Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Synergistic luminescence effect and high-pressure optical properties of CsPbBr2Cl@EuMOFs nanocomposites

Open Access Open Access

Abstract

Metal-organic frameworks (MOFs) are a class of highly porous materials that have garnered significant attention in the field of optoelectronics due to their exceptional properties. In this study, CsPbBr2Cl@EuMOFs nanocomposites were synthesized using a two-step method. The fluorescence evolution of the CsPbBr2Cl@EuMOFs was investigated under high pressure, revealing a synergistic luminescence effect between CsPbBr2Cl and Eu3+. The study found that the synergistic luminescence of CsPbBr2Cl@EuMOFs remains stable even under high pressure, and there is no energy transfer among different luminous centers. These findings provide a meaningful case for future research on nanocomposites with multiple luminescent centers. Additionally, CsPbBr2Cl@EuMOFs exhibit a sensitive color-changing mechanism under high pressure, making them a promising candidate for pressure calibration via the color change of the MOF materials.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

As a new type of semiconducting materials in optoelectronics and photovoltaics [14], all-inorganic CsPbX3 (X = Cl, Br, I) perovskite quantum dots (PeQDs) have shown exceptional linear optical properties like narrow full width at half maximum (FWHM) and composition-dependent bandgaps over the entire visible spectral region [59]. Recently, many studies indicate that direct doping of rare earth ions is an effective way to regulate the luminescence properties of PeQDs [1012]. By replacing Pb2+ in PeQDs with rare earth ions, the structural stability, as well as the overall luminous efficiency of PeQDs can be greatly improved, such as the fluorescence quantum yield of Yb3+ doped CsPbCl3 PeQDs can reach about 200% due to the efficient energy transfer from PeQDs to Yb ions [1315]. However, the tolerance of PeQDs to rare earth ions varies greatly depending on the halogen used [16], which significantly restricts the development of rare earth ions doping in PeQDs with all halogens.

To achieve a stable doping concentration of rare earth ions and combine it with the excellent optical properties of PeQDs, nanocomposite preparation is a feasible option. The common nanocomposites includes rare earth/PeQDs co-doped glass [1719], MOFs [2022], upconversion nanoparticles (UCNPs) [2325] and so on. Co-doping rare earth ions and PeQDs in composite materials can result in exceptional multicolor emission properties and long-lasting luminescence stability. The luminous mechanism of nanocomposites can be classified as energy transfer or synergistic luminescence based on the coupling of luminescent centers [2628]. In general, regardless of whether PeQDs have experienced energy transfer to rare earth ions, the fluorescence emission of rare earth/PeQDs nanocomposites is almost identical, which can be regarded as beneficial optical features and provides additional potential for diverse practical applications.

As one of the nanocomposites, MOFs possess versatile advantages, including different coordination geometry and large aperture, which enable them to support multiple luminescent centers simultaneously. Recently, MOFs have attracted much attention with special utilization as host matrices for photoluminescence [29,30]. In particular, studies on lanthanide MOFs (Ln-MOFs) are extensively reported and many guest members including the carbon dots, metal nanoparticles, and dye molecules, have also been incorporated into MOFs to design new multifunctional MOF-based nanocomposites [3135]. In order to exploit the novel optical properties of MOFs, the diamond anvil cell (DAC) can be used as an effective tool. The interatomic distance and electronic configuration of nanocomposites can be efficiently modified under high pressure, increasing their electronic structure and photoelectric properties. The photoelectric characteristics of MOFs materials, particularly Ln-MOFs, are rapidly being discovered as high-pressure technology advances [36]. However, the QDs/Ln-MOFs with multiple emission centers have better photoelectric performance, but their optical properties under high pressure have not been studied yet.

Herein, we synthesized the CsPbBr2Cl@EuMOFs nanocomposites by two-step method. The CsPbBr2Cl@EuMOFs is a rod-like crystal structure with high crystallization and multichannel fluorescence emission. Under high pressure, both CsPbBr2Cl PeQDs and Eu3+ have a common optics behavior that fluorescence enhancement and attenuation in nanocomposites. By comparing with the fluorescence properties of EuMOFs under high pressure, the result is confirmed that the CsPbBr2Cl@EuMOFs have a significantly stable synergistic luminescence property, despite the fact that the intermolecular interaction distance has been compressed under high pressure. Furthermore, the spot color of CsPbBr2Cl@EuMOFs varies with pressure, indicating that this phenomenon has the potential to be used as an optical application to calibrate pressure value.

2. Results and discussion

The CsPbBr2Cl@EuMOFs were synthesized by a two-step method based on the previous reports, which could be seen in Supplement 1 [37,38]. As shown in Fig. 1(a), the TEM images exhibit that rod-like standard CsPbBr2Cl@EuMOFs are highly uniform crystals with the width of 0.35 µm and the length of 1.75 µm. As to be expected, numerous CsPbBr2Cl PeQDs are grown in-situ on the surface of the nanocomposites, and the TEM image is shown in Fig. 1(b). The HR-TEM image (Fig. 1(c)) exhibits a well-crystallized structure with a lattice fringe spacing of 0.28 nm, which is consistent with the spacing of the (200) plane in cubic CsPbBr2Cl PeQDs. The prepared CsPbBr2Cl@EuMOFs elemental mapping, which is depicted in Fig. 1(e)–(j), reveals a uniform distribution of the elements O, Eu, Cs, Pb, Br, and Cl. Figure 2(d) depicts the XRD patterns of the CsPbBr2Cl@EuMOFs. The triangle in the figure represents the EuMOF diffraction peaks, and the main peak position corresponds to previous research findings [37]. The red star represents the CsPbBr2Cl PeQDs diffraction peak, which is consistent with the standard diffraction results (PDF#75-0412). The addition of Cl ions causes a blue shift in the peak position of the CsPbBr2Cl PeQDs compared to the crystal plane diffraction peak of CsPbBr3. The results show that CsPbBr2Cl@EuMOFs have good crystallinity and are composite materials, which can be verified by the as-observed diffraction patterns.

 figure: Fig. 1.

Fig. 1. (a) TEM images, (b) TEM magnification images, and (c) HR-TEM images of the CsPbBr2Cl@EuMOFs. (d) The XRD diffraction patterns of CsPbBr2Cl@EuMOFs. (e-j) EDX spectra of CsPbBr2Cl@EuMOFs nanocomposites, testing element including O, Eu, Cs, Pb, Br, Cl.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. (a) Structural composition model of the CsPbBr2Cl@EuMOFs. (b) PL spectra of the CsPbBr2Cl@EuMOFs and the EuMOFs at 405 nm excitation. (c) PL spectra of the CsPb(BrxCl1-x) 3@EuMOFs vary with PbBr2 doping concentration.

Download Full Size | PDF

The basic composition of the CsPbBr2Cl@EuMOFs nanocomposites is shown in Fig. 2(a). In brief, the framework of the CsPbBr2Cl@EuMOFs is composed of Eu-doped benzene ring and carbon chain, and the CsPbBr2Cl PeQDs are filled in the voids of EuMOFs. In the framework, Pb2+ preferentially binds to O atoms in the framework rather than directly to Eu ions during the growth of nanocomposites [3941]. The carbon skeleton serves as the foundation for the nanocomposite structure, and combines with Eu ions and PeQDs. The PL spectra of EuMOFs and CsPbBr2Cl@EuMOFs at room temperature are shown in Fig. 2(b). Under the excitation of 405 nm laser, the fluorescence emission of the EuMOFs at 606 nm, 634 nm and 706 nm are electron transitions from 5D07F1, 5D07F2 and 5D07F4, respectively. In contrast, the CsPbBr2Cl@EuMOFs exhibit strong blue and red fluorescence emission, and the blue fluorescence emission at 485 nm is mainly the band-edge exciton recombination of CsPbBr2Cl PeQDs. The effect of the Pb2+ doping concentration on the luminescence of CsPbBr2Cl@EuMOFs is shown in Fig. 2(d). The concentration of Br ions increases with increasing PbBr2 doping concentration, resulting in emission wavelength redshift due to anion imbalance at a certain concentration of Cs + . Obviously, the CsPbBr2Cl@EuMOFs has the best luminescence performance when the optimal concentration of Pb2+ is 0.5 mmol.

Considering the unique fluorescence characteristics of the CsPbBr2Cl@EuMOFs nanocomposites, high-pressure PL experiments were carried out to trace the pressure-induced optical evolution. The PL intensity change of the CsPbBr2Cl@EuMOFs nanocomposites during compression process is shown in Fig. 3(a) and Fig. 3(b), respectively. With the increase of pressure, the PL intensity of the CsPbBr2Cl PeQDs in nanocomposites increases temporarily before 0.9 GPa, then gradually decreases until it disappears at 2.3 GPa. During PL intensity variations, its fluorescence wavelength maintains a steady red shift. In comparison, the PL intensity of the Eu3+ in nanocomposites increases before 5.7 GPa, then decreases gradually with pressure increasing. The fluorescence wavelength of the the Eu3+ in nanocomposites changes little during pressurization. The relative changes of each fluorescence peak in the CsPbBr2Cl@EuMOFs are shown in Fig. 3(c) and the inset. The maximum PL intensity of CsPbBr2Cl PeQDs and Eu3+ in nanocomposites during the pressurization process is 1.3 times and 2.9 times that of their initial fluorescence, respectively. Notably, the pink part in Fig. 3(c) represents the overlapping region of the decreasing PL intensity of CsPbBr2Cl PeQDs and the rising PL intensity of Eu3+ in nanocomposites, which is similar to the fluorescence change of Eu3+ directly doped perovskite under pressure [42]. The peak position changes of the CsPbBr2Cl@EuMOFs are shown in Fig. 3(d). Under high pressure, the initial fluorescence wavelength of the CsPbBr2Cl PeQDs is 485 nm and gradually redshifts to 489 nm in nanocomposites. The fluorescence wavelength of Eu3+ in nanocomposites changes only slightly, and each wavelength is redshifted by around 0.5 nm after compression.

 figure: Fig. 3.

Fig. 3. (a) (b) Evolution of PL optical properties of CsPbBr2Cl@EuMOFs under high pressure, and (c) the changes in relative PL intensity of corresponding peaks, and (d) the changes in relative position of corresponding peaks. The fluctuation of the PL intensity caused by errors in experimental measurement are reasonably optimized.

Download Full Size | PDF

To investigate the luminescence mechanism further, we took the CsPbBr2Cl PeQDs in CsPbBr2Cl@EuMOFs nanocomposites as an example to specifically study its time-resolved fluorescence decay. Figure 4(a) shows the PL intensity changes of the CsPbBr2Cl PeQDs in nanocomposites under high pressure, and Fig. 4(b) shows its fluorescence lifetime changes under corresponding pressure. Before 0.9 GPa, the increase of the fluorescence lifetime can be attributed to the defect passivation at a low pressure, which significantly reduces the non-radiative recombination rate and increases the PL intensity of CsPbBr2Cl PeQDs [43,44]. As the pressure further increases, the symmetrical lattice structure of CsPbBr2Cl PeQDs gradually distorts, resulting in the prolong of carrier diffusion length and the weakening of PL intensity [45,46]. The fluorescence lifetimes of CsPbBr2Cl PeQDs increase from 4.3 ns to 6.7 ns under compression process, which is greatly beneficial to promote efficiently the photon-generated carrier charge separation in photovoltaic applications [47]. Notably, efficient non-radiative energy transfer usually results in a reduced fluorescence lifetime of the donor, implying that the luminescence of the CsPbBr2Cl@EuMOFs may not be a process of energy transfer. As shown in Fig. 4(c) and the inset, the PL intensity of EuMOFs enhances remarkably before 5.1 GPa and then decreases gradually. By comparing Fig. 4(d) with Fig. 3(c), it is clear that the PL changes between the CsPbBr2Cl@EuMOFs and the EuMOFs under high pressure are nearly identical. As for enhanced fluorescence of Eu3+, a similar pattern of results is obtained in TbMOFs [36]. The increased fluorescence of Eu3+ could be due to the pressure optimizes the energy transfer process from ligand H3BTC to Eu3+. As a result, the luminescence mechanism of CsPbBr2Cl@EuMOFs is synergistic luminescence, and the luminescence mechanism is extremely stable even when the distance between the luminescent centers is compressed by pressure.

 figure: Fig. 4.

Fig. 4. The changes in (a) PL intensity and (b) PL lifetime decay of CsPbBr2Cl PeQDs in nanocomposites under high pressure. (c) Evolution of PL optical properties of the EuMOFs under high pressure, and (d) statistics of PL intensity corresponding to peak values. The fluctuation of the PL intensity of EuMOFs caused by errors in experimental measurement are reasonably optimized.

Download Full Size | PDF

The synergistic luminescence mechanism of the CsPbBr2Cl@EuMOFs is shown in the Fig. 5. Under excitation, the PL emission of the CsPbBr2Cl PeQDs and the Eu3+ in nanocomposites is mutual independent, and does not interact with each other. Under high pressure, the fluorescence attenuation of CsPbBr2Cl PeQDs overlaps with the fluorescence enhancement of Eu3+ on the timescale, while the original luminescence mechanism remains unchanged. Compared with Eu3+ directly doped PeQDs, the contact distance between PeQDs and Eu3+ in CsPbBr2Cl@EuMOFs may not match the requirement of Förster resonance energy transfer even under high pressure. The findings show that MOF-based nanocomposites integrate the distinct emission properties of each luminescence center and are not constrained by coupling distance. Furthermore, the CsPbBr2Cl@EuMOFs can precisely replicate the excellent optical properties of the rare earth ions directly doped PeQDs, solving the problem of low rare earth ion doping content in PeQDs.

 figure: Fig. 5.

Fig. 5. The luminescence mechanism of the CsPbBr2Cl@EuMOFs under high pressure.

Download Full Size | PDF

We recorded the spot color changes of the CsPbBr2Cl@EuMOFs with value of pressure, as shown in Fig. 6. The spot color on the DAC plane corresponds to the PL intensity change of the nanocomposite under high pressure (Fig. 3). The CsPbBr2Cl@EuMOFs exhibit a sensitive color-changing mechanism under high pressure, indicating that the CsPbBr2Cl@EuMOFs can be used as a color scale for pressure detection. Figure 7 shows the reusable property of the nanocomposites. Due to the inherent ionic structure, the PeQDs in MOF matrix can be destroyed by polar solvents impregnation, thus quenching the luminescence of the PeQDs. Then, highly luminescent PeQDs can be formed quickly and simply by using a halide salt trigger that reacts with the Pb@EuMOFs, thus the recovery of fluorescence is achieved. Accordingly, the peak positions of the emission intensities remain nearly unchanged after multiple cycles processes. Through our conversion strategy, the nanocomposites can be selectively used to avoid the problem of fluorescence weakening caused by long time placement of perovskites in nanocomposites, which is conducive to its further application.

 figure: Fig. 6.

Fig. 6. The spot color changes of the CsPbBr2Cl@EuMOFs under high pressure.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. The photos of CsPbBr2Cl@EuMOFs nanocomposites during the impregnation-recovery process reacted with Ethanol and CsX(Br/Cl). The variation of emission intensities, peak position, and FWHM values of the CsPbBr2Cl PeQDs in nanocomposites during the impregnation-recovery cycles for several cycle numbers.

Download Full Size | PDF

3. Conclusions

In conclusion, we have successfully developed the CsPbBr2Cl@EuMOFs nanocomposites with multichannel emission and displayed its excellent fluorescence emission property.Under high pressure, the CsPbBr2Cl@EuMOFs nanocomposites have unique synergistic luminescence properties, and there is no efficient energy transfer occuring between the CsPbBr2Cl PeQDs and the Eu3+. The optical responses of PeQDs and Eu3+ in the CsPbBr2Cl@EuMOFs are different under high pressure, resulting in various color changes of their composite fluorescence. Through the color variations, this phenomenon can be viewed as a potential use for pressure calibration.

Funding

National Natural Science Foundation of China (61307067, 61905106); Taishan Scholar Project of Shandong Province (tsqn201812098, ts201511055); Natural Science Foundation of Shandong Province (ZR2019MF057, ZR2019MA066).

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

Supplemental document

See Supplement 1 for supporting content.

References

1. R. Begum, M. R. Parida, A. L. Abdelhady, B. Murali, N. M. Alyami, G. H. Ahmed, M. N. Hedhili, O. M. Bakr, and O. F. Mohammed, “Engineering interfacial charge transfer in CsPbBr3 perovskite nanocrystals by heterovalent doping,” J. Am. Chem. Soc. 139(2), 731–737 (2017). [CrossRef]  

2. Q. A. Akkerman, G. Rainò, M. V. Kovalenko, and L. Manna, “Genesis, challenges and opportunities for colloidal lead halide perovskite nanocrystals,” Nat. Mater. 17(5), 394–405 (2018). [CrossRef]  

3. V. M. Burlakov, Y. Hassan, M. Danaie, H. J. Snaith, and A. Goriely, “Competitive nucleation mechanism for CsPbBr3 perovskite nanoplatelet growth,” J. Phys. Chem. Lett. 11(16), 6535–6543 (2020). [CrossRef]  

4. P. Ijaz, M. Imran, M. M. Soares, H. C. Tolentino, B. Martín-García, C. Giannini, I. Moreels, L. Manna, and R. Krahne, “Composition-, size-, and surface functionalization-dependent optical properties of lead bromide perovskite nanocrystals,” J. Phys. Chem. Lett. 11(6), 2079–2085 (2020). [CrossRef]  

5. Z. Liu, Y. Bekenstein, X. Ye, S. C. Nguyen, J. Swabeck, D. Zhang, S.-T. Lee, P. Yang, W. Ma, and A. P. Alivisatos, “Ligand mediated transformation of cesium lead bromide perovskite nanocrystals to lead depleted Cs4PbBr6 nanocrystals,” J. Am. Chem. Soc. 139(15), 5309–5312 (2017). [CrossRef]  

6. L. Protesescu, S. Yakunin, M. I. Bodnarchuk, F. Krieg, R. Caputo, C. H. Hendon, R. X. Yang, A. Walsh, and M. V. Kovalenko, “Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): novel optoelectronic materials showing bright emission with wide color gamut,” Nano Lett. 15(6), 3692–3696 (2015). [CrossRef]  

7. J. Song, J. Li, X. Li, L. Xu, Y. Dong, and H. Zeng, “Quantum dot light-emitting diodes based on inorganic perovskite cesium lead halides (CsPbX3),” Adv. Mater. 27(44), 7162–7167 (2015). [CrossRef]  

8. Y. Li, H. Huang, Y. Xiong, S. V. Kershaw, and A. L. Rogach, “Revealing the formation mechanism of CsPbBr3 perovskite nanocrystals produced via a slowed-down microwave-assisted synthesis,” Angew. Chem. Int. Ed. 57(20), 5833–5837 (2018). [CrossRef]  

9. Q. A. Akkerman, V. D’Innocenzo, S. Accornero, A. Scarpellini, A. Petrozza, M. Prato, and L. Manna, “Tuning the optical properties of cesium lead halide perovskite nanocrystals by anion exchange reactions,” J. Am. Chem. Soc. 137(32), 10276–10281 (2015). [CrossRef]  

10. G. Pan, X. Bai, D. Yang, X. Chen, P. Jing, S. Qu, L. Zhang, D. Zhou, J. Zhu, and W. Xu, “Doping lanthanide into perovskite nanocrystals: highly improved and expanded optical properties,” Nano Lett. 17(12), 8005–8011 (2017). [CrossRef]  

11. G. H. Debnath, B. P. Bloom, S. Tan, and D. H. Waldeck, “Room temperature doping of Ln3+ in perovskite nanoparticles: a halide exchange mediated cation exchange approach,” Nanoscale 14(16), 6037–6051 (2022). [CrossRef]  

12. X. Zhang, Y. Zhang, X. Zhang, W. Yin, Y. Wang, H. Wang, M. Lu, Z. Li, Z. Gu, and W. Y. William, “Yb3+ and Yb3+/Er3+ doping for near-infrared emission and improved stability of CsPbCl3 nanocrystals,” J. Mater. Chem. C 6(37), 10101–10105 (2018). [CrossRef]  

13. A. Lesage, M. van der Laan, L. Gomez, and T. Gregorkiewicz, “Substitutional Doping of Yb3+ in CsPbBrxCl3–x Nanocrystals,” J. Phys. Chem. C 124(11), 6413–6417 (2020). [CrossRef]  

14. T. J. Milstein, D. M. Kroupa, and D. R. Gamelin, “Picosecond quantum cutting generates photoluminescence quantum yields over 100% in ytterbium-doped CsPbCl3 nanocrystals,” Nano Lett. 18(6), 3792–3799 (2018). [CrossRef]  

15. K. Xu, D. Chen, D. Huang, and H. Zhu, “Thermally stable emission from Yb3+-doped CsPbCl3 nanocrystals,” J. Lumin. 240, 118464 (2021). [CrossRef]  

16. Q. Li, Y. Liu, P. Chen, J. Hou, Y. Sun, G. Zhao, N. Zhang, J. Zou, J. Xu, and Y. Fang, “Excitonic luminescence engineering in tervalent-europium-doped cesium lead halide perovskite nanocrystals and their temperature-dependent energy transfer emission properties,” J. Phys. Chem. C 122(50), 29044–29050 (2018). [CrossRef]  

17. Y. Cheng, C. Shen, L. Shen, W. Xiang, and X. Liang, “Tb3+, Eu3+ co-doped CsPbBr3 QDs glass with highly stable and luminous adjustable for white LEDs,” ACS Appl. Mater. Interfaces 10(25), 21434–21444 (2018). [CrossRef]  

18. C. Wang, H. Lin, X. Xiang, Y. Cheng, Q. Huang, Y. Gao, X. Cui, and Y. Wang, “CsPbBr3/EuPO4 dual-phase devitrified glass for highly sensitive self-calibrating optical thermometry,” J. Mater. Chem. C 6(37), 9964–9971 (2018). [CrossRef]  

19. P. Li, Y. Duan, Y. Lu, A. Xiao, Z. Zeng, S. Xu, and J. Zhang, “Nanocrystalline structure control and tunable luminescence mechanism of Eu-doped CsPbBr3 quantum dot glass for WLEDs,” Nanoscale 12(12), 6630–6636 (2020). [CrossRef]  

20. C. Zhang, B. Wang, W. Li, S. Huang, L. Kong, Z. Li, and L. Li, “Conversion of invisible metal-organic frameworks to luminescent perovskite nanocrystals for confidential information encryption and decryption,” Nat. Commun. 8(1), 1138 (2017). [CrossRef]  

21. H. He, Y. Cui, B. Li, B. Wang, C. Jin, J. Yu, L. Yao, Y. Yang, B. Chen, and G. Qian, “Confinement of Perovskite-QDs within a Single MOF Crystal for Significantly Enhanced Multiphoton Excited Luminescence,” Adv. Mater. 31(6), 1806897 (2019). [CrossRef]  

22. K. Wang, N. Li, J. Zhang, Z. Zhang, and F. Dang, “Size-selective QD@ MOF core-shell nanocomposites for the highly sensitive monitoring of oxidase activities,” Biosens. Bioelectron. 87, 339–344 (2017). [CrossRef]  

23. W. Zheng, P. Huang, Z. Gong, D. Tu, J. Xu, Q. Zou, R. Li, W. You, J.-C. G. Bünzli, and X. Chen, “Near-infrared-triggered photon upconversion tuning in all-inorganic cesium lead halide perovskite quantum dots,” Nat. Commun. 9(1), 3462 (2018). [CrossRef]  

24. H. Che, P. Dang, Y. Wei, Z. Cheng, and J. Lin, “Encapsulation of lead halide perovskite quantum dots in mesoporous NaYF4 matrices with enhanced stability for anti-counterfeiting,” Dalton Trans. 50(29), 10299–10309 (2021). [CrossRef]  

25. M. Li, W. Liu, T. Yang, Q. Xu, H. Mu, J. Han, K. Cao, M. Jiao, M. Liu, and S. Zhang, “Multi-color UCNPs/CsPb (Br1-xIx)3 for upconversion luminescence and dual-modal anticounterfeiting,” Opt. Express 31(2), 2956–2966 (2023). [CrossRef]  

26. K. Wei, P. Li, Y. Duan, S. Zhang, L. Chen, S. Xu, and J. Zhang, “Temperature-dependent color-tunable luminescence in CsPbBr3: Dy3+ glass ceramic,” J. Non-Cryst. Solids 570, 121022 (2021). [CrossRef]  

27. Y. Yu, G. Shao, L. Ding, H. Zhang, X. Liang, J. Liu, and W. Xiang, “Ultra-stable Eu3+-doped CsPbCl2Br perovskite quantum dots glass for optical temperature sensing,” J. Rare Earths 39(12), 1497–1505 (2021). [CrossRef]  

28. C. Lu, Y. Duan, P. Li, Y. Lu, S. Xu, and J. Zhang, “Polychromatic tunable luminescence of Eu3+ doped CsPbBr3 quantum dot glass ceramic induced by mechanical crystallization,” Ceram. Int. 48(10), 13826–13832 (2022). [CrossRef]  

29. C. Rösler and R. A. Fischer, “Metal–organic frameworks as hosts for nanoparticles,” CrystEngComm. 17(2), 199–217 (2015). [CrossRef]  

30. Y. Cui, B. Li, H. He, W. Zhou, B. Chen, and G. Qian, “Metal–organic frameworks as platforms for functional materials,” Acc.Chem.Res. 49(3), 483–493 (2016). [CrossRef]  

31. Y. Dong, J. Cai, Q. Fang, X. You, and Y. Chi, “Dual-emission of lanthanide metal–organic frameworks encapsulating carbon-based dots for ratiometric detection of water in organic solvents,” Anal. Chem. 88(3), 1748–1752 (2016). [CrossRef]  

32. R. Kaur, K. Vellingiri, K.-H. Kim, A. Paul, and A. Deep, “Efficient photocatalytic degradation of rhodamine 6 G with a quantum dot-metal organic framework nanocomposite,” Chemosphere 154, 620–627 (2016). [CrossRef]  

33. T. Xia, T. Song, Y. Cui, Y. Yang, and G. Qian, “A dye encapsulated terbium-based metal–organic framework for ratiometric temperature sensing,” Dalton Trans. 45(46), 18689–18695 (2016). [CrossRef]  

34. G. W. Xu, Y. P. Wu, W. W. Dong, J. Zhao, X. Q. Wu, D. S. Li, and Q. Zhang, “A multifunctional Tb-MOF for highly discriminative sensing of Eu3+/Dy3+ and as a catalyst support of Ag nanoparticles,” Small 13(22), 1602996 (2017). [CrossRef]  

35. Z.-G. Gu and J. Zhang, “Epitaxial growth and applications of oriented metal–organic framework thin films,” Coord.Chem.Rev. 378, 513–532 (2019). [CrossRef]  

36. Y. Wang, X. Yang, C. Liu, Z. Liu, Q. Fang, F. Bai, S. Wang, X. Hou, B. Feng, and B. Chen, “Maximized Green Photoluminescence in Tb-Based Metal–Organic Framework via Pressure-Treated Engineering,” Angew. Chem. Int. Ed. 61(48), e202210836 (2022). [CrossRef]  

37. D. Zhang, W. Zhou, Q. Liu, and Z. Xia, “CH3NH3PbBr3 perovskite nanocrystals encapsulated in lanthanide metal–organic frameworks as a photoluminescence converter for anti-counterfeiting,” ACS Appl. Mater. Interfaces 10(33), 27875–27884 (2018). [CrossRef]  

38. B. Xu, H. Guo, S. Wang, Y. Li, H. Zhang, and C. Liu, “Solvothermal synthesis of luminescent Eu(BTC)(H2O) DMF hierarchical architectures,” CrystEngComm 14(8), 2914–2919 (2012). [CrossRef]  

39. D. Chen, W. Shen, S. Wu, C. Chen, X. Luo, and L. Guo, “Ion exchange induced removal of Pb (II) by MOF-derived magnetic inorganic sorbents,” Nanoscale 8(13), 7172–7179 (2016). [CrossRef]  

40. J. Zhang, J. Zeng, Y. Liu, L. Sun, F. Xu, W. You, and Y. Sawada, “Th·al decomposition kinetics of the synthetic complex Pb (1, 4-BDC)·(DMF)(H2O),” J. Therm. Anal. Calorim. 91(1), 189–193 (2008). [CrossRef]  

41. C. Zhang, Z. S. Li, X. Y. Dong, Y. Y. Niu, and S. Q. Zang, “Multiple responsive CPL switches in an enantiomeric pair of perovskite confined in lanthanide MOFs,” Adv. Mater. 34(11), 2109496 (2022). [CrossRef]  

42. X. Jing, D. Zhou, R. Sun, Y. Zhang, Y. Li, X. Li, Q. Li, H. Song, and B. Liu, “Enhanced Photoluminescence and Photoresponsiveness of Eu3+ Ions-Doped CsPbCl3 Perovskite Quantum Dots under High Pressure,” Adv. Funct. Mater. 31(31), 2100930 (2021). [CrossRef]  

43. S. Guo, Y. Zhao, K. Bu, Y. Fu, H. Luo, M. Chen, M. P. Hautzinger, Y. Wang, S. Jin, and W. Yang, “Pressure-suppressed carrier trapping leads to enhanced emission in two-dimensional perovskite (HA)2(GA)Pb2I7,” Angew. Chem. Int. Ed 132(40), 17686–17692 (2020). [CrossRef]  

44. T. Yang, Q. Xu, H. Mu, T. Lu, M. Jiao, J. Zhang, H. Wang, S. Zhang, K. Wang, and C. Yang, “Pressure-Induced Distinct Excitonic Properties of 2D Perovskites with Isometric Organic Molecules for Spacer Cations,” Nanoscale 15(13), 6234–6242 (2023). [CrossRef]  

45. L. Zhang, Q. Zeng, and K. Wang, “Pressure-induced structural and optical properties of inorganic halide perovskite CsPbBr3,” J. Phys. Chem. Lett. 8(16), 3752–3758 (2017). [CrossRef]  

46. G. Xiao, Y. Cao, G. Qi, L. Wang, C. Liu, Z. Ma, X. Yang, Y. Sui, W. Zheng, and B. Zou, “Pressure effects on structure and optical properties in cesium lead bromide perovskite nanocrystals,” J. Am. Chem. Soc. 139(29), 10087–10094 (2017). [CrossRef]  

47. W. Li, Z. Chen, J. Tang, and O. V. Prezhdo, “Anti-correlation between band gap and carrier lifetime in lead halide perovskites under compression rationalized by ab initio quantum dynamics,” Chem. Mater. 32(11), 4707–4715 (2020). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Materials

Data availability

No data were generated or analyzed in the presented research.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. (a) TEM images, (b) TEM magnification images, and (c) HR-TEM images of the CsPbBr2Cl@EuMOFs. (d) The XRD diffraction patterns of CsPbBr2Cl@EuMOFs. (e-j) EDX spectra of CsPbBr2Cl@EuMOFs nanocomposites, testing element including O, Eu, Cs, Pb, Br, Cl.
Fig. 2.
Fig. 2. (a) Structural composition model of the CsPbBr2Cl@EuMOFs. (b) PL spectra of the CsPbBr2Cl@EuMOFs and the EuMOFs at 405 nm excitation. (c) PL spectra of the CsPb(BrxCl1-x) 3@EuMOFs vary with PbBr2 doping concentration.
Fig. 3.
Fig. 3. (a) (b) Evolution of PL optical properties of CsPbBr2Cl@EuMOFs under high pressure, and (c) the changes in relative PL intensity of corresponding peaks, and (d) the changes in relative position of corresponding peaks. The fluctuation of the PL intensity caused by errors in experimental measurement are reasonably optimized.
Fig. 4.
Fig. 4. The changes in (a) PL intensity and (b) PL lifetime decay of CsPbBr2Cl PeQDs in nanocomposites under high pressure. (c) Evolution of PL optical properties of the EuMOFs under high pressure, and (d) statistics of PL intensity corresponding to peak values. The fluctuation of the PL intensity of EuMOFs caused by errors in experimental measurement are reasonably optimized.
Fig. 5.
Fig. 5. The luminescence mechanism of the CsPbBr2Cl@EuMOFs under high pressure.
Fig. 6.
Fig. 6. The spot color changes of the CsPbBr2Cl@EuMOFs under high pressure.
Fig. 7.
Fig. 7. The photos of CsPbBr2Cl@EuMOFs nanocomposites during the impregnation-recovery process reacted with Ethanol and CsX(Br/Cl). The variation of emission intensities, peak position, and FWHM values of the CsPbBr2Cl PeQDs in nanocomposites during the impregnation-recovery cycles for several cycle numbers.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.