; Chen, X.-B.; Morrison, J.L.; Huso, J.; Purdy, A.P. Photoluminescence Dynamics in Ensembles of Wide-Band-Gap Nanocrystallites and Powders. J. Appl. Phys. 2004, 96, 675–682. [Google Scholar] [CrossRef]
  35. Kuok, M.H.; Tan, L.S.; Shen, Z.X.; Huan, C.H.; Mok, K.F. A Raman Study of RbSnBr3. Solid State Commun. 1996, 97, 497–501. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Sb3+-doped Cs4SnBr6-SnF2 powders with various doping concentrations.
Figure 1. XRD patterns of Sb3+-doped Cs4SnBr6-SnF2 powders with various doping concentrations.
Materials 18 05324 g001
Figure 2. (a) SEM image and elemental mapping of Cs, Br, Sn, Sb, and F in the S-0.1 sample; (b) elemental composition evolution as a function of Sb3+ concentration.
Figure 2. (a) SEM image and elemental mapping of Cs, Br, Sn, Sb, and F in the S-0.1 sample; (b) elemental composition evolution as a function of Sb3+ concentration.
Materials 18 05324 g002
Figure 3. (a) PL emission spectra of Cs4SnBr6–SnF2 powders with different Sb3+ concentrations. The inset shows excitation spectra monitored at 525 nm; (b) PL spectra of the Cs4SnBr6 sample S-0.1 recorded under different excitation wavelengths.
Figure 3. (a) PL emission spectra of Cs4SnBr6–SnF2 powders with different Sb3+ concentrations. The inset shows excitation spectra monitored at 525 nm; (b) PL spectra of the Cs4SnBr6 sample S-0.1 recorded under different excitation wavelengths.
Materials 18 05324 g003
Figure 4. PLQY of the Cs4SnBr6 sample containing 0.1 mmol Sb3+. The inset shows the reference PLQY of the undoped Cs4SnBr6 sample. The PLQY is determined from the ratio of the number of emitted photons to the number of absorbed photons (Q = Nem/Nabs). The number of emitted photons (Nem) is obtained from the integrated area of the spectrally corrected emission of the Sb3+-doped sample after subtracting that of the reference (Nem = Aem sample-Aem ref). The number of absorbed photons (Nabs) is calculated from the difference in the integrated Rayleigh scattering peak areas of the reference and doped samples (Nabs = Ascat ref-Ascat sample).
Figure 4. PLQY of the Cs4SnBr6 sample containing 0.1 mmol Sb3+. The inset shows the reference PLQY of the undoped Cs4SnBr6 sample. The PLQY is determined from the ratio of the number of emitted photons to the number of absorbed photons (Q = Nem/Nabs). The number of emitted photons (Nem) is obtained from the integrated area of the spectrally corrected emission of the Sb3+-doped sample after subtracting that of the reference (Nem = Aem sample-Aem ref). The number of absorbed photons (Nabs) is calculated from the difference in the integrated Rayleigh scattering peak areas of the reference and doped samples (Nabs = Ascat ref-Ascat sample).
Materials 18 05324 g004
Figure 5. Time-resolved PL decay curves of Cs4SnBr6-SnF2 samples with various Sb3+ concentrations recorded at 525 nm. All measurements were performed under 375 nm excitation using laser pulses with a duration of 70 ps. The instrument response function of our setup is on the nanosecond scale.
Figure 5. Time-resolved PL decay curves of Cs4SnBr6-SnF2 samples with various Sb3+ concentrations recorded at 525 nm. All measurements were performed under 375 nm excitation using laser pulses with a duration of 70 ps. The instrument response function of our setup is on the nanosecond scale.
Materials 18 05324 g005
Figure 6. (a) Power-dependent PL intensity of S-0.1 powder with k = 1.02, consistent with excitonic emission. The inset shows corresponding PL spectra under different excitation powers; (b) temperature-dependent PL spectra (80–300 K) of S-0.1 powder; (c) temperature evolution of FWHM fitted with the Huang–Rhys model for S-0.1 powder. The inset displays the Raman spectrum showing Sn-Br stretching at 130 cm−1; (d) Arrhenius fitting of integrated PL intensity for S-0.1 powder yielding an exciton binding energy of 610 meV.
Figure 6. (a) Power-dependent PL intensity of S-0.1 powder with k = 1.02, consistent with excitonic emission. The inset shows corresponding PL spectra under different excitation powers; (b) temperature-dependent PL spectra (80–300 K) of S-0.1 powder; (c) temperature evolution of FWHM fitted with the Huang–Rhys model for S-0.1 powder. The inset displays the Raman spectrum showing Sn-Br stretching at 130 cm−1; (d) Arrhenius fitting of integrated PL intensity for S-0.1 powder yielding an exciton binding energy of 610 meV.
Materials 18 05324 g006
Figure 7. Normalized PL intensity of the S-0.1 sample recorded during heating–cooling cycles.
Figure 7. Normalized PL intensity of the S-0.1 sample recorded during heating–cooling cycles.
Materials 18 05324 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, H.; Zhou, W.; Huang, R.; Song, J.; Lin, Z.; Zhang, Y.; Qiu, T.; Li, H. Robust Self-Trapped Exciton Emission in Sb3+-Engineered Lead-Free Cs4SnBr6 Zero-Dimensional Perovskites. Materials 2025, 18, 5324. https://doi.org/10.3390/ma18235324

AMA Style

Wu H, Zhou W, Huang R, Song J, Lin Z, Zhang Y, Qiu T, Li H. Robust Self-Trapped Exciton Emission in Sb3+-Engineered Lead-Free Cs4SnBr6 Zero-Dimensional Perovskites. Materials. 2025; 18(23):5324. https://doi.org/10.3390/ma18235324

Chicago/Turabian Style

Wu, Haixia, Wendi Zhou, Rui Huang, Jie Song, Zhenxu Lin, Yi Zhang, Tianpei Qiu, and Hongliang Li. 2025. "Robust Self-Trapped Exciton Emission in Sb3+-Engineered Lead-Free Cs4SnBr6 Zero-Dimensional Perovskites" Materials 18, no. 23: 5324. https://doi.org/10.3390/ma18235324

APA Style

Wu, H., Zhou, W., Huang, R., Song, J., Lin, Z., Zhang, Y., Qiu, T., & Li, H. (2025). Robust Self-Trapped Exciton Emission in Sb3+-Engineered Lead-Free Cs4SnBr6 Zero-Dimensional Perovskites. Materials, 18(23), 5324. https://doi.org/10.3390/ma18235324

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop