Next Article in Journal
Dispersion Management Nonlinear Multimode Interference Mode-Locked Ytterbium Fiber Laser
Next Article in Special Issue
Enhanced Photoluminescence of Crystalline Alq3 Micro-Rods Hybridized with Silver Nanowires
Previous Article in Journal
Recent Advances in Reconfigurable Metasurfaces: Principle and Applications
Previous Article in Special Issue
Ultrasound-Promoted Abatement of Formaldehyde in Liquid Phase with Electrospun Nanostructured Membranes: The Synergy of Combined AOPs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Enhancing Photoluminescence of CsPb(ClxBr1−x)3 Perovskite Nanocrystals by Fe2+ Doping

1
Jiangsu Key Laboratory of Micro and Nano Heat Fluid Flow Technology and Energy Application, School of Physical Science and Technology, Suzhou University of Science and Technology, Suzhou 215009, China
2
School of Environmental Science and Engineering, Suzhou University of Science and Technology, Suzhou 215009, China
3
School Optoelect Engn, Zaozhuang University, Zaozhuang 277160, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 533; https://doi.org/10.3390/nano13030533
Submission received: 30 December 2022 / Revised: 19 January 2023 / Accepted: 20 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Advance in Photoactive Nanomaterials)

Abstract

:
The doping of impurity ions into perovskite lattices has been scrupulously developed as a promising method to stabilize the crystallographic structure and modulate the optoelectronic properties. However, the photoluminescence (PL) of Fe2+-doped mixed halide perovskite NCs is still relatively unexplored. In this work, the Fe2+-doped CsPb(ClxBr1−x)3 nanocrystals (NCs) are prepared by a hot injection method. In addition, their optical absorption, photoluminescence (PL), PL lifetimes, and photostabilities are compared with those of undoped CsPb(Br1−xClx)3 NCs. We find the Fe2+ doping results in the redshift of the absorption edge and PL. Moreover, the full width at half maximums (FWHMs) are decreased, PL quantum yields (QYs) are improved, and PL lifetimes are extended, suggesting the defect density is reduced by the Fe2+ doping. Moreover, the photostability is significantly improved after the Fe2+ doping. Therefore, this work reveals that Fe2+ doping is a very promising approach to modulate the optical properties of mixed halide perovskite NCs.

1. Introduction

Lead halide ABX3 (X = Cl, Br, I) perovskites are very competitive semiconductors for solar cells, lighting, and photodetectors owing to their tremendous optical properties, such as a tunable fluorescent color across the whole visible range, strong absorption, high photoluminescence (PL) quantum yields (QYs), narrow PL bandwidths, and suppressed PL blinking [1,2,3,4,5,6,7,8]. Although the perovskites exhibit extraordinary potential and there are already many exciting progresses, the further commercialization of these promising semiconductors still confronts severe challenges. One is the inclusion of toxic Pb and another one is the instability against moisture, oxygen, heat, and electric current/irradiation [9,10,11,12,13,14]. In particular, for the mixed halide perovskites, the photo- or electric-induced ion migration is ineluctable, resulting in phase segregation [15,16,17]. Therefore, during the past years, many research works have been conducted to decrease the use of toxic Pb and improve the stability of perovskites [18,19,20]. To avoid the utilization of toxic lead, various lead-free perovskites, such as Sn2+, Sn4+, Mn2+, and Cu2+-based perovskites have been exploited [18,19,20]. Among them, tin-based perovskites have been the most explored [19]. However, the Sn2+ in perovskite is easily oxidized to Sn4+, causing poor stability [21]. The PL QYs of tin-based perovskites and power conversion efficiency (PCE) of solar cells based on tin-based perovskites are very low [11].
Doping of other metal ions to partially replace Pb is another effective way to reduce the utilization of toxic lead, while maintaining or even enhancing the excellent optical and photoelectrical properties. The incorporation of appropriate impurity ions into host lattices has been exploited as a promising method to stabilize the crystallographic phases while modulating the optical, electronic, and magnetic properties of diverse semiconductors [22,23,24,25,26,27,28]. Regarding the halide perovskites, the partial substitution of Pb2+ by divalent metal ions, such as Cu2+, Mg2+, Fe2+, Co2+, Ni2+, and Mn2+ at the B sites of the perovskite lattice have been demonstrated [29,30,31,32,33,34,35,36]. Klug et al revealed that the perovskite films retain an excellent photovoltaic performance if less than 3% Pb2+ ions are substituted by homovalent metal species due to the high tolerance of the perovskite lattices [31]. To date, many different research groups have recently demonstrated that Mn2+ ions can be doped into the B sites of perovskite lattices by using a facile approach [32,33,34,35]. Moreover, a broad PL peak at about 600 nm is induced by the 4T1 to 6A1 transition of Mn2+, which can be applied for multicolor luminescence [35]. After anion exchange reactions between Mn-doped CsPbCl3 and CsPbBr3, fluorescence color gamut almost covering the entire visible spectrum are obtained [35]. Thanks to the high stability and wide color gamut, color converters for light-emitting diodes (LEDs) were constructed. Similarly, the doping of Ni2+ ions into all inorganic perovskite nanocrystals (NCs) can also modulate the PL. Sun et al have a general strategy for the synthesis of Ni-doped CsPbCl3 NCs, which shows a strong single-color violet emission with a maximum PL QY of 96.5% [36].
Fe ions, as the earth-abundant elements, are eco-friendly and low-cost. The doping of Fe ions in perovskites has also attracted research interests. CH3NH3PbCl3 single crystals with different concentrations of Fe2+/Fe3+ doping were synthesized by Cheng et al [37]. In addition, the crystal structure, optical, and optoelectronic properties were investigated. They revealed that Fe2+ is prone to replacing Pb2+ and the optoelectronic properties are seriously deteriorated. On the contrary, Hu et al. reported that an appropriate amount of Fe2+ doping into the lattice of CsPbCl3 NCs not only improved the homogeneity of the size of the NCs, but also enhanced the PL QY and average PL lifetimes [38]. Therefore, the impact of Fe2+ doping on the optoelectronic properties of perovskite NCs is still unclear.
In this work, the Fe2+-doped CsPb1−xFex(Br1−xClx)3 NCs are prepared by mixing FeCl2 (x mmol) and PbBr2 (1−x mmol) during the hot injection process. When x is not zero, Pb2+ at the B sites are partially replaced by Fe2+; meanwhile, the X sites are also partially occupied by Cl-. Therefore, the PL properties of Fe2+-doped mixed halide perovskite NCs are investigated. To our knowledge, the PL of Fe2+-doped mixed halide perovskite NCs is studied for the first time. The morphology and size distribution of the Fe2+-doped perovskite NCs are investigated by a transmission electron microscope (TEM). The optical absorption, PL, PL lifetimes, and photostabilities of the Fe2+-doped perovskite NCs are measured, which are compared with those of the undoped CsPb(Br1−xClx)3 NCs.

2. Materials and Methods

2.1. Synthesis of CsPb1−xFex(Br1−xClx)3 (x = 0, 0.1, 0.2, and 0.3) NCs

Cs-oleate was prepared by dispersing Cs2CO3 powders (0.407 g, 1.25 mmol) into a mixture of 18 mL octadecene (ODE, Aladdin, 90%) and 1.74 mL oleic acid (OA, Aldrich, 90%), which was then heated to 150 °C under N2 atmosphere until all Cs2CO3 powders were reacted. Then, FeCl2 (x mmol, x = 0, 0.1, 0.2, and 0.3) and PbBr2 (1−x mmol) were mixed with OLA (1.5 mL), oleylamine (OLA, 70%, 1.5 mL), trioctylphosphine (TOP, 90%, 1 mL), and ODE (10 mL) in a 100 mL 3-neck flask. The mixture was then degassed at 110 °C for 40 min. After that, the mixture was heated to 170 °C under N2 atmosphere. After reaction for 15 s, the hot Cs-oleate precursor (1 mL) was quickly injected. Subsequently, an ice-water bath was used to cool down the reaction. The resulting solution was centrifuged at 5000 r/m for 5 min, the supernatant was discarded. To remove the residual reactants, the solids were redispersed in hexane and centrifuged again for 5 min. When x = 0.4, well-shaped perovskite nanocrystals cannot be obtained. The reaction was conducted under protection of N2 and no other oxidant was added. Thus, the oxidation of feeding Fe2+ to Fe3+ is negligible. Moreover, Fe2+ is expected to replace Pb2+ as confirmed previously [37,38]. It is reasonable that the main valence of doping iron is bivalent.

2.2. Characterizations

Transmission electron microscope (TEM) images were captured on the JEM-2100F electron microscope (JOEL, Japan Electronics Co., Ltd, Tokyo, Japan). Elemental analysis was conducted using energy-dispersive X-ray spectroscopy (EDS) coupled on the TEM. The absorption spectra were measured by using a Shimazu UV2600 UV-Vis spectrophotometer, Shimazu Co., Ltd, Tokyo, Japan). The PL spectra were measured by a Maya 2000 Pro high sensitivity spectrometer (Ocean Optics Co., Ltd, Orlando, USA). The time-resolved PL decay curves were measured by a PicoHarp 300 time-correlated single photon counting (TCSPC) system (Pico Quant Co., Ltd, Berlin, Germany).

3. Results and Discussion

3.1. Structure

Figure 1a,b show the TEM and size distribution of the CsPb(Br0.8Cl0.2)3 NCs without the Fe2+ doping. The Fe2+-doped perovskite NCs were prepared by a hot injection method. The feeding ratio between the FeCl2 (x mmol) and PbBr2 (1−x mmol) was adjusted to simultaneously control the doping concentration and halide composition. Figure 1c shows the TEM image of the CsPb1−xFex(Br1−xClx)3 NCs when x = 0.2. It shows that the NCs are monodispersed with regular cubic shapes. In addition, a size distribution is obtained based on the TEM image. As shown in Figure 1d, the sizes of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs are about 9.3 nm with a relatively uniform distribution. The elemental distributions are measured by energy-dispersive X-ray spectroscopy (EDS) equipped on the TEM. The result shown in Figure 1e suggests the homogeneous presence of Fe, indicating the Fe2+ ions are successfully doped into NCs. The actual atomic ratio between Pb and Fe determined by the whole EDS mapping of Figure 1e is about 4.3:1 in Figure 1f. Both the shape and size distribution of the CsPb(Br1−xClx)3 NCs are very similar to those of the CsPb1−xFex(Br1−xClx)3 NCs. Therefore, we found the doping of the Fe2+ ion at a low concentration has a negligible influence on the growth of perovskite NCs.
The absorption spectra of the CsPb1−xFex(Br1−xClx)3 NCs are shown in Figure 2a. All the absorption spectra show a sharp edge with a negligible Urbach tail, indicating the low density of defect trapping. When x is not zero, Pb2+ at the B sites are partially replaced by Fe2+; meanwhile, the X sites are also partially occupied by Cl. Typically, as the fraction of Cl increases, the bandgap of the perovskite increases and the PL peak position blueshifts [39]. The relationship between the absorption edge position and x value is presented in Figure 2b. Surprisingly, the absorption edge of the CsPb0.9Fe0.1(Br0.9Cl0.1)3 NCs redshift slightly compared to that of the CsPbBr3. As x increases from 0 to 0.3, the absorption edge slightly redshifts first and then blueshifts. Mixed halide CsPb(Br1−xClx)3 perovskite NCs without the Fe doping are also synthesized by mixing PbCl2 (x mmol) and PbBr2 (1−x mmol) during the preparation. The absorption edge positions of the CsPb(Br1−xClx)3 NCs without the Fe2+ doping are added for comparison. The absorption edge position gradually blueshifts as x increases. All the absorption edges of the Fe2+-doped NCs redshift compared to those of the corresponding perovskite NCs without the Fe2+ doping.

3.2. Photoluminescence Properties

Photos of the CsPb1−xFex(Br1−xClx)3 NC solutions under natural daylight and UV light are shown in Figure 3a. As x changes from 0 to 0.3, the fluorescent color turns from green to cyan. The corresponding PL spectra of the CsPb1−xFex(Br1−xClx)3 NCs with x = 0 to 0.3 are displayed in Figure 3b. For x = 0, the PL peak position of the CsPbBr3 NCs appears at about 519 nm, which is in good agreement with previous work. Similar to the shift of the absorption edge, when x = 0.1, the PL peak position of the CsPb1−xFex(Br1−xClx)3 NCs redshifts to 521 nm. As x increases to 0.2 and 0.3, the PL peak positions apparently blueshift. When x = 0.3, the peak position appears at 490 nm, explaining the cyan fluorescent color. The PL spectra of the CsPb(Br1−xClx)3 NCs without the Fe2+ doping are shown in Figure 3c. The PL peak position monotonously blueshifts as the x value increases without any exception. The dependencies of the PL peak position on the x value for these two groups of perovskite NCs are plotted in Figure 3d. For the same x value, the PL peak position of the CsPb1−xFex(Br1−xClx)3 NCs always redshifts compared to that of the CsPb(Br1−xClx)3 NCs, which is again in good agreement with the shift of the absorption edge. Therefore, we can conclude that the Fe2+ doping results in the redshift of PL.
Many different effects can be responsible for the PL redshift. For example, the large size means a narrowed bandgap due to the size effect, which can explain the PL redshift [39]. Moreover, photon reabsorption is another possible reason for the PL redshift [40,41]. As revealed previously [41], evident photon reabsorption prevailingly exists in various halide perovskite materials due to their high absorption coefficient and small Stokes shift. The Fe2+ doping may increase the aggregation of NCs, leading to the PL redshift. However, the TEM images do not show either an increased size or increased aggregation, excluding these factors. In addition, the element doping may expand the lattice spacings, causing the shrinkage of the bandgap. However, here, the doping element is Fe2+ with a radius of ~0.76 Å, which is smaller than the Pb2+ with a radius of ∼1.33 Å. In fact, previous work has reported that the doping of Fe2+ in CsPbCl3 NCs results in a slight blueshift of the PL peak position [38]. Furthermore, the surface defects generally induce shallow traps, leading to a narrowing of the bandgap. The Fe2+ doping may induce shallow traps in the perovskite lattices, which leads to the PL redshift. This kind of possibility is evaluated by our further investigation.
The PL spectra of the CsPb1xFex(Br1xClx)3 NCs and CsPb(Br1xClx)3 NCs are separately compared in Figure 4a–c for x = 0.1 to 0.3. It can be clearly seen that, not only the PL peak position redshifts but also the bandwidth decreases in the Fe2+-doped samples. The full widths at half maximums (FWHMs) are compared in Figure 4d. Especially when x = 0.3, the FWHM of the PL of the CsPb(Br1xClx)3 NCs is about 26 nm, which decreases to about 18 nm after the Fe2+ doping. The PL bandwidth depends on both intrinsic effects, such as electron–phonon interactions, shallow defect states, and extrinsic factors, such as size polydispersity. Herein, the size distribution is little changed by the Fe2+ doping. Therefore, the reduced FWHM is mainly attributed to the intrinsic effects. It is well supposed that both the doping of Fe2+ improves the crystallinity and diminishes the defect states. To verify this point, the PL QYs of the perovskite NCs before and after the Fe2+ doping were measured and compared. As shown in Figure 4e, the PL QYs are improved after different concentrations of the Fe2+ doping. Therefore, both the FWHMs and PL QYs suggest the density of defects is reduced by the appropriate Fe2+ doping.
To further understand the PL of the CsPb1-xFex(Br1xClx)3 NCs, the time-resolved PL decay spectra were measured. As shown in Figure 5a, all the time-resolved PL decays are fitted by an exponential function. The PL lifetimes are shown in Figure 5b. The PL lifetimes of the undoped CsPb(Br1–xClx)3 NCs are also added for comparison. The PL lifetimes of the undoped CsPb(Br1−xClx)3 monotonously decrease as x increases, which has been widely reported previously [39]. However, for the CsPb1−xFex(Br1−xClx)3 NCs, the PL lifetimes increase to about 70 ns when x = 0.1 and 0.2, which are significantly longer than those of the undoped CsPb(Br1−xClx)3. The time-resolved PL decay spectra indicate the PL lifetimes of the mixed halide perovskite NCs are increased by the Fe doping. The defect trapping usually leads to a short PL lifetime [42,43]. Thus, a longer PL lifetime implies a lower defect density [42,43]. Therefore, the PL lifetime further suggests the defect density is decreased after the Fe2+ doping, which is in good agreement with the decreased FWHMs and improved PL QYs. Back to the PL peak positions shown in Figure 3, the redshift of the emission cannot be attributed to the shallow traps caused by the Fe2+ doping.
As is well known, the photostability of the CsPb(Br1−xClx)3 NCs is generally worse than that of the CsPbBrCl3 NCs due to the phase segregation [15,16,17]. After continuous irradiation, the mixed CsPb(Br1−xClx)3 NCs transform to a separated Br rich phase and Cl rich phase, and the PL shows typical redshifts and quenches largely. As shown in Figure 6a, the PL intensity of the CsPb(Br1−xClx)3 NCs quenches to about half of the original intensity after 1 W UV irradiation for 30 minutes. After the Fe2+ doping, we find the photostability is significantly improved. The PL intensity only decreases by ca.15% after UV irradiation for 30 minutes. The doping of the appropriate concentration of Fe2+ (∼0.76 Å) with an ionic radius smaller than Pb2+ (∼1.33 Å) can enhance the formation energies of perovskite lattices and, thus, essentially improve the structural stability [44,45]. Moreover, the stability usually depends on the crystal quality. The improved photostability is consistent with the decreased defect density.

4. Conclusions

In summary, the Fe2+-doped perovskite NCs are prepared by a hot injection method. In addition, their optical properties, including absorption, PL, and PL lifetimes are compared with those of the undoped CsPb(Br1−xClx)3 NCs. We find that Fe2+ doping results in the redshift of the absorption edge and PL. Moreover, the FWHMs are decreased and PL QYs are improved by the Fe2+ doping, suggesting the density of defects is reduced. The extended PL lifetimes further verify the defect density is decreased after the Fe2+ doping. Moreover, the photostability is significantly improved after the Fe doping. Therefore, this work reveals that Fe2+ doping is a very promising approach to modulate the optical properties of mixed halide perovskite NCs.

Author Contributions

Conceptualization, C.W. and Y.L.; methodology, Z.X., C.J. and Y.T.; writing draft, J.Z.; editing, C.M.; project administration, J.G. All authors have agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (No. 62004136), and Natural Science Foundation of the Jiangsu Higher Education Institutions of China (20KJB140019 and 19KJB150018). This work is supported by the Innovation and Entrepreneurship Training Program for College Students (202110332020Z). This work is supported by the Doctor of Entrepreneurship and Innovation in Jiangsu Province (No. (2020)30790).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All the experimental/calculation date that support the findings of this study are available from the corresponding authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jeon, N.J.; Na, H.; Jung, E.H.; Yang, T.-Y.; Lee, Y.G.; Kim, G.; Shin, H.-W.; Il Seok, S.; Lee, J.; Seo, J. A fluorene-terminated hole-transporting material for highly efficient and stable perovskite solar cells. Nat. Energy 2018, 3, 682–689. [Google Scholar] [CrossRef]
  2. Akihiro, K.; Kenjiro, T.; Yasuo, S.; Tsutomu, M. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050. [Google Scholar]
  3. Kim, D.; Jung, H.J.; Park, I.J.; Larson, B.W.; Dunfield, S.P.; Xiao, C.; Kim, J.; Tong, J.; Boonmongkolras, P.; Ji, S.G.; et al. Efficient, stable silicon tandem cells enabled by anion-engineered wide-bandgap perovskites. Science 2020, 368, 155–160. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, X.L.; Xu, B.; Zhang, J.B.; Gao, Y.; Zheng, Y.J.; Wang, K.; Sun, X.W. All-inorganic perovskite nanocrystals for high-efficiency light emitting diodes: Dual-phase CsPbBr3-CsPb2Br5 composites. Adv. Funct. Mater. 2016, 26, 4595. [Google Scholar] [CrossRef]
  5. Song, J.Z.; Li, J.H.; Li, X.M.; Xu, L.M.; Dong, Y.H.; Zeng, H.B. Quantum dot light-emitting diodes based on inorganic perovskite cesium lead halides (CsPbX3). Adv. Mater. 2015, 27, 7162. [Google Scholar] [CrossRef]
  6. Fang, H.J.; Li, J.W.; Ding, J.; Sun, Y.; Li, Q.; Sun, J.L.; Wang, L.D.; Yan, Q.F. An origami perovskite photodetector with spatial recognition ability. ACS Appl. Mater. Interfaces 2017, 9, 10921. [Google Scholar] [CrossRef]
  7. Cao, F.R.; Tian, W.; Wang, M.; Li, L. Polarized ferroelectric field-enhanced self-powered perovskite photodetector. ACS Photonics 2018, 5, 3731. [Google Scholar] [CrossRef]
  8. Ramasamy, P.; Lim, D.H.; Kim, B.; Lee, S.H.; Lee, M.S.; Lee, J.S. All-inorganic cesium lead halide perovskite nanocrystals for photodetector applications. Chem. Commun. 2016, 52, 2067. [Google Scholar] [CrossRef]
  9. Hao, F.; Stoumpos, C.C.; Cao, D.H.; Chang, R.P.H.; Kanatzidis, M.G. Lead-free solid-state organic−inorganic halide perovskite solar cells. Nat. Photonics 2014, 8, 489–494. [Google Scholar] [CrossRef]
  10. Volonakis, G.; Filip, M.R.; Haghighirad, A.A.; Sakai, N.; Wenger, B.; Snaith, H.J.; Giustino, F. Lead-free halide double perovskites via heterovalent substitution of noble metals. J. Phys. Chem. Lett. 2016, 7, 1254–1259. [Google Scholar] [CrossRef] [Green Version]
  11. Savory, C.N.; Walsh, A.; Scanlon, D.O. Pb-free halide double perovskites dupport high-efficiency solar cells. ACS Energy Lett. 2016, 1, 949–955. [Google Scholar] [CrossRef] [PubMed]
  12. Zhao, X.G.; Yang, J.H.; Fu, Y.H.; Yang, D.W.; Xu, Q.L.; Yu, L.P.; Wei, S.H.; Zhang, L.J. Lead-free inorganic halide perovskites for solar cells via cation-transmutation. J. Am. Chem. Soc. 2017, 139, 2630–2638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wu, X.L.; Xiong, S.J.; Liu, Z.; Chen, J.; Shen, J.C.; Li, T.H.; Wu, P.H.; Chu, P.K. Green light stimulates terahertz emission from mesocrystal microspheres. Nat. Nanotech. 2011, 6, 103–106. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, B.; Chen, J.S.; Hong, F.; Mao, X.; Zheng, K.B.; Yang, S.Q.; Li, Y.J.; Pullerits, T.; Deng, W.Q.; Han, K.L. Lead-free, air-stable all-inorganic cesium bismuth halide perovskite nanocrystals. Angew. Chem. Int. Ed. 2017, 56, 12471. [Google Scholar] [CrossRef]
  15. Unger, E.L.; Hoke, E.T.; Bailie, C.D.; Nguyen, W.H.; Bowring, A.R.; Heumuller, T.; Christoforo, M.G.; McGehee, M.D. Hysteresis and transient behavior in current-voltage measurements of hybrid-perovskite absorber solar cells. Energy Environ. Sci. 2014, 7, 3690–3698. [Google Scholar] [CrossRef]
  16. Yuan, Y.; Huang, J. Ion migration in organometal trihalide perovskite and its impact on photovoltaic efficiency and stability. Acc. Chem. Res. 2016, 49, 286–293. [Google Scholar] [CrossRef] [Green Version]
  17. Brennan, M.C.; Ruth, A.; Kamat, P.V.; Kuno, M. Photoinduced anion segregation in mixed halide perovskites. Trends Chem. 2020, 2, 282–301. [Google Scholar] [CrossRef] [Green Version]
  18. Zhao, R.; Yang, C.; Wang, H.G.; Jiang, K.; Wu, H.; Shen, S.P.; Wang, L.; Sun, Y.; Jin, K.J.; Gao, J.; et al. Emergent multiferroism with magnetodielectric coupling in EuTiO3 created by a negative pressure control of strong spin-phonon coupling. Nat. Commun. 2022, 13, 1–9. [Google Scholar]
  19. 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. [Google Scholar] [CrossRef] [Green Version]
  20. Grandhi, G.K.; Matuhina, A.; Liu, M.N.; Annurakshita, S.; Ali-Loytty, H.; Bautista, G.; Vivo, P. Lead-free cesium titanium bro mide double perovskite nanocrystals. Nanomaterials 2021, 11, 1458. [Google Scholar] [CrossRef]
  21. Noel, N.K.; Stranks, S.D.; Abate, A.; Wehrenfennig, C.; Guarnera, S.; Haghighirad, A.A.; Sadhanala, A.; Eperon, G.E.; Pathak, S.K.; Johnston, M.B.; et al. Lead-free organic–inorganic tin halide perovskites for photovoltaic applications. Energy Environ. Sci. 2014, 7, 3061–3068. [Google Scholar] [CrossRef]
  22. Wang, F.; Han, Y.; Lim, C.S.; Lu, Y.H.; Wang, J.; Xu, J.; Chen, H.Y.; Zhang, C.; Hong, M.H.; Liu, X.G. Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping. Nature 2010, 463, 1061. [Google Scholar] [CrossRef] [PubMed]
  23. Dong, H.; Sun, L.D.; Wang, Y.F.; Ke, J.; Si, R.; Xiao, J.W.; Lyu, G.M.; Shi, S.; Yan, C.H. Efficient tailoring of upconversion selectivity by engineering local structure of lanthanides in NaxREF3+x nanocrystals. J. Am. Chem. Soc. 2015, 137, 6569. [Google Scholar] [CrossRef]
  24. Feng, X.D.; Sayle, D.C.; Wang, Z.L.; Paras, M.S.; Santora, B.; Sutorik, A.C.; Sayle, T.X.T.; Yang, Y.; Ding, Y.; Wang, X.D.; et al. Converting ceria polyhedral nanoparticles into single-crystal nanospheres. Science 2006, 312, 1504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Begum, R.; Parida, M.R.; Abdelhady, A.L.; Murali, B.; Alyami, N.M.; Ahmed, G.H.; Hedhili, M.N.; Bakr, O.M.; Mohammed, O.F. Engineering interfacial charge transfer in CsPbBr3 perovskite nanocrystals by heterovalent doping. J. Am. Chem. Soc. 2017, 139, 731. [Google Scholar] [CrossRef]
  26. Mao, K.H.; Zhang, J.L.; Guo, Z.J.; Liu, L.Z.; Ma, H.; Chin, Y.Y.; Lin, H.J.; Bao, S.S.; Xie, H.Q.; Yang, R.; et al. Constructing asymmetrical Ni-centered {NiN2O4} octahedra in layered metal−organic structures for near-room-temperature single-phase magnetoelectricity. J. Am. Chem. Soc. 2020, 142, 12841–12849. [Google Scholar] [CrossRef] [PubMed]
  27. Guo, Z.J.; Mao, K.H.; Ma, G.D.; Li, G.A.; Wu, Q.F.; Chen, J.; Bao, S.S.; Yu, G.L.; Li, S.H.; Zhang, J.L.; et al. Light-Induced Tunable Ferroelectric Polarization in DipoleEmbedded Metal−Organic Framework. Nano Lett. 2022, 22, 10018–10024. [Google Scholar] [CrossRef]
  28. Jiang, Y.C.; He, A.P.; Luo, K.; Zhang, J.L.; Liu, G.Z.; Zhao, R.; Zhang, Q.; Wang, Z.; Zhao, C.; Wang, L.; et al. Giant bipolar unidirectional photomagnetoresistance. Proc. Natl. Acad. Sci. USA 2022, 27, 119. [Google Scholar] [CrossRef] [PubMed]
  29. Xia, Z.Y.; Li, Y.; Ji, C.; Jiang, Y.C.; Ma, C.L.; Gao, J.; Zhang, J.L. Slow relaxation behavior of a mononuclear Co(II) complex featuring long axial Co-O bond. Nanomaterials. 2022, 12, 707. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, W.Z.; Li, X.; Peng, C.C.; Yang, F.; Lian, L.Y.; Guo, R.D.; Zhang, J.B.; Wang, L. CsPb(Br/Cl)3 perovskite nanocrystals with bright blue emission synergistically modified by calcium halide and ammonium ion. Nanomaterials 2022, 12, 2026. [Google Scholar] [CrossRef]
  31. Klug, M.T.; Osherov, A.; Haghighirad, A.A.; Stranks, S.D.; Brown, P.R.; Bai, S.; Wang, J.T.W.; Dang, X.N.; Bulovic, V.; Snaith, H.J.; et al. Tailoring metal halide perovskites through metal substitution: Influence on photovoltaic and material properties. Science 2017, 10, 236–246. [Google Scholar] [CrossRef]
  32. Parobek, D.; Roman, B.J.; Dong, Y.; Jin, H.; Lee, E.; Sheldon, M.; Son, D.H. Exciton-to-dopant energy transfer in Mn-doped cesium lead halide perovskite nanocrystals. Nano Lett. 2016, 16, 7376. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, H.W.; Wu, Z.N.; Shao, J.R.; Yao, D.; Gao, H.; Liu, Y.; Yu, W.L.; Zhang, H.; Yang, B. CsPbxMn1–xCl3 perovskite quantum Dots with high Mn substitution ratio. ACS Nano 2017, 11, 2239. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, W.Y.; Lin, Q.L.; Li, H.B.; Wu, K.F.; Robel, I.; Pietryga, J.M.; Klimov, V.I. Mn2+-doped lead halide perovskite nanocrystals with dual-color emission controlled by halide content. J. Am. Chem. Soc. 2016, 138, 14954. [Google Scholar] [CrossRef] [PubMed]
  35. Ma, K.W.; Sheng, Y.H.; Wang, G.Z.; Zhang, X.W.; Di, Y.S.; Liu, C.H.; Yu, L.Y.; Dong, L.F.; Gan, Z.X. Stable and multicolor solid-state luminescence of Mn doped CsPb (Cl/Br)3perovskite quantum dots and its application in light-emitting diodes. J. Lumin. 2022, 243, 118622. [Google Scholar] [CrossRef]
  36. Yong, Z.J.; Guo, S.Q.; Ma, J.P.; Zhang, J.Y.; Li, Z.Y.; Chen, Y.M.; Zhang, B.B.; Zhou, Y.; Shu, J.; Gu, J.L.; et al. Doping-enhanced short-range order of perovskite nanocrystals for near-unity violet luminescence quantum yield. J. Am. Chem. Soc. 2018, 140, 9942–9951. [Google Scholar] [CrossRef] [Green Version]
  37. Cheng, X.H.; Jing, L.; Yuan, Y.; Du, S.J.; Zhang, J.; Zhan, X.Y.; Ding, J.X.; Shi, G.D.T.H. Fe2+/Fe3+ doped into MAPbCl3 single crystal: Impact on crystal growth and optical and photoelectronic properties. J. Phys. Chem. C 2019, 123, 1669–1676. [Google Scholar] [CrossRef]
  38. Hu, Y.; Zhang, X.Y.; Yang, C.Q.; Lia, J.; Wang, L. Fe2+doped in CsPbCl3 perovskite nanocrystals: Impact on the luminescence and magnetic properties. RSC Adv. 2019, 9, 33017–33022. [Google Scholar] [CrossRef] [Green Version]
  39. 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. [Google Scholar] [CrossRef] [Green Version]
  40. Gan, Z.X.; Wen, X.M.; Chen, W.J.; Zhou, C.H.; Yang, S.; Cao, G.Y.; Ghiggino, K.P.; Zhang, H.; Jia, B.H. The dominant energy transport pathway in halide perovskites: Photon recycling or carrier diffusion. Adv. Energy Mater. 2019, 9, 1900185. [Google Scholar] [CrossRef]
  41. Gan, Z.X.; Chen, W.J.; Yuan, L.; Cao, G.Y.; Zhou, C.H.; Huang, S.J.; Wen, X.M.; Jia, B.H. External stokes shift of perovskite nanocrystals enlarged by photon recycling. Appl. Phys. Lett. 2019, 114, 011906. [Google Scholar] [CrossRef]
  42. Wenger, B.; Nayak, P.K.; Wen, X.M.; Kesava, S.V.; Noel, N.K.; Snaith, H.J. Consolidation of the optoelectronic properties of CH3NH3PbBr3 perovskite single crystals. Nat. Commun. 2017, 8, 590. [Google Scholar] [CrossRef]
  43. Sheng, Y.H.; Liu, C.H.; Yu, L.Y.; Yang, Y.Y.; Hu, F.R.; Sheng, C.; Di, Y.S.; Dong, L.F.; Gan, Z.X. Microsteganography on all inorganic perovskite micro-platelets by direct laser writing. Nanoscale 2021, 13, 14450. [Google Scholar] [CrossRef]
  44. Zou, S.H.; Liu, Y.S.; Li, J.H.; Liu, C.P.; Feng, R.; Jiang, F.L.; Li, Y.X.; Song, J.Z.; Zeng, H.B.; Hong, M.C.; et al. Stabilizing cesium lead halide perovskite lattice through Mn (II) substitution for air-stable light-emitting diodes. J. Am. Chem. Soc. 2017, 139, 11443–11450. [Google Scholar] [CrossRef] [PubMed]
  45. Cherevkov, S.; Azizov, R.; Sokolova, A.; Nautran, V.; Miruschenko, M.; Arefina, I.; Baranov, M.; Kurdyukov, D.; Stovpiaga, E.; Golubev, V.; et al. Interface chemical modification between all-inorganic perovskite nanocrystals and porous silica microspheres for composite materials with improved emission. Nanomaterials 2021, 11, 119. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a,b) TEM image (a) and size distribution (b) of the CsPb(Br0.8Cl0.2)3 NCs. (c,d) TEM image and size distribution of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs. (e) Elemental mappings of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs. (f) Atomic ratio between Pb and Fe of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs.
Figure 1. (a,b) TEM image (a) and size distribution (b) of the CsPb(Br0.8Cl0.2)3 NCs. (c,d) TEM image and size distribution of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs. (e) Elemental mappings of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs. (f) Atomic ratio between Pb and Fe of the CsPb0.8Fe0.2(Br0.8Cl0.2)3 NCs.
Nanomaterials 13 00533 g001
Figure 2. (a) Absorption spectra of the CsPb1xFex(Br1xClx)3 NCs. (b) Absorption band edge positions of the CsPb1xFex(Br1xClx)3 NCs at different x. Absorption band edge positions of the CsPb1xFex(Br1xClx)3 are also plotted for comparison.
Figure 2. (a) Absorption spectra of the CsPb1xFex(Br1xClx)3 NCs. (b) Absorption band edge positions of the CsPb1xFex(Br1xClx)3 NCs at different x. Absorption band edge positions of the CsPb1xFex(Br1xClx)3 are also plotted for comparison.
Nanomaterials 13 00533 g002
Figure 3. (a) Photographs of CsPb1−xFex(Br1−xClx)3 NC solutions under natural light (top) and UV light (bottom). (b) Normalized PL spectra of CsPb1−xFex(Br1−xClx)3 NC solutions. (c) Normalized PL spectra of CsPb(Br1−xClx)3 NC solutions. (d) Comparison of PL peak positions of CsPb1−xFex(Br1−xClx)3 and CsPb(Br1−xClx)3 NC solutions at different x (x = 0, 0.1, 0.2, 0.3). The excitation wavelength kept as 405 nm.
Figure 3. (a) Photographs of CsPb1−xFex(Br1−xClx)3 NC solutions under natural light (top) and UV light (bottom). (b) Normalized PL spectra of CsPb1−xFex(Br1−xClx)3 NC solutions. (c) Normalized PL spectra of CsPb(Br1−xClx)3 NC solutions. (d) Comparison of PL peak positions of CsPb1−xFex(Br1−xClx)3 and CsPb(Br1−xClx)3 NC solutions at different x (x = 0, 0.1, 0.2, 0.3). The excitation wavelength kept as 405 nm.
Nanomaterials 13 00533 g003
Figure 4. (ac) Comparison of PL spectra of the CsPb1−xFex(Br1−xClx)3 and CsPb(Br1−xClx)3 NC solutions with x = 0.1 (a), 0.2 (b), and 0.3 (c). (d) FWHM of the two kinds of NCs with different x values. (e) PL QY of the two kinds of NCs with different x values.
Figure 4. (ac) Comparison of PL spectra of the CsPb1−xFex(Br1−xClx)3 and CsPb(Br1−xClx)3 NC solutions with x = 0.1 (a), 0.2 (b), and 0.3 (c). (d) FWHM of the two kinds of NCs with different x values. (e) PL QY of the two kinds of NCs with different x values.
Nanomaterials 13 00533 g004
Figure 5. (a) Time-resolved PL spectra of the CsPb1−xFex(Br1−xClx)3; (b) PL lifetime parameters of the CsPb1−xFex(Br1−xClx)3 obtained from numerical fitting on (a). PL lifetimes of the CsPb(Br1−xClx)3 are also plotted for comparison.
Figure 5. (a) Time-resolved PL spectra of the CsPb1−xFex(Br1−xClx)3; (b) PL lifetime parameters of the CsPb1−xFex(Br1−xClx)3 obtained from numerical fitting on (a). PL lifetimes of the CsPb(Br1−xClx)3 are also plotted for comparison.
Nanomaterials 13 00533 g005
Figure 6. (a) PL spectra of CsPb(Br1−xClx)3 NCs under continuous UV irradiation. (b) PL spectra of CsPb1−xFex(Br1−xClx)3 NCs under continuous UV irradiation. Insets: PL intensity versus irradiation time.
Figure 6. (a) PL spectra of CsPb(Br1−xClx)3 NCs under continuous UV irradiation. (b) PL spectra of CsPb1−xFex(Br1−xClx)3 NCs under continuous UV irradiation. Insets: PL intensity versus irradiation time.
Nanomaterials 13 00533 g006
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, C.; Li, Y.; Xia, Z.; Ji, C.; Tang, Y.; Zhang, J.; Ma, C.; Gao, J. Enhancing Photoluminescence of CsPb(ClxBr1−x)3 Perovskite Nanocrystals by Fe2+ Doping. Nanomaterials 2023, 13, 533. https://doi.org/10.3390/nano13030533

AMA Style

Wu C, Li Y, Xia Z, Ji C, Tang Y, Zhang J, Ma C, Gao J. Enhancing Photoluminescence of CsPb(ClxBr1−x)3 Perovskite Nanocrystals by Fe2+ Doping. Nanomaterials. 2023; 13(3):533. https://doi.org/10.3390/nano13030533

Chicago/Turabian Style

Wu, Chang, Yan Li, Zhengyao Xia, Cheng Ji, Yuqian Tang, Jinlei Zhang, Chunlan Ma, and Ju Gao. 2023. "Enhancing Photoluminescence of CsPb(ClxBr1−x)3 Perovskite Nanocrystals by Fe2+ Doping" Nanomaterials 13, no. 3: 533. https://doi.org/10.3390/nano13030533

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