Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Enhancement of the efficiency and thermal stability of the double perovskite Cs2AgInCl6 single crystal by Sc substitution

Liyan Chen ab, Hangjie Jiang ac, Zhaohua Luo *a, Guoqiang Liu a, Xianhui Wu ac, Yongfu Liu a, Peng Sun a and Jun Jiang *a
aNingbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China. E-mail: luozhaohua@nimte.ac.cn; jjun@nimte.ac.cn
bNano Science and Technology Institute, University of Science and Technology of China, Suzhou 215123, P. R. China
cSchool of Material Science and Engineering, Jiangxi University of Science and Technology, Ganzhou, 341000, P. R. China

Received 8th February 2022 , Accepted 12th April 2022

First published on 21st April 2022


Abstract

Cs2AgInCl6 lead-free double perovskite materials have potential applications in lighting and high-energy ray detection. However, the transition of Cs2AgInCl6 single crystals (SCs) is parity forbidden, resulting in low photoluminescence (PL) efficiency. In this study, we provide a strategy to break the parity forbidden transition by substituting Sc for In. The PL efficiency is greatly increased by 43%. In addition, the thermal stability is improved from 7.4%@475 K to 51.3%@475 K, which is attributed to the formation of Sc–Cl covalent bonds instead of In–Cl ionic ones. These results demonstrate that a suitable method makes a great contribution to promoting the possible application of Cs2AgInCl6 in the future.


Introduction

The lead-halide perovskites have recently shown tremendous application potential in solar cells, light-emitting diodes, photo-detectors and lasers1–6 due to their excellent performances, such as high photoluminescence quantum yield (PLQY), low-cost synthesis process, convenient tunability of emission colour and high tolerance to defects.7–9 However, the toxicity of Pb and its unstable characteristics in common working conditions with a humid environment, light and high temperatures are the biggest problems hindering their commercial applications.10–12 Therefore, it has become a focus for researchers to develop a new alternative material, which is non-toxic and chemically stable. Sn2+ and Ge2+ were first chosen to replace Pb2+, but these two elements are easily oxidized into a tetravalent state, which makes the corresponding halide materials more unstable.13

Recently, all-inorganic lead-free double perovskites were proposed and widely studied, in which two Pb2+ sites were simultaneously substituted by a monovalent and a trivalent cation, abbreviated as A2M+M3+X6 (A = Cs+; M+ = Ag+, Na+ or Cu+; M3+ = Bi3+, In3+or Sb3+; X = Cl, Br or I).14–18 The formation of a double perovskite structure can not only avoid toxicity, but also improve the stability of the material.19 Among them, lead-free double perovskite Cs2AgInCl6 has attracted wide attention since it was first proposed by Volonakis et al.20 This is because of its direct band gap character and suitable forbidden bandwidth, which makes rare earth and transition metal ion doping possible.21–25 According to previous research, the luminescence mechanism of Cs2AgInCl6 is related to self-trapped excitons (STEs), arising from the strong Jahn–Teller distortion of the AgCl6 octahedron.26 This is proved by both theoretical calculations and experiments. However, the STEs and free excitons in Cs2AgInCl6 have the same orbital parity, which leads to a parity-forbidden transition.7 As a result, Cs2AgInCl6 always suffers from a low photoluminescence quantum yield (PLQY < 0.1%). How to improve its PLQY has become one of the most important subjects to be solved.

The doping of ions is considered to be a promising method to endow and adjust the optical and optoelectronic properties of inorganic materials in the visible region.22 Liu et al. increased the PLQY to 11.4% by doping Bi in Cs2AgInCl6 NCs.27 It was also reported that Mn was doped into Cs2AgInCl6 NCs to obtain orange-red emission, and the PLQY was as high as 16%.28 Most previous reports on Cs2AgInCl6 targeted nanocrystals. Nevertheless, the double-doped Cs2Ag0.6Na0.4InCl6: 0.04%Bi single crystal prepared by Luo et al. has bright white light emission, with a PLQY of 86.2% ± 5%. They break the parity forbidden transition by partially replacing Ag+ with Na+ in the Cs2AgInCl6 lattice and changing the wave function of STEs, so as to significantly improve the PL emission intensity. Furthermore, the STEs produced by the strong Jahn–Teller distortion of the AgCl6 octahedron in the excited state obtain a wider visible light emission.7

Although Na and Bi double-doped Cs2AgInCl6 single crystals have achieved excellent optical properties, previous studies mainly focused on the intrinsic band gap and photoluminescence (PL) behaviors,20,29–31 but little on their thermal stability. In the actual application of Cs2AgInCl6 materials, not only the luminescence properties, such as quantum yield, but also stability must be considered, and thermal stability is an extremely important aspect.

Inspired by these results, we propose for the first time to dope Scandium (Sc) into Cs2AgInCl6, with the aim to improve its PL efficiency and thermal stability. On the one hand, Sc3+ ions have the same valence state and similar ionic radius as that of In3+, which makes it possible to be alloyed into Cs2AgInCl6, and subsequently, its PL efficiency would be improved by the broken symmetry of the crystal structure. More importantly, when Sc3+ is doped, the electrons on the Sc d orbital and those on the Cl p orbital tend to form a covalent bond-like structure. Therefore, the thermal stability of Cs2AgIn1−xScxCl6 (x > 0) would be effectively improved. We believe that the alloying of Sc3+ would prove to be a new way to help to promote the real application of Cs2AgInCl6.

Results and discussion

1. Morphology and crystal structure of Cs2AgIn1−xScxCl6

Fig. S1 (ESI) presents the picture of the Cs2AgInCl6 crystal, which exhibits a regular octahedral shape with a size of about 1 mm. To confirm the crystalline character, the powder X-ray diffraction (XRD) patterns of Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4, 0.6, 0.8 and 1) are collected and shown in Fig. 1a. It is found that almost all the peaks match well with the reference standard pattern (Cs2AgInCl6, (ICSD-244519)) when the Sc doping amount x ≤ 0.4, indicating that Sc3+ ions have entered the lattice and formed a solid solution with the Cs2AgInCl6 substrate. In addition, from the enlarged view of the characteristic (220) diffraction peak as shown in Fig. 1a, it also can be found that the diffraction peak shifts to a higher angle as the amount of x increases from 0 to 0.8. Noted that, the ionic radius for 6-coordinated Sc3+ (r = 0.745 Å) is slightly smaller than that of In3+ (r = 0.8 Å). According to the Bragg equation, the substitution of Sc3+ for In3+ would cause shrinkage of the lattice and result in a peak shift towards a larger angle, which is supported by the results of the calculated lattice constant as shown in Fig. 1b. However, when the amount of Sc3+ is more than 60% (x = 0.6), the AgCl second phase appears, and only the AgCl phase can be found when x = 1, which means that the solid solubility of Sc3+ in the Cs2AgInCl6 lattice is less than 60%.
image file: d2ma00139j-f1.tif
Fig. 1 (a) XRD patterns and selected diffraction peaks at around 24° of the as-prepared Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4, 0.6, 0.8 and 1), and (b) the change of cell parameters with Sc doping content.

In order to clarify the influence of Sc3+ ions on the crystal structure, samples of Cs2AgInCl6, Cs2AgIn0.8Sc0.2Cl6, Cs2AgIn0.6Sc0.4Cl6 and Cs2AgIn0.4Sc0.6Cl6 are taken for further XRD measurement and refined by Rietveld refinement, as presented in Fig. S2 (ESI). The detailed structure parameters of Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4 and 0.6) determined with Rietveld refinement and the reliability factors are listed in Table S1 (ESI). The reliability factors in Table S1 (ESI) indicate that the results of Rietveld refinements are reliable. According to the Rietveld refinement results, Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4 and 0.6) belongs to the cubic crystal system with the Fm[3 with combining macron]m space group. Moreover, the cell parameter decreases from 10.502 Å to 10.495 Å with the introduction of 60% Sc3+ (x = 0.6) ions, as plotted in Fig. 1b. It should be noted that, according to the results of Rietveld refinement, the occupancies of Sc3+ in In3+ sites are much less than the designed values (as shown in Table S1, ESI).

To identify the composition uniformity, the Cs2AgIn0.6Sc0.4Cl6 single-crystal particle with extremely small size was selected and analyzed by EDS-mapping, which is given in Fig. 2. As shown in the SEM image of Fig. 2a, the Cs2AgIn0.6Sc0.4Cl6 single-crystal presents a regular octahedron morphology. The pictures of EDS mapping show that Cs, Ag, In, Cl and Sc elements are homogeneously distributed within the selected particle. However, the Sc intensity is very weak, which is consistent with the results of Rietveld refinement as shown in Table S1 (ESI).


image file: d2ma00139j-f2.tif
Fig. 2 (a) SEM image of the Cs2AgInCl6 SC. (b–e) EDS-mapping of Cs, Ag, In and Cl for the selected Cs2AgInCl6 particle. (f) SEM image of the Cs2AgIn0.6Sc0.4Cl6 SC. (g–k) EDS-mapping of Cs, Ag, In, Sc and Cl for the selected Cs2AgIn0.6Sc0.4Cl6 particle.

2. Spectral characteristics of Cs2AgIn1−xScxCl6 single crystals

With the aim to understand the effect of Sc3+ doping on the optical properties of Cs2AgInCl6, the normalized steady-state absorption spectra of Cs2AgIn1−xScxCl6 single crystals with x varying from 0 to 0.6 are measured and shown in Fig. 3a. As for Cs2AgInCl6, a sharp absorption edge starting at around 400 nm is observed, which is attributed to the allowed transition and consistent with the previous reports. With the increase of the Sc doping level, this absorption edge shifts to a longer wavelength. In addition, there is another broad absorption with the edge starting around 900 nm for samples with Sc doping. The absorption intensity enhanced, and the absorption edges of Cs2AgIn1−xScxCl6 broadened with the increase of the Sc content. In order to determine the optical bandgaps of Cs2AgIn1−xScxCl6 crystals, the following equation was used:
 
[F(R)*]n = A(Eg)(1)
where is the photon energy; A is a proportional constant; Eg represents the value of the band gap; n = 2 since the Cs2AgIn[thin space (1/6-em)]Cl6 host absorption is a direct transition; and F(R) is the Kubelka–Munk function defined as:
 
F(R) = (1 − R)2/R(2)
where R represents the reflection coefficient.32 Tauc plots of Cs2AgIn1−xScxCl6 single crystals (x = 0, 0.2, 0.4 and 0.6) are presented in Fig. 3b. It shows that the indirect bandgap appears with the doping of Sc3+, and it increases from 1.0 to 1.18 eV as the amount of Sc3+ increases from 0.2 to 0.6. The variation of the bandgaps would be attributed to the change in electronic structure that we will discuss in the next section. In addition, with the increase in the Sc doping level, the energy difference from the VBM to the second conduction band decreases from 3.29 eV for Cs2AgInCl6 to 2.44 eV for Cs2AgIn0.4Sc0.6Cl6.

image file: d2ma00139j-f3.tif
Fig. 3 The normalized steady-state absorption spectra (a), the Tauc plot of Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4, 0.6 and 0.8) single crystals (b), and photoluminescence (PL) spectra (c).

The steady-state photoluminescence spectra of the Cs2AgIn1−xCl6: xSc3+ samples are shown in Fig. 3c. Excited by ultraviolet radiation of 370 nm, all samples exhibit broadband emission character with a peak at around 620 nm, and the position of these peaks is almost unchanged upon tuning the content of Sc (as shown in Fig. S4, ESI). In general, the broadband emission usually comes from STEs or excited state reorganization, which exists in semiconductors with localized carriers and soft crystal lattices, as reported previously.33–36 The STEs that exist in Cs2AgInCl6 crystals are thought to be caused by the Jahn–Teller distortion of the AgCl6 octahedron, which emits effectively and stably. Moreover, the integral intensity for the PL spectra increases at first and then decreases with a maximum increment of 43% when x = 0.4 (as shown in Fig. S4, ESI). The reason can be attributed to the octahedral AgCl6 distortion with the doping of Sc, which would help to break the reverse symmetry and modify the parity forbidden transition subsequently.

3. Electronic structure of Cs2AgScCl6 single crystals

In order to have a deeper understanding of the changes brought about by Sc doping, we analyze the electronic structure of Cs2AgScCl6, as shown in Fig. 4.
image file: d2ma00139j-f4.tif
Fig. 4 The calculated band structure (a), and density of states (b) for Cs2AgScCl6.

Fig. 4a and b, respectively, present the calculated band structure and density of states (DOS) for Cs2AgScCl6. The calculations were performed using the full-potential package of WIEN2K with an 8 × 8 × 8 K-point mesh. The previous theoretical calculation indicated that Cs2AgInCl6 is a direct gap semiconductor.29Fig. 4a shows that Cs2AgScCl6 is an indirect semiconductor, while the valence band maximum (VBM) and conduction band minimum (CBM) appear at the L and X points, respectively. The obtained band gap is about 2.5 eV, which is smaller than that of Cs2AgInCl6 in the previous report.27,37–39 Therefore, it is reasonable to deduce that the direct bandgap of Cs2AgInCl6 would decrease with the doping of Sc and that is in accordance with our results as shown in Fig. 3b.

The DOS of Cs2AgScCl6 and the partial DOS for Sc are displayed in Fig. 4b. As may be seen, the conduction band around 3 eV is mainly from the orbital of Sc. As for a pure Sc3+ ion, its d-orbital should be fully empty and is located at the position of the conduction band. However, the partial DOS of Sc shows a peak around −3.5 eV. This fact indicates that Sc-d and Cl-p orbitals form a valence bond rather than an ionic bond. Usually, the bond strength of valence bonding is stronger than ionic bonding.40,41 For this reason, the lattice constant of Cs2AgScCl6 is decreased relative to Cs2AgInCl6, and the thermal stability would also be improved.

4. Thermal stability of Cs2AgIn1−xScxCl6 single crystals

In real applications, such as solar photovoltaics and semiconductor lighting, heat generation in photoelectric conversion materials is unavoidable. As a result, it requires the materials to have robust thermal stability, which was often neglected in the past. The temperature-dependent PL spectra of Cs2AgIn1−xScxCl6 (x = 0, 0.2, 0.4 and 0.6) single crystals are shown in Fig. 5a–d, where a slight blue shift is observed for the emission peaks upon the increase of the temperature (as shown in Fig. S5, ESI). Here, we consider that both electron–phonon interaction and crystal lattice thermal expansion would be responsible for this blue shift, and a similar phenomenon has been observed in lead/copper chalcogenide semiconductors and lead-halide perovskites.42–44
image file: d2ma00139j-f5.tif
Fig. 5 Temperature-dependent PL spectra (a–d) and the relationship between PL intensity and temperature for Cs2AgIn0.8Sc0.2Cl6, Cs2AgIn0.6Sc0.4Cl6 and Cs2AgIn0.4Sc0.6Cl6, respectively (e).

To clarify the thermal stability of Cs2AgIn1−xScxCl6 single crystals, the relationship between integral PL intensities and temperature is presented in Fig. 5e. It clearly shows that the emission intensity decreases with the rise of temperature, owing to the thermally activated nonradiative recombination. Moreover, it can also be observed that the rate of PL intensity decline slows down with the increase of Sc doping concentration, indicating that the thermal stability is effectively improved. For example, the PL intensity of Cs2AgIn0.4Sc0.6Cl6 at 475 K still maintained 51.3% of that at room temperature, while the PL intensity of Cs2AgInCl6 only maintained 7.4%. There are mainly two reasons for this enhancement. First, the electrons on the In-s orbital and Cl-p orbital in Cs2AgInCl6 are usually in the form of ionic bonds, but when Sc3+ is doped, electrons on the Sc-d orbital and Cl-p orbital tend to form a structure similar to a covalent bond, which is pointed out in the analysis of the DOS results in the previous section. Second, the incorporation of Sc3+ causes the contraction of the Cs2AgInCl6 crystal lattice, and the bond length becomes shorter. As a result, the interaction between ions would be enhanced, and the rigidity of the crystal lattice will be improved.

Conclusions

In conclusion, Sc-alloyed Cs2AgInCl6 SCs have been successfully synthesized by hydrothermal synthesis. Results of Rietveld refinement show that Sc3+ ions take the place of In3+ sites, but their actual occupancies are considerably lower than the designed values. With the incorporation of Sc, the PL intensity of Cs2AgInCl6 SCs can be enhanced by 43%. This may be caused by the local structure distortion that would break the reverse symmetry and subsequently modify the parity forbidden transition of STEs. In addition, the thermal stability of Cs2AgInCl6 is also enhanced with the doping of Sc. The PL intensity of Cs2AgIn0.4Sc0.6Cl6 at 475 K is 51.3% of that at room temperature, while it drops to 7.4% for Cs2AgInCl6. According to the calculated results of electronic structure, the enhancement of the thermal stability is attributed to the formation of a covalent bond for Sc–Cl other than the ionic bond for In–Cl. There is a reason to believe that the Sc doping would provide an effective solution to enhance the problems of low PL and thermal stability for Cs2AgInCl6 single crystals.

Experimental section

Materials

Silver chloride (AgCl, 99.5%), sodium chloride (NaCl, 99.99%), anhydrous indium chloride (InCl3, 99.99%), hydrochloric acid and isopropanol were purchased from Aladdin. Cesium chloride (CsCl, AR) and scandium nitrate hydrate (Sc(NO3)3, 99.9%) were purchased from Macklin. All chemicals were used without any further purification.

Synthesis of Cs2AgIn1−xScxCl6 single crystals

1 mmol of AgCl, 1 − x mmol of anhydrous InCl3, x mmol Sc (NO3)3 and 2 mmol of CsCl were first dissolved in 10 ml M HCl solution in a beaker. After stirring for 30 minutes, the solution was then transferred to a 25 ml Teflon autoclave. The Teflon autoclave was heated at 180 °C for 12 h in a stainless steel Parr Auto-Claveand, and then, the solution was stably cooled to 50 °C at a speed of 5 °C h−1. The as-prepared crystals were then filtered out, washed twice with isopropanol and dried in a furnace at 65 °C.

Characterization

Powder XRD measurements were performed by grinding Cs2AgIn1−x ScxCl6 crystals into fine powders in a mortar, and then using a Bruker D8 X-ray diffractometer with Cu Kα radiation (λ = 1.54056 Å) at 40 kV and 40 mA. The crystal structures were determined with Rietveld refinement by using a general structure analysis system (GSAS).45 A field-emission scanning electron microscope (FE-SEM, Hitachi S-4800) equipped with an energy-dispersive X-ray spectroscopy (EDS) system was used to measure the morphology. The photoluminescence and PLE measurements were carried out using a Horiba FL3-111 fluorescence spectrometer with a xenon lamp as a light source. The temperature-dependent photoluminescence spectra were measured using the same instrument excited by a 360 nm wavelength and at a temperature ranging from 300 to 475 K. The absorption spectra were measured at room temperature using a LAMBDA 950 ultraviolet-visible-near infrared spectrophotometer.

Author contributions

Liyan Chen: investigation, validation, formal analysis, data curation, writing – original draft; Hangjie Jiang: investigation, data curation; Zhaohua Luo: investigation, data curation, writing – review & editing, funding acquisition, supervision; Guoqiang Liu: investigation, data curation; Xianhui Wu: data curation; Yongfu Liu: investigation, data curation, funding acquisition; Peng Sun: investigation, data curation; Jun Jiang: resources, writing – review & editing, supervision, funding acquisition, project administration.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work is financially supported by the “Pioneer” and “Leading Goose” R&D Program of Zhejiang (2022C01046), and the National Natural Science Foundation of China (12074393, 62105342).

References

  1. L. Protesescu, S. Yakunin, M. I. Bodnarchuk, F. Krieg, R. Caputo, C. H. Hendon, R. X. Yang, A. Walsh and M. V. Kovalenko, Nano Lett., 2015, 15, 3692–3696 CrossRef CAS PubMed.
  2. H. Huang, M. I. Bodnarchuk, S. V. Kershaw, M. V. Kovalenko and A. L. Rogach, ACS Energy Lett., 2017, 2, 2071–2083 CrossRef CAS PubMed.
  3. R. E. Brandt, J. R. Poindexter, P. Gorai, R. C. Kurchin, R. L. Z. Hoye, L. Nienhaus, M. W. B. Wilson, J. A. Polizzotti, R. Sereika, R. Zaltauskas, L. C. Lee, J. L. MacManus-Driscoll, M. Bawendi, V. Stevanovic and T. Buonassisi, Chem. Mater., 2017, 29, 4667–4674 CrossRef CAS.
  4. J. Kang and L. W. Wang, J. Phys. Chem. Lett., 2017, 8, 489–493 CrossRef CAS PubMed.
  5. Z. Jia, C. Yuan, Y. Liu, X. J. Wang, P. Sun, L. Wang, H. Jiang and J. Jiang, Light: Sci. Appl., 2020, 9, 86 CrossRef CAS PubMed.
  6. Q. Yao, P. Hu, P. Sun, M. Liu, R. Dong, K. Chao, Y. Liu, J. Jiang and H. Jiang, Adv. Mater., 2020, 32, e1907888 CrossRef PubMed.
  7. J. Luo, X. Wang, S. Li, J. Liu, Y. Guo, G. Niu, L. Yao, Y. Fu, L. Gao, Q. Dong, C. Zhao, M. Leng, F. Ma, W. Liang, L. Wang, S. Jin, J. Han, L. Zhang, J. Etheridge, J. Wang, Y. Yan, E. H. Sargent and J. Tang, Nature, 2018, 563, 541–545 CrossRef CAS PubMed.
  8. Y. Luo, N. Liu, X. Li, J. C. Hone and S. Strauf, 2D Mater., 2019, 6 Search PubMed.
  9. X. Wang, W. Meng, W. Liao, J. Wang, R. G. Xiong and Y. Yan, J. Phys. Chem. Lett., 2019, 10, 501–506 CrossRef PubMed.
  10. C. Bi, S. V. Kershaw, A. L. Rogach and J. Tian, Adv. Funct. Mater., 2019, 29 Search PubMed.
  11. Y. Wei, Z. Cheng and J. Lin, Chem. Soc. Rev., 2019, 48, 310–350 RSC.
  12. H. C. Wang, S. Y. Lin, A. C. Tang, B. P. Singh, H. C. Tong, C. Y. Chen, Y. C. Lee, T. L. Tsai and R. S. Liu, Angew. Chem., Int. Ed., 2016, 55, 7924–7929 CrossRef CAS PubMed.
  13. S. Ghosh and B. Pradhan, ChemNanoMat, 2019, 5, 300–312 CrossRef CAS.
  14. Q. A. Akkerman, G. Raino, M. V. Kovalenko and L. Manna, Nat. Mater., 2018, 17, 394–405 CrossRef CAS PubMed.
  15. F. Wei, Z. Deng, S. Sun, F. Zhang, D. M. Evans, G. Kieslich, S. Tominaka, M. A. Carpenter, J. Zhang, P. D. Bristowe and A. K. Cheetham, Chem. Mater., 2017, 29, 1089–1094 CrossRef CAS.
  16. Z. W. Xiao, W. W. Meng, J. B. Wang, D. B. Mitzi and Y. F. Yan, Mater. Horiz., 2017, 4, 206–216 RSC.
  17. A. H. Slavney, T. Hu, A. M. Lindenberg and H. I. Karunadasa, J. Am. Chem. Soc., 2016, 138, 2138–2141 CrossRef CAS PubMed.
  18. D. H. Fabini, T. Hogan, H. A. Evans, C. C. Stoumpos, M. G. Kanatzidis and R. Seshadri, J. Phys. Chem. Lett., 2016, 7, 376–381 CrossRef CAS PubMed.
  19. K. Dave, M. H. Fang, Z. Bao, H. T. Fu and R. S. Liu, Chem. – Asian J., 2020, 15, 242–252 CrossRef CAS PubMed.
  20. G. Volonakis, A. A. Haghighirad, R. L. Milot, W. H. Sio, M. R. Filip, B. Wenger, M. B. Johnston, L. M. Herz, H. J. Snaith and F. Giustino, J. Phys. Chem. Lett., 2017, 8, 772–778 CrossRef CAS PubMed.
  21. Y. Mahor, W. J. Mir and A. Nag, J. Phys. Chem. C, 2019, 123, 15787–15793 CrossRef CAS.
  22. Y. Liu, X. Rong, M. Li, M. S. Molokeev, J. Zhao and Z. Xia, Angew. Chem., Int. Ed., 2020, 59, 11634–11640 CrossRef CAS PubMed.
  23. Z. Zeng, B. Huang, X. Wang, L. Lu, Q. Lu, M. Sun, T. Wu, T. Ma, J. Xu, Y. Xu, S. Wang, Y. Du and C. H. Yan, Adv. Mater., 2020, 32, 2004506 CrossRef CAS PubMed.
  24. Q. Liao, J. Chen, L. Zhou, T. Wei, L. Zhang, D. Chen, F. Huang, Q. Pang and J. Z. Zhang, J. Phys. Chem. Lett., 2020, 11, 8392–8398 CrossRef CAS PubMed.
  25. P. Han, X. Zhang, C. Luo, W. Zhou, S. Yang, J. Zhao, W. Deng and K. Han, ACS Cent. Sci., 2020, 6, 566–572 CrossRef CAS PubMed.
  26. Y. Liu, A. Nag, L. Manna and Z. Xia, Angew. Chem., Int. Ed., 2021, 60, 11592–11603 CrossRef CAS PubMed.
  27. Y. Liu, Y. Jing, J. Zhao, Q. Liu and Z. Xia, Chem. Mater., 2019, 31, 3333–3339 CrossRef CAS.
  28. W. Lee, S. Hong and S. Kim, J. Phys. Chem. C, 2019, 123, 2665–2672 CrossRef CAS.
  29. J. Zhou, Z. G. Xia, M. S. Molokeev, X. W. Zhang, D. S. Peng and Q. L. Liu, J. Mater. Chem. A, 2017, 5, 15031–15037 RSC.
  30. J. Luo, S. Li, H. Wu, Y. Zhou, Y. Li, J. Liu, J. Li, K. Li, F. Yi, G. Niu and J. Tang, ACS Photonics, 2017, 5, 398–405 CrossRef.
  31. T. T. Tran, J. R. Panella, J. R. Chamorro, J. R. Morey and T. M. McQueen, Mater. Horiz., 2017, 4, 688–693 RSC.
  32. F. Y. Zhao, Z. Song, J. Zhao and Q. L. Liu, Inorg. Chem. Front., 2019, 6, 3621–3628 RSC.
  33. Z. Yuan, C. Zhou, Y. Tian, Y. Shu, J. Messier, J. C. Wang, L. J. van de Burgt, K. Kountouriotis, Y. Xin, E. Holt, K. Schanze, R. Clark, T. Siegrist and B. Ma, Nat. Commun., 2017, 8, 14051 CrossRef CAS PubMed.
  34. C. Zhou, Y. Tian, Z. Yuan, H. Lin, B. Chen, R. Clark, T. Dilbeck, Y. Zhou, J. Hurley, J. Neu, T. Besara, T. Siegrist, P. Djurovich and B. Ma, ACS Appl. Mater. Interfaces, 2017, 9, 44579–44583 CrossRef CAS PubMed.
  35. C. Zhou, Y. Tian, M. Wang, A. Rose, T. Besara, N. K. Doyle, Z. Yuan, J. C. Wang, R. Clark, Y. Hu, T. Siegrist, S. Lin and B. Ma, Angew. Chem., Int. Ed., 2017, 56, 9018–9022 CrossRef CAS PubMed.
  36. C. Zhou, H. Lin, Y. Tian, Z. Yuan, R. Clark, B. Chen, L. J. van de Burgt, J. C. Wang, Y. Zhou, K. Hanson, Q. J. Meisner, J. Neu, T. Besara, T. Siegrist, E. Lambers, P. Djurovich and B. Ma, Chem. Sci., 2018, 9, 586–593 RSC.
  37. J. C. Dahl, W. T. Osowiecki, Y. Cai, J. K. Swabeck, Y. Bekenstein, M. Asta, E. M. Chan and A. P. Alivisatos, Chem. Mater., 2019, 31, 3134–3143 CrossRef CAS.
  38. B. Nath, B. Pradhan and S. K. Panda, New J. Chem., 2020, 44, 18656–18661 RSC.
  39. P. G. Han, X. Zhang, X. Mao, B. Yang, S. Q. Yang, Z. C. Feng, D. H. Wei, W. Q. Deng, T. Pullerits and K. L. Han, Sci. China: Chem., 2019, 62, 1405–1413 CrossRef CAS.
  40. A. Burawoy, Trans. Faraday Soc., 1943, 39, 79–90 RSC.
  41. H. A. Skinner, Thermochim. Acta, 1986, 109, 165–174 CrossRef CAS.
  42. S. Li, H. Y. Wang, P. Yang, L. L. Wang, X. R. Cheng and K. Yang, J. Alloys Compd., 2021, 854 CAS.
  43. P. Dey, J. Paul, J. Bylsma, D. Karaiskaj, J. M. Luther, M. C. Beard and A. H. Romero, Solid State Commun., 2013, 165, 49–54 CrossRef CAS.
  44. Y. Li, Z. F. Shi, L. Z. Lei, F. Zhang, Z. Z. Ma, D. Wu, T. T. Xu, Y. T. Tian, Y. T. Zhang, G. T. Du, C. X. Shan and X. J. Li, Chem. Mater., 2018, 30, 6744–6755 CrossRef CAS.
  45. Z. H. Luo, Y. F. Liu, C. H. Zhang, J. X. Zhang, H. M. Qin, H. C. Jiang and J. Jiang, Inorg. Chem., 2016, 55, 3040–3046 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ma00139j

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.