Next Article in Journal
Sol-Gel Synthesis and Photoluminescence Properties of a Far-Red Emitting Phosphor BaLaMgTaO6:Mn4+ for Plant Growth LEDs
Previous Article in Journal
Unveiling the Effects of Quicklime on the Properties of Sulfoaluminate Cement–Ordinary Portland Cement–Mineral Admixture Repairing Composites and Their Sulphate Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Improved Photoluminescence of SnF2-Derived CsSnCl3-SnF2:Mn2+ Perovskites via Rapid Thermal Treatment

School of Materials Science and Engineering, Hanshan Normal University, Chaozhou 521041, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(11), 4027; https://doi.org/10.3390/ma16114027
Submission received: 4 May 2023 / Revised: 25 May 2023 / Accepted: 25 May 2023 / Published: 28 May 2023

Abstract

:
We report a rapid synthesis method for producing CsSnCl3:Mn2+ perovskites, derived from SnF2, and investigate the effects of rapid thermal treatment on their photoluminescence properties. Our study shows that the initial CsSnCl3:Mn2+ samples exhibit a double luminescence peak structure with PL peaks at approximately 450 nm and 640 nm, respectively. These peaks originate from defect-related luminescent centers and the 4T1→6A1 transition of Mn2+. However, as a result of rapid thermal treatment, the blue emission is significantly reduced and the red emission intensity is increased nearly twofold compared to the pristine sample. Furthermore, the Mn2+-doped samples demonstrate excellent thermal stability after the rapid thermal treatment. We suggest that this improvement in photoluminescence results from enhanced excited-state density, energy transfer between defects and the Mn2+ state, as well as the reduction of nonradiative recombination centers. Our findings provide valuable insights into the luminescence dynamics of Mn2+-doped CsSnCl3 and open up new possibilities for controlling and optimizing the emission of rare-earth-doped CsSnCl3.

1. Introduction

In recent years, all-inorganic lead halide perovskite (CsPbX3, X=Cl, Br, I) has garnered significant attention in optoelectronic devices due to its outstanding photoelectric properties, such as high photoluminescent quantum yield (PLQY), excellent defect tolerance, and tunable light emission [1,2,3,4]. Currently, LED devices based on CsPbX3 have surpassed 20% external quantum efficiency. Further, the stability issues related to lead-based perovskites have been effectively improved via packaging, increasing their lifespan from a few hours to hundreds of hours. However, the high toxicity of lead poses a threat to human health and the ecological environment, thereby impeding their practical use. Therefore, there is a rising demand for non-toxic or less toxic elements to replace lead, such as equivalent substitution (Sn2+, Ge2+) and isovalent substitution (Bi3+, Ag+) to form lead-free perovskite luminescent materials [5,6,7,8,9,10]. It is proposed to replace Pb in lead–halide perovskite with divalent Sn and Ge cations as they both satisfy the prerequisites for coordination and charge balance. Sn2+ has an ion radius comparable to Pb2+ (1.35 Å for Sn2+ and 1.49 Å for Pb2+), resulting in the avoidance of significant lattice vibration caused by substitution. Additionally, Sn2+ possesses a similar ns2 electronic configuration to Pb2+, making it possible to form a perovskite with the formula ASnX3, akin to lead-based counterparts [11,12,13,14,15,16,17]. However, the formation energy of defects in Sn-based perovskites is relatively low (~250 meV), making it easy to form a defect density of up to 1019 cm–3 in Sn-based perovskites, which results in a PLQY of as low as 0.14% in these Sn-based perovskites. Additionally, Sn2+ is susceptible to oxidation into Sn4+ in air, leading to structural instability of perovskites and luminescence quenching. Recently, Zhang et al. employed SnF2 instead of easily oxidized SnBr2 as a tin source and effectively improved the structural stability of Cs4SnBr6 perovskite through introducing F to inhibit the oxidation of Sn2+ in tin-based perovskite [18]. Despite the progress, the luminescence stability of cesium tin chloride perovskite is still poor, and the luminescence efficiency is too low to meet the application requirements. In particular, the synthesis and photoluminescence properties of SnF2-derived cesium–tin perovskite (CsSnCl3-SnF2) have not been reported so far.
On the other hand, impurity doping is considered an effective method to improve the luminescence properties and stability of tin-based perovskites. For instance, Dawson et al. successfully improved the stability of tin-based perovskites through replacing Sn2+ sites with Mn2+ [19]. Similarly, Hou et al. discovered that Mn2+ doping could improve the electron energy level structure and stability of tin halide perovskites and also greatly enhance the luminescence efficiency of CsSnCl3:Mn2+ to around 2% [20]. However, the low-temperature chemical synthesis of Mn-doped tin-based perovskites complicates the full activation of the Mn2+ luminescent centers. Due to tin-based halide perovskite materials’ poor thermal stability, conventional thermal annealing may not effectively activate the Mn ion luminescent centers in this kind of material. Rapid thermal treatment (RTT), a transient process that is characterized by rapid heating and cooling, provides an alternative way to thermally activate the Mn ion luminescent centers of tin-based perovskites. Here, we describe a rapid synthesis method for SnF2-derived CsSnCl3:Mn2+ and investigate the effect of RTT on the photoluminescent (PL) properties of CsSnCl3:Mn2+ perovskites. Pristine CsSnCl3:Mn2+ displays a double luminescence peak structure, which includes PL peaks at around 450 nm and 640 nm, respectively. These peaks originate from defect-related luminescent centers and the 4T1→6A1 transition of Mn2+, respectively. We found that RTT significantly reduced the blue emission and increased the red emission intensity nearly twofold compared to the pristine sample. Furthermore, the Mn2+-doped samples exhibited excellent thermal stability after RTT processing. We discuss the improved PL based on our analysis of PL excitation spectra, XRD patterns, and PL decay traces.

2. Materials and Methods

Mn2+-doped CsSnCl3 were synthesized using a water-assisted wet ball-milling method. The reactant precursors included cesium chloride (4 mmol, CsCl, Aladdin, 99.9%), stannous fluoride (1 mmol, SnF2, Macklin, 99.9%), ammonium bromide (1 mmol, NH4Cl, Aladdin, 99.99%), and manganous chloride (MnCl2, Macklin, 99%). To obtain Mn2+-doped CsSnCl3 with different levels, the molar ratios of CsCl, SnF2, and NH4Cl were maintained at 1:1:2 mmol, whereas the molar ratio of MnCl2 was kept at 1 mmol. Firstly, the precursors were loaded into a jar and mixed with 60 μL of deionized water. A ball milling process was then performed for 30 min at a speed of 600 rpm. The resulting product was subsequently dried in a vacuum-drying oven for 120 min at room temperature and annealed at 200 °C using a simple rapid thermal treatment (RTT) process. The RTT process was carried out on a rapid thermal processor, heating the sample to the annealing temperature at a rate of 10 °C s−1. After remaining at the annealing temperature of 200 °C for 60–300 s, the system was rapidly cooled down to room temperature. Upon cooling to room temperature, the Mn2+-doped CsSnCl3 powder was obtained via ball milling for 30 min at a speed of 600 rpm. Figure 1 illustrates the Mn2+-doped CsSnCl3 synthesis process using water-assisted ball-milling at room temperature, followed by the RTT process. The crystal structures of Mn2+-doped CsSnCl3 were characterized using X-ray diffraction (XRD) (Bruker D8 Advance, Karlsruhe, Germany) at 35 kV and 35 mA. The compositions of the Mn2+-doped CsSnCl3 were analyzed through energy dispersive spectroscopy (EDS) (Bruker EDS QUANTAX, Karlsruhe, Germany). The structure of the Mn2+-doped CsSnCl3 sample (S-Mn2+-doping) was characterized via scanning electron microscopy (SEM, Hitachi SU5000, Tokyo, Japan). PL measurements, including temperature-dependent PL spectra, PL excitation (PLE) spectra, and time-resolved PL spectra, were carried out using a PL spectrometer (FLS1000, Edinburgh Instrument, Livingstone, UK).

3. Results and Discussion

Figure 2a displays the SEM image obtained from the Mn2+-doped CsSnCl3 powders prior to the RTT process. The EDS elemental mapping of the Mn2+-doped CsSnCl3 reveals a well-distributed presence of Cs, Pb, Cl, Mn, and F elements. Figure 2b indicates that the CsSnCl3 powder retains a uniform distribution of Cs, Sn, Cl, Mn, and F elements even after the RTT process.
Figure 3a depicts the PL spectrum obtained from the Mn2+-doped CsSnCl3 powders prior to the RTT process. Pristine CsSnCl3:Mn2+ shows a double luminescence peak structure with PL peaks at ~450 nm and ~640 nm, respectively. As portrayed in the inset of Figure 3a, the 580 nm PL band from CsSnCl3 powders without Mn2+ doping is due to the radiative recombination of self-localized excitons [21]. The addition of Mn2+ doping into CsSnCl3 results in a significant 60 nm redshift in the PL, which is ascribed from the 4T1→6A1 transition of Mn2+ [22]. To comprehend the origin of the blue emission, the temperature-dependent PL spectra of the pristine CsSnCl3:Mn2+ were measured from 78 to 298 K. As illustrated in Figure 3b, the blue emission displayed an increase in PL intensity and a reduction in the full width at half maximum (FWHM) with decreasing temperature, which is due to thermal quenching. Nevertheless, its peak barely changes with temperature, suggesting a typical defect-related luminescence behavior and implying that the blue emission arises from the defect-related luminescence centers. In contrast, the red emission intensity displays almost no recognizable changes with temperature, and only the PL peak redshifts slightly to 670 nm with decreasing temperature, indicating a competitive relationship between the blue and red emission in the pristine CsSnCl3:Mn2+. To clarify the PL characteristics, PL decay curves were measured under an excitation wavelength of 375 nm (375 nm, 70 ps excitation pulses LASER), portrayed in Figure 3c,d. The PL decay curve for blue emission was well-fitted with a biexponential decay function, while the PL decay curve for red emission was well-fitted with a triexponential decay function [23]:
I ( t ) = I 0 + A 1 e x p ( t τ 1 ) + A 2 e x p ( t τ 2 ) + A 3 e x p ( t τ 3 )
where I0 represents the background level; τ1, τ2, and τ3 represent the lifetime of each exponential decay component, and A1, A2, and A3 denote the corresponding amplitudes, respectively. Therefore, the intensity-weighted average PL lifetimes are determined using ( A 1 τ 1 2 + A 2 τ 2 2 + A 3 τ 3 2 ) / ( A 1 τ 1 + A 2 τ 2 + A 3 τ 3 ) [24]. As demonstrated in Figure 3c, the blue emission exhibits a fast decay dynamic with a lifetime of 4.93 ns, while the red emission displays a slow decay behavior with a lifetime of 0.16 ms, which is five orders of magnitude longer than that of the blue emission. These findings adequately explain the competitive relationship between the defect-related luminescence centers and the Mn2+ state in the pristine CsSnCl3:Mn2+. Upon light excitation, the excitons lifted from the ground state to the excited state relax rapidly to the defect-related luminescence centers due to the fast decay dynamic, thereby significantly intensifying the blue emission.
The PL and PLE spectra of CsSnCl3:Mn2+ powders with and without the RTT process are depicted in Figure 4. Interestingly, it was observed that the RTT process led to a remarkable enhancement in red emission while significantly reducing the blue emission. Moreover, the intensity of the red emission increased with an increase in RTT time. Specifically, after the RTT process, the excitation peaks of CsSnCl3:Mn2+ at 355, 418, and 469 nm, corresponding to the 6A1(6S)→4T2(4D), 4T2(4G), and 4T1(4G) transitions of the Mn2+ ion [25], respectively, were observed, as shown in Figure 4. This strongly indicates that the RTT process effectively activates Mn2+ in CsSnCl3, leading to a significant enhancement in excited-state density, as revealed in Figure 4. Therefore, the enhancement in excited-state density is suggested to be responsible for the improved red emission. Interestingly, from Figure 4, it was found that the blue emission wavelength overlapped with the transition from 6A1(6S)→4T2(4G) and 4T1(4G) of the Mn2+ ion. This indicates that the reduction in blue emission may have resulted from the energy transfer from the defect-related luminescent state to the 6A1(6S)→4T2(4G) and 4T1(4G) transitions of the Mn2+ ion, which in turn contributes to the improvement in red PL intensity. Thus, the significant enhancement in red emission can be attributed to the enhanced excited-state density as well as energy transfer between the defect-related luminescent state and the Mn2+ state.
Figure 5 presents the PL decay curves of the CsSnCl3:Mn2+ powders with and without the RTT process. It was found that the average lifetime rapidly increased from 0.16 ms to 4.73 ms after the RTT process. Furthermore, with an increase in the RTT processing time from 60 s to 300 s, the average lifetime gradually increased from 4.73 ms to 6.81 ms, indicating a reduction in nonradiative recombination centers. This was confirmed via the XRD patterns of the CsSnCl3:Mn2+ powders with and without the RTT process, as shown in Figure 6. For the CsSnCl3:Mn2+ powders without the RTT process, the diffraction peaks at 22.10°, 23.95°, 30.269°, and 31.142° corresponded to (400), (020), (411), and (002) crystal planes of cubic CsSnCl3 (PDF# 71-2023), respectively. In addition, there were still small amounts of characteristic peaks of CsCl. However, after the RTT process, the diffraction peaks corresponding to (400), (020), (411), and (002) crystal planes of cubic CsSnCl3 became stronger and sharper with an increase in the RTT time from 60 s to 300 s. This strongly indicates an improved crystallinity of the CsSnCl3:Mn2+ powders after the RTT process, which coincided well with the increasing lifetime with an increase in the RTT time. Evidently, the RTT not only effectively activated the Mn2+ but also improved the crystallinity of the CsSnCl3:Mn2+ powders, thus reducing the nonradiative recombination centers and contributing to the enhancement in red emission. Raman spectra in Figure 7 show an array of symmetric (125–175 cm−1) and asymmetric (185–260 cm−1) stretching peaks for the SnCl3 group, and a peak at ~275 cm−1, attributable to symmetric Mn–Cl stretching.
To gain a deeper understanding of the PL characteristics, the temperature-dependent PL spectra of the CsSnCl3:Mn2+ powders with 300 s of RTT processing were measured, and the results are shown in Figure 8. It was observed that with a decrease in temperature, there was a significant increase in the red PL intensity along with a slight increase in the blue PL intensity. This is in contrast to the behavior observed in the CsSnCl3:Mn2+ powders without the RTT process, as shown in Figure 3b. This increase can be attributed to the energy transfer from the defect-related luminescent state to the 6A1(6S)→4T2(4G) and 4T1(4G) transitions of the Mn2+ ion. The temperature dependence of the integrated red PL intensity IPL(T) can be fitted using the Arrhenius equation given below [22]:
I P L ( T ) = I P L ( T 0 ) 1 + β exp E b / k B T
where IPL(T0) is the integrated PL intensity at 10 K, β is a constant related to the density of radiative recombination centers, kB is Boltzmann’s constant, and Eb is the exciton binding energy. Using this equation, the exciton binding energy Eb was empirically derived to be 31.6 meV, as shown in the inset of Figure 8. This further confirms that the red emission is from the Mn2+ state rather than from self-trapped excitons.
To evaluate the thermal stability of the CsSnCl3:Mn2+ powders, the temperature-dependent integrated PL intensities were monitored during thermal recycling. As shown in Figure 9, both samples experienced thermal quenching of PL as temperatures were increased from 25 to 165 °C. However, for pure CsSnCl3 powders, a reduction of over 90% in PL intensity was observed after the heating and cooling cycles. In contrast, CsSnCl3:Mn2+ powders with RTT processing showed a slight enhancement in PL intensity after the heating and cooling cycles. Clearly, the CsSnCl3:Mn2+ powders with RTT processing displayed superior thermal and structural stability.

4. Conclusions

We demonstrated a rapid synthesis method for SnF2-derived CsSnCl3:Mn2+ perovskite and investigated the effect of RTT on its PL properties. We observed that pristine CsSnCl3:Mn2+ exhibited a double luminescence peak structure, attributed to defect-related luminescent centers and the 4T1→6A1 transition of Mn2+. RTT was found to significantly reduce the blue emission and enhance the red emission intensity almost twofold compared to the pristine sample. This improvement in PL was suggested to arise from increased excited-state density, energy transfer between the defect-related state and the Mn2+ state, as well as a reduction of nonradiative recombination centers. Furthermore, we demonstrated that the Mn2+-doped samples after RTT exhibit excellent thermal stability. Our findings clearly demonstrate that RTT not only can avoid thermal decomposition of tin-based halide perovskite materials during the heating process but also improve the crystallinity of CsSnCl3:Mn2+ powders, and more importantly provides an alternative method for thermal activation of the Mn ion luminescence center of tin-based perovskite, which has implications for the future development of optoelectronic devices.

Author Contributions

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

Funding

This research was funded by the Guangdong Basic and Applied Basic Research Foundation (Grant Nos. 2020A1515010432), Project of Guangdong Province Key Discipline Scientific Research Level Improvement (Grant Nos. 2021ZDJS039, 2022ZDJS067), Special Innovation Projects of Guangdong Provincial Department of Education (Grant Nos. 2019KTSCX096, 2020KTSCX076), and the Special Fund for Science and Technology Innovation Strategy of Guangdong Province (Grant Nos. pdjh2023a0336).

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 conflict of interest.

References

  1. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed]
  2. Li, X.; Wu, Y.; Zhang, S.; Cai, B.; Gu, Y.; Song, J.; Zeng, H. CsPbX3 quantum dots for lighting and displays: Room-temperature synthesis, photoluminescence superiorities, underlying origins and white light-emitting diodes. Adv. Funct. Mater. 2016, 26, 2435–2445. [Google Scholar] [CrossRef]
  3. Lin, K.; Xing, J.; Quan, L.N.; Arquer, F.P.G.; Gong, X.W.; Lu, J.X.; Xie, L.Q.; Zhao, W.; Zhang, J.D.; Yan, C.Z.; et al. Perovskite light-emitting diodes with external quantum efficiency exceeding 20 percent. Nature 2018, 562, 245–248. [Google Scholar] [CrossRef]
  4. Liu, M.; Wan, Q.; Wang, H.; Carulli, F.; Sun, X.; Zheng, W.; Kong, L.; Zhang, Q.; Zhang, C.; Zhang, Q.; et al. Suppression of temperature quenching in perovskite nanocrystals for efficient and thermally stable light-emitting diodes. Nat. Photonics 2021, 15, 379–385. [Google Scholar] [CrossRef]
  5. Fan, Q.; Biesold-McGee, G.V.; Ma, J.; Xu, Q.; Pan, S.; Peng, J.; Lin, Z. Lead-free halide perovskite nanocrystals: Crystal structures, synthesis, stabilities, and optical properties. Angew. Chem. Int. Ed. 2020, 59, 1030–1046. [Google Scholar] [CrossRef] [PubMed]
  6. Li, X.; Gao, X.; Zhang, X.; Shen, X.; Lu, M.; Wu, J.; Shi, Z.; Colvin, V.L.; Hu, J.; Bai, X.; et al. Lead-Free Halide Perovskites for Light Emission: Recent Advances and Perspectives. Adv. Sci. 2021, 8, 2003334. [Google Scholar] [CrossRef]
  7. Jing, Y.; Liu, Y.; Zhao, J.; Xia, Z. Sb3+ doping-induced triplet self-trapped excitons emission in lead-free Cs2SnCl6 nanocrystals. J. Phys. Chem. Lett. 2019, 10, 7439–7444. [Google Scholar] [CrossRef]
  8. Lian, L.; Zheng, M.; Zhang, W.; Yin, L.; Du, X.; Zhang, P.; Zhang, X.; Gao, J.; Zhang, D.; Gao, L.; et al. Efficient and reabsorption-free radioluminescence in Cs3Cu2I5 nanocrystals with self-trapped excitons. Adv. Sci. 2020, 7, 2000195. [Google Scholar] [CrossRef]
  9. Cong, M.; Zhang, Q.; Yang, B.; Chen, J.; Xiao, J.; Zheng, D.; Zheng, T.; Zhang, R.; Qing, G.; Zhang, C.; et al. Bright triplet self-trapped excitons to dopant energy transfer in halide double-perovskite nanocrystals. Nano Lett. 2021, 21, 8671–8678. [Google Scholar] [CrossRef]
  10. Su, B.; Jin, J.; Peng, Y.; Molokeev, M.S.; Yang, X.; Xia, Z. Zero-Dimensional Organic Copper(I) Iodide Hybrid with High Anti-Water Stability for Blue-Light-Excitable Solid-State Lighting. Adv. Opt. Mater. 2022, 10, 2102619. [Google Scholar] [CrossRef]
  11. 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]
  12. Liu, Q.; Yin, J.; Zhang, B.-B.; Chen, J.-K.; Zhou, Y.; Zhang, L.-M.; Wang, L.-M.; Zhao, Q.; Hou, J.; Shu, J.; et al. Theory-guided synthesis of highly luminescent colloidal cesium tin halide perovskite nanocrystals. J. Am. Chem. Soc. 2021, 143, 5470–5480. [Google Scholar] [CrossRef]
  13. Chiara, R.; Ciftci, Y.O.; Queloz, V.I.E.; Nazeeruddin, M.K.; Grancini, G.; Malavasi, L. Green-emitting lead-free Cs4SnBr6 zero-dimensional perovskite nanocrystals with improved air stability. J. Phys. Chem. Lett. 2020, 11, 618–623. [Google Scholar] [CrossRef] [PubMed]
  14. 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] [PubMed]
  15. Zhang, X.; Wang, H.; Wang, S.; Hu, Y.; Liu, X.; Shi, Z.; Colvin, V.L.; Wang, S.; Yu, W.W.; Zhang, Y. Room temperature synthesis of all inorganic lead-free zero-dimensional Cs4SnBr6 and Cs3KSnBr6 perovskites. Inorg. Chem. 2019, 59, 533–538. [Google Scholar] [CrossRef] [PubMed]
  16. Tan, L.; Wang, W.; Li, Q.; Luo, Z.; Zou, C.; Tang, M.; Zhang, L.; He, J.; Quan, Z. Colloidal syntheses of zero-dimensional Cs4SnX6 (X = Br, I) nanocrystals with high emission efficiencies. Chem. Commun. 2020, 56, 387–390. [Google Scholar] [CrossRef]
  17. Xu, K.; Wei, Q.; Wang, H.; Yao, B.; Zhou, W.; Gao, R.; Chen, H.; Li, H.; Wang, J.; Ning, Z. The 3D-structure-mediated growth of zero-dimensional Cs4SnX6 nanocrystals. Nanoscale 2022, 14, 2248–2255. [Google Scholar] [CrossRef]
  18. 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]
  19. Dawson, M.; Ribeiro, C.; Morelli, M.R. MnCl2 doping increases phase stability of tin halide perovskites. Mater. Sci. Semicond. Process. 2021, 132, 105908. [Google Scholar] [CrossRef]
  20. Hou, L.; Zhu, Y.; Zhu, J.; Gong, Y.; Li, C. Mn-doped 2D Sn-based perovskites with energy transfer from self-trapped excitons to dopants for warm white light-emitting diodes. J. Mater. Chem. C 2020, 8, 8502–8506. [Google Scholar] [CrossRef]
  21. Voloshinovskii, A.S.; Myagkota, S.V.; Pidzyrailo, N.S.; Tokarivskii, M.V. Luminescence and structural transformations of CsSnCl3 Crystals. J. Appl. Spectrosc. 1994, 60, 226–228. [Google Scholar] [CrossRef]
  22. Lin, Z.; Wang, A.; Huang, R.; Wu, H.; Song, J.; Lin, Z.; Hou, D.; Zhang, Z.; Guo, Y.; Lan, S. 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]
  23. Lin, K.; Liou, S.; Chen, W.; Wu, C.; Lin, G.; Chang, Y. Tunable and stable UV-NIR photoluminescence from annealed SiOx with Si nanoparticles. Opt. Express 2013, 21, 23416–23424. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, Z.; Lin, Z.; Guo, Y.; Wu, H.; Song, J.; Zhang, Y.; Zhang, W.; Li, H.; Hou, D.; Huang, R. Effect of a-SiCxNy:H encapsulation on the stability and photoluminescence property of CsPbBr3 quantum dots. Nanomaterials 2023, 13, 1228. [Google Scholar] [CrossRef]
  25. Liu, W.-R.; Chiu, Y.-C.; Yeh, Y.-T.; Jang, S.-M.; Chen, T.-M. Luminescence and Energy Transfer Mechanism in Ca10K(PO4)7: Eu2+, Mn2+ Phosphor. J. Electrochem. Soc. 2009, 156, J165. [Google Scholar] [CrossRef]
Figure 1. Schematic showing the method for (a) synthesizing Mn2+-doped CsSnCl3 and (b) RTT process.
Figure 1. Schematic showing the method for (a) synthesizing Mn2+-doped CsSnCl3 and (b) RTT process.
Materials 16 04027 g001
Figure 2. SEM image and EDS elemental maps of Cs, Sn, Mn, Cl, and F for a typical Mn2+-doped CsSnCl3 powder (a) before RTT process and (b) after RTT process for 300 s, respectively.
Figure 2. SEM image and EDS elemental maps of Cs, Sn, Mn, Cl, and F for a typical Mn2+-doped CsSnCl3 powder (a) before RTT process and (b) after RTT process for 300 s, respectively.
Materials 16 04027 g002
Figure 3. (a) PL spectrum from the Mn2+-doped CsSnCl3 powders without RTT process. Inset shows PL spectrum from the CsSnCl3 powders. (b) Temperature-dependent PL spectra of the pristine CsSnCl3:Mn2+ measured in the range of 78 to 298 K. Time-resolved PL decay trace of (c) blue PL recorded at 450 nm and (d) red PL recorded at 640 nm in the pristine CsSnCl3:Mn2+.
Figure 3. (a) PL spectrum from the Mn2+-doped CsSnCl3 powders without RTT process. Inset shows PL spectrum from the CsSnCl3 powders. (b) Temperature-dependent PL spectra of the pristine CsSnCl3:Mn2+ measured in the range of 78 to 298 K. Time-resolved PL decay trace of (c) blue PL recorded at 450 nm and (d) red PL recorded at 640 nm in the pristine CsSnCl3:Mn2+.
Materials 16 04027 g003
Figure 4. PL and PLE spectra of CsSnCl3:Mn2+ powders with and without RTT process, respectively.
Figure 4. PL and PLE spectra of CsSnCl3:Mn2+ powders with and without RTT process, respectively.
Materials 16 04027 g004
Figure 5. Time-resolved PL decay trace of (a) red PL recorded at corresponding PL peak and (b) blue PL recorded at 454 nm in the CsSnCl3:Mn2+ powders with and without RTT process, measured under an excitation wavelength of 375 nm (375 nm, 70 ps excitation pulses LASER).
Figure 5. Time-resolved PL decay trace of (a) red PL recorded at corresponding PL peak and (b) blue PL recorded at 454 nm in the CsSnCl3:Mn2+ powders with and without RTT process, measured under an excitation wavelength of 375 nm (375 nm, 70 ps excitation pulses LASER).
Materials 16 04027 g005
Figure 6. XRD patterns obtained for the CsSnCl3:Mn2+ powders with and without RTT process, The symbol “*”represents the corresponding crystal plane.
Figure 6. XRD patterns obtained for the CsSnCl3:Mn2+ powders with and without RTT process, The symbol “*”represents the corresponding crystal plane.
Materials 16 04027 g006
Figure 7. Raman spectra of CsSnCl3:Mn2+ powders with and without RTT process.
Figure 7. Raman spectra of CsSnCl3:Mn2+ powders with and without RTT process.
Materials 16 04027 g007
Figure 8. Temperature-dependent PL spectra of the CsSnCl3:Mn2+ powders with RTT processing for 300 s measured in the range of 78 to 298 K. Inset shows the integrated red PL intensities at different temperatures (orange solid symbols) and the fitting of the experimental data (red curve).
Figure 8. Temperature-dependent PL spectra of the CsSnCl3:Mn2+ powders with RTT processing for 300 s measured in the range of 78 to 298 K. Inset shows the integrated red PL intensities at different temperatures (orange solid symbols) and the fitting of the experimental data (red curve).
Materials 16 04027 g008
Figure 9. Heating and cooling cycling measurements at various temperatures: (a) pure CsSnCl3; (b) CsSnCl3:Mn2+ powers with RTT processing for 300 s.
Figure 9. Heating and cooling cycling measurements at various temperatures: (a) pure CsSnCl3; (b) CsSnCl3:Mn2+ powers with RTT processing for 300 s.
Materials 16 04027 g009
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

Xu, J.; Wu, H.; Lu, X.; Huang, Y.; Chen, J.; Zhou, W.; Lin, Z.; Song, J.; Li, H.; Huang, R. Synthesis and Improved Photoluminescence of SnF2-Derived CsSnCl3-SnF2:Mn2+ Perovskites via Rapid Thermal Treatment. Materials 2023, 16, 4027. https://doi.org/10.3390/ma16114027

AMA Style

Xu J, Wu H, Lu X, Huang Y, Chen J, Zhou W, Lin Z, Song J, Li H, Huang R. Synthesis and Improved Photoluminescence of SnF2-Derived CsSnCl3-SnF2:Mn2+ Perovskites via Rapid Thermal Treatment. Materials. 2023; 16(11):4027. https://doi.org/10.3390/ma16114027

Chicago/Turabian Style

Xu, Jisheng, Haixia Wu, Xinye Lu, Yaqian Huang, Jianni Chen, Wendi Zhou, Zewen Lin, Jie Song, Hongliang Li, and Rui Huang. 2023. "Synthesis and Improved Photoluminescence of SnF2-Derived CsSnCl3-SnF2:Mn2+ Perovskites via Rapid Thermal Treatment" Materials 16, no. 11: 4027. https://doi.org/10.3390/ma16114027

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