Next Article in Journal
Solvent Dependent Disorder in M2(BzOip)2(H2O)·Solvate (M = Co or Zn)
Next Article in Special Issue
A Feasible and Effective Post-Treatment Method for High-Quality CH3NH3PbI3 Films and High-Efficiency Perovskite Solar Cells
Previous Article in Journal
Liquid Crystal Ordering of Four-Base-Long DNA Oligomers with Both G–C and A–T Pairing
Previous Article in Special Issue
Atomic Characterization of Byproduct Nanoparticles on Cesium Lead Halide Nanocrystals Using High-Resolution Scanning Transmission Electron Microscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Stability and Photoluminescence Yield of Mn2+-Doped CH3NH3PbCl3 Perovskite Nanocrystals

1
Department of Chemistry, Shantou University, Guangdong 515063, China
2
Department of Chemistry and Key Laboratory for Preparation and Application of Ordered Structural Materials of Guangdong Province, Shantou University, Guangdong 515063, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Crystals 2018, 8(1), 4; https://doi.org/10.3390/cryst8010004
Submission received: 21 November 2017 / Revised: 18 December 2017 / Accepted: 20 December 2017 / Published: 23 December 2017

Abstract

:
Organic–inorganic CH3NH3PbCl3 perovskite nanocrystals (PNCs) doped with Mn2+, CH3NH3PbxMn1−xCl3, have been successfully prepared using a reprecipitation method at room temperature. Structural and morphological characterizations reveal that the CH3NH3PbxMn1−xCl3 PNCs with cubic phase transforms from particles to cubes and increases in size from 16.2 ± 4.4 nm in average diameter to 25.3 ± 7.2 nm in cubic length after the addition of Mn2+ precursor. The CH3NH3PbxMn1−xCl3 PNCs exhibit a weak exciton emission at ~405 nm with a low absolute quantum yield (QY) of around 0.4%, but a strong Mn2+ dopant emission at ~610 nm with a high QY of around 15.2%, resulting from efficient energy transfer from the PNC host to the Mn2+ dopant via the 4T16A1 transition. In addition, the thermal and air stability of CH3NH3PbxMn1−xCl3 PNCs are improved due to the passivation with (3-aminopropyl) triethoxysilane (APTES), which is important for applications such as light emitting diodes (LEDs).

Graphical Abstract

1. Introduction

As a promising candidate for photovoltaics and lighting application, lead halide perovskite (LHP) afford many intriguing advantages over traditional semiconductor materials such as simple fabrication [1,2,3,4], abundant precursors source, defect tolerance [5,6], high photoluminescence (PL) quantum yields (QY) [5,7,8,9], halogen dependent emission over the entire visible spectral range with narrow full width at half maximum (FWHM) [10,11], as well as long carrier lifetime and diffusion length [12]. Compared with the intensive investigations of lead bromide and iodide perovskite [13,14,15,16,17], lead chloride perovskites have received less attention due to its low PL QY and large bandgap (~2.97 eV), which is less desirable for light emitting diodes (LEDs) and solar cells applications [18].
Doping metal ions into nanocrystals is an effective approach to tuning their optical, electronic and magnetic properties, as has been widely applied to II-VI metal chalcogenide nanocrystals [19,20]. Recently, several heterovalent or isovalent metal ions such as Bi3+ [21], Al3+ [22], Sn2+ [23], Cd2+ [23], Zn2+ [23], and Mn2+ [24,25,26,27,28,29,30] have been introduced into LHP as dopants to modulate their optical properties. Among these metal ions, Mn2+-doped all inorganic CsPbCl3 perovskite nanocrystals (PNCs) have attracted increasing attention owing to the appropriate band alignment of CsPbCl3 and d-d transition of Mn2+ (2.15 eV, 6A14T1), resulting in efficient intraparticle energy transfer between the host exciton and dopant ions [25,31]. Moreover, the incorporation of Mn2+ may provide the possibility of imparting paramagnetism to the host materials [31].
In general, hot injection approach has been applied in the preparation of Mn2+-doped lead halide PNCs because of the homogeneous products and relatively simple purification process [32,33]. However, the highly strict conditions of this method and poor stability of the resulting PNCs greatly hinder their applications. In contrast, organic-inorganic hybrid CH3NH3PbCl3 PNCs can be easily prepared via solution-based reprecipitation approach at room temperature when (3-aminopropyl) triethoxysilane (APTES) is used as a capping ligand [34,35]. Additionally, the SiO2 shell resulting from the hydrolysis of APTES capping ligand can greatly improve the product yield and stability of CH3NH3PbCl3 PNCs towards water and air [34].
In this work, we demonstrate that Mn2+ can be doped into CH3NH3PbCl3 PNCs via a reprecipitation method. By adjusting the doping concentration of Mn2+, CH3NH3PbxMn1−xCl3 PNCs with a strong dopant emission located at 610 nm were obtained. Besides the greatly improved overall QY of CH3NH3PbxMn1−xCl3 PNCs (from ~0.4% to ~15.6%) upon Mn2+ doping, the thermal and air stability have also been enhanced by the SiO2 shell formed from hydrolysis of APTES. The combined stability and high PL QY make these PNCs potentially useful for various photonics applications including LEDs.

2. Experimental Section

2.1. Materials

All the chemicals were used as received without further purification, including PbCl2 (99.99%, Aladdin Chemical Co., Ltd., Shanghai, China), CH3NH2 (40%, TCI Chemical Co., Ltd., Tokyo, Japan), MnCl2 (99%, Aladdin Chemical Co., Ltd., Shanghai, China), APTES (99%, Aladdin Chemical Co., Ltd., Shanghai, China), HCl (37.5%, Xilong Scientific Co., Ltd., Shantou, China), toluene (AR, Xilong Scientific Co., Ltd., Shantou, China), N,N-dimethylformamide (DMF, spectroscopic grade, 99.9%, Aladdin Chemical Co., Ltd., Shanghai, China), oleic acid (85%, TCI Chemical Co., Ltd., Tokyo, Japan), octylamine (99%, Fisher Scientific Co., Ltd., Pittsburgh, PA, US), ethyl acetate (99%, Aladdin Chemical Co., Ltd., Shanghai, China).
Mn2+-doped CH3NH3PbCl3 PNCs: CH3NH3Cl was prepared according to a reported procedure [18]. The Mn2+-doped CH3NH3PbCl3 PNCs were synthesized following a method reported previously with some small changes [1]. 1.5 mL precursor solution of PNCs was prepared by dissolving 0.05 mmol CH3NH3Cl, 0.05 mmol (PbCl2 + MnCl2), 0.05 mmol APTES or octylamine and 40 μL oleic acid in 1.5 mL DMF solvent and ultrasonicated until the solution became transparent. Then, 10 μL, 25 μL, 50 μL and 100 μL of the precursor solution were injected slowly into 5 mL toluene to form CH3NH3PbCl3 PNCs solution. Mn2+-doped CH3NH3PbCl3 PNCs with different doping concentration of Mn2+ were prepared by varying the ratio between PbCl2 and MnCl2. The solid samples were obtained by adding ethyl acetate in toluene to precipitate CH3NH3PbCl3 PNCs and centrifuged at 4000 rpm subsequently, then re-dispersed in toluene. These steps were repeated three times. At last, the dry samples were collected by placing the samples at 50 °C oven overnight.
Mn2+-doped CH3NH3PbCl3 PNCs films: The films were prepared by dropping purified solution on microslides and dried at room temperature.

2.2. Characterization

Absorption and photoluminescence properties of the samples were collected with circular dichroism spectra (MOS450, Biologic Science Instruments, Seyssinet-Pariset, France) and fluorescence spectra (PTI QM-TM, Photon Technology International, Ontario, Canada), respectively. Absolute QY was recorded on HAMAMATSU C11347 spectrometer (Hamamatsu Photonics Instruments, Hamamatsu City, Japan) by testing samples in toluene. Luminescence lifetime measurements were collected on EDINBURGH INSTRUMENTS (FLS920, Edinburgh Instruments Ltd., Livingston, UK). X-Ray diffraction (XRD, MiniFlex 600, Rigaku Corporation, Tokyo, Japan) analysis was used to obtain the crystalline phase. The scanning angle range was 10–50° (2θ) with a rate of 3°/min. Temperature-dependent Powder XRD (Ultima IV, Rigaku Corporation, Tokyo, Japan) were studied to indicate the thermal stability of samples. Transmission electron microscopy (TEM, Tecnai G2 F20, FEI Technologies, Hillsboro, OR, USA) were carried out to obtain the morphology and interlayer spacing of samples at an accelerating voltage of 200 kV. The elemental analysis was conducted on inductively coupled plasma atomic emission spectroscopy (ICP-AES, Agilent 720, Agilent Technologies, Santa Clara, CA, USA).

3. Results and Discussion

3.1. Structural and Morphological Characterizations

The structure and morphology of the prepared samples were determined by XRD and TEM measurements, respectively. For the nominal doping concentration of Mn2+ in 0~75 at.% range, the XRD pattern (Figure 1) shows a series of diffraction peaks that can all be attributed to cubic phase CH3NH3PbCl3 perovskite structure (space group: Pm 3 ¯ m), demonstrating that the incorporation of Mn2+ has little effect on the crystal structure of CH3NH3PbCl3 [18]. With the addition of 90 at. % MnCl2, some diffraction peaks attributed to CH3NH3Cl and MnCl2 precursors could be clearly observed due to the excess amount of precursors. Note that the broad band ranging from 15° to 38° is characteristic of amorphous silica resulting from the hydrolysis of APTES, which has been clearly elucidated in our previous work [34]. To further determine the concentration of Mn2+ dopant, ICP-AES was used, with results summarized in Table S1. In the synthesis of Mn2+-doped CH3NH3PbCl3 PNCs, it was found that APTES would react with MnCl2 to form a puce complex at a high concentration of Mn2+ precursor, as shown in Figure S1. The color of CH3NH3PbxMn1−xCl3 PNCs capped with APTES under room light is deeper than that of CH3NH3PbxMn1−xCl3 PNCs passivated with octylamine owing to the adsorption of Mn2+-APTES complex, indicating that the detected concentration in Table S1 are higher than the actual concentration of Mn2+ in the crystal lattice.
To determine the morphology evolution and size distribution, TEM images were recorded, as shown in Figure 2. For undoped CH3NH3PbCl3 PNCs (Figure 2a), the PNCs are mostly spherical with an average diameter of 16.2 ± 4.4 nm. The PNCs are embedded in amorphous silica matrix (Figure S2), which results from the hydrolysis of APTES. When MnCl2 precursor was added, the resulting Mn2+-doped PNCs are mostly cubic with an average dimension increasing from 10.4 ± 5.1 nm (Figure 2b) to 25.3 ± 7.2 nm (Figure 2e) as the Mn2+ concentration was increased from 25 to 90 at. %. Due to the complex reaction between APTES and the Mn2+ precursor, the concentration of APTES ligands decreased after adding MnCl2 precursor. Higher concentration of APTES can facilitate the formation of small spherical PNCs due to the large steric hindrance of branched APTES [34]. Likewise, with the increase of Mn2+ precursor concentration, the lower concentration of APTES will result in large cubic PNCs, which is consistent with the TEM results of this work. High-resolution TEM (HRTEM, Figure 2f) reveals that the lattice space of undoped PNCs is around 0.28 nm, corresponding to the (200) crystal plane of cubic CH3NH3PbCl3.

3.2. Optical Properties

Similar to Mn2+-doped all inorganic CsPbCl3 PNCs, new emission band may be introduced in Mn2+-doped organic-inorganic CH3NH3PbCl3 PNCs, as shown in the PL spectra in Figure 3, where the electronic absorption spectra are also shown. For undoped CH3NH3PbCl3 PNCs, the absorption spectrum exhibits a sharp rise starting around 400 nm and an excitonic peak at ~373 nm. In comparison, the excitonic peak positions of all the Mn2+-doped CH3NH3PbCl3 PNCs shift 4–10 nm towards long wavelength owing to a significant increase in average particle size. For the PL spectra, Mn2+-doped CH3NH3PbCl3 PNCs feature a broad emission band peaked at 610 nm, attributed to the spin-forbidden 4T16A1 transition of Mn2+ [31], along with a narrow luminescence band located at ~400 nm, which is assigned to intrinsic excitonic emission of the CH3NH3PbCl3 host. Compared to the Mn2+ dopant emission (~590 nm) in CsPbCl3 PNCs [28,31], Mn2+ dopant emission in the CH3NH3PbCl3 system significantly red shifts by ~20 nm, possibly caused by the crystal lattice difference between cubic CH3NH3PbCl3 and cubic/tetragonal CsPbCl3. The lattice parameters of cubic CH3NH3PbCl3 (space group: Pm 3 ¯ m, a = b = c = 5.685 Å) are a little larger than that of cubic CsPbCl3 (a = b = c = 5.605 Å) and tetragonal CsPbCl3 (a = b = 5.590 Å, c = 5.630 Å) [18,36], resulting in the smaller energy difference (ΔO) of Mn2+ d-orbitals.
In addition, all excitonic emissions of Mn2+-doped CH3NH3PbCl3 PNCs show a slight red-shift (~4 nm) with respect to that of undoped CH3NH3PbCl3 PNCs, consistent with red-shifted absorption spectra, attributed to the increase in size of PNCs. Intriguingly, with the increased concentration of Mn2+ dopant, the PL intensity of host increases first and then decreases, which is highly unusual. On one hand, the Mn2+ dopant ions may remove the preexisting structural defects and enhance the excitonic emission of hosts [25,37]. At the low doping concentration, both the intensity of excitionic emission and dopant emission are increased after 25 at. % Mn2+ doping, attributed to the dominant defect passivation of Mn2+ and exciton-to-Mn2+ energy transfer. On the other hand, the excitonic emission will be quenched upon the incorporation of Mn2+. With the addition of a higher Mn2+ concentration, the transfer of excitonic energy from the host to Mn2+ dopants becomes more dominant, giving rise to the decrease of excitonic emission intensity and enhancement of dopant emission intensity. Importantly, the overall PL QY was highly improved from 0.4% (undoped) to 15.6% (75 at. % Mn2+-doped), indicating efficient excitons-to-Mn2+ energy transfer, which is desired for light emitting applications.
To demonstrate that the Mn2+ dopant emission does not originate from the Mn2+-APTES complex, contrast test was conducted as shown in Figure S3. Without the formation of CH3NH3PbCl3 perovskite, no absorption and emission band can be observed. In the range of 380–700 nm, no emission can be seen for MnCl2 precursor, indicating that the Mn2+ dopant emission results from the host absorption. In addition, PLE spectra are collected to further illustrate the emission mechanism of Mn2+-doped CH3NH3PbCl3 PNCs. As shown in Figure S4, a sharp onset located at ~400 nm can be clearly observed for all Mn2+-doped CH3NH3PbCl3 PNCs, consistent with the absorption of undoped CH3NH3PbCl3 PNCs. These results display that the emission band centered at 610 nm originates from the absorption of CH3NH3PbCl3 PNCs, manifesting the energy transfer from host to dopants.
The volume ratio between precursor solution and anti-solvent (toluene) is an important factor in the synthesis of CH3NH3PbCl3 PNCs when adopting the reprecipitation strategy. Figure S5 displays the PL spectrum of Mn2+-doped CH3NH3PbCl3 PNCs solution prepared with different volume ratios. The Mn2+ dopant emission greatly increases in the range of 50–200 volume ratio after normalizing to excitonic emission, implying that more Mn2+ ions are incorporated into the crystal lattice of CH3NH3PbCl3. However, the intensity of dopant emission is diminished when reaching 500 toluene/precursor volume ratio, owing to the poor crystallinity of the host at high Mn2+ dopant concentration, thereby lower the transfer efficiency of photo-induced excitons between the host and dopants due to the competition of charge transfer to trap states [28].
To investigate the PL decay of CH3NH3PbxMn1−xCl3 PNCs, time-resolved PL spectra are obtained. All the PL decay kinetics were fitted with a single exponential curve, as shown in Figure 4. An expected decrease of excitonic lifetime from 2.40 ns to 2.04 ns can be observed with increasing the Mn2+ concentration from 0% to 75% (Figure 4a), due to the energy transfer from the host to Mn2+ dopants. Interestingly, the small change of exciton emission lifetime indicates the insignificant effect of dopant states on the overall exciton decay in the CH3NH3PbCl3 PNCs. Since the introduction of Mn2+ dopants provides a new and efficient decay pathway of the excitons, which leads to the radiative decay of Mn2+ via the 4T16A1 transition, the overall PL QY of the doped PNCs was improved [28].
PL decay curve of Mn2+ dopant emission located at 610 nm was also studied and presented in Figure 4b. A very long lifetime of 1.2 ms was observed resulting from the spin-forbidden nature of Mn2+ 4T16A1 transition, suggesting the successful incorporation of Mn2+ [28,38].

3.3. Stability Investigation

Although moisture stability of CH3NH3PbCl3 has been greatly improved, their practical applications are still restricted by their low decomposition temperature, which is determined by their low formation energies [39]. A recent study found that the formation energy of Mn2+-doped CsPbBr3 is calculated to be ~5% larger than that of pure CsPbBr3 using DFT simulation, suggesting the better thermal stability of CsPbBr3 after Mn2+ doping with a specific concentration [39]. In this work, the thermal stability of pure CH3NH3PbCl3 and CH3NH3PbxMn1−xCl3 PNCs has also been studied using temperature-dependent XRD, as presented in Figure 5. Without the addition of MnCl2 precursor, CH3NH3PbCl3 PNCs capped with octylamine (Figure 5a) show poor thermal stability. Some diffraction peaks due to impurities could be observed when the temperature was increased up to 170 °C. CH3NH3PbxMn1−xCl3 PNCs passivated with octylamine were relatively stable under 170 °C, but still to be decomposed at 180 °C (Figure 5b). In contrast, both CH3NH3PbCl3 PNCs and CH3NH3PbxMn1−xCl3 PNCs passivated with APTES show better thermal stability, as shown in Figure 5c,d. The decomposition temperature was significantly improved by 20–30 °C compared to PNCs passivated with octylamine. The improved thermal stability of CH3NH3PbCl3 PNCs passivated with APTES may be due to the SiO2 coating layer, which maintains the morphology and crystal structure of CH3NH3PbCl3 as a result of the higher thermal stability of SiO2.
The air stability of CH3NH3PbxMn1−xCl3 PNCs films was also studied, as shown in Figure S6. The PL intensity CH3NH3PbxMn1−xCl3 PNCs passivated with APTES and octylamine drop ~9% and ~73% after one week, indicating the better air stability of CH3NH3PbxMn1−xCl3 PNCs capped with APTES.

4. Conclusions

Doping of Mn2+ into organic-inorganic CH3NH3PbCl3 PNCs has been demonstrated using a versatile reprecipitation approach. The PL QY of the obtained samples are greatly improved after Mn2+ doping as a result of efficient energy transfer from the host to dopants. Furthermore, both thermal and air stability are enhanced due to the presence of SiO2 on the PNC surface. Importantly, compared to the Mn2+ dopant emission (~590 nm) in all inorganic CsPbCl3 system, the larger crystal lattice of CH3NH3PbCl3 results in a lower energy splitting of Mn2+ d-orbitals, thereby emitting at a longer wavelength (~610 nm). In this manner, Mn2+ dopant emission may be tuned by modulating the crystal lattice through substitution of different components in perovskites.

Supplementary Materials

Supplementary materials can be found at www.mdpi.com/2073-4352/8/1/4/s1, Figure S1: Photographs of 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs passivated with octylamine (left) and APTES (right) under room light (A) and 365 nm UV light (B); Figure S2: TEM image of CH3NH3PbCl3 PNCs passivated with APTES; Figure S3: (a) Absorption and (b) PL spectra of solutions prepared using MnCl2 and APTES precursors (black line), and using MnCl2, APTES, PbCl2 and CH3NH3Cl precursors (red line). (c) PL spectra of MnCl2ex = 360 nm); Figure S4: PLE spectra of Mn doped CH3NH3PbCl3 PNCs with different doping concentration (λem = 610 nm); Figure S5: PL spectra of 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs solutions with different toluene/precursor volume ratios; Figure S6: Air stability test of Mn2+-doped CH3NH3PbCl3 PNCs passivated with octylamine and APTES; Table S1. Elemental analysis of CH3NH3PbxMn1-xCl3 PNCs passivated with APTES under different doping concentration using ICP-AES.

Acknowledgments

This project was supported by the support of STU Scientific Research Foundation for Talents (NTF17001) and Science and technology planning project of Guangdong Province (No. 2014A020216045).

Author Contributions

X.L. and B.L. conceived and designed the experiments; X.L. and Y.G. performed the experiments and analyzed the data; B.L. wrote the paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, F.; Zhong, H.; Chen, C.; Wu, X.; Hu, X.; Huang, H.; Han, J.; Zou, B.; Dong, Y. Brightly luminescent and color-tunable colloidal CH3NH3PbX3 (X = Br, I, Cl) quantum dots: Potential alternatives for display technology. ACS Nano 2015, 9, 4533–4542. [Google Scholar] [CrossRef] [PubMed]
  2. Luo, B.; Naghadeh, S.B.; Allen, A.L.; Li, X.; Zhang, J.Z. Peptide-passivated lead halide perovskite nanocrystals based on synergistic effect between amino and carboxylic functional groups. Adv. Funct. Mater. 2017, 27, 1604018. [Google Scholar] [CrossRef]
  3. Xuan, T.; Yang, X.; Lou, S.; Huang, J.; Liu, Y.; Yu, J.; Li, H.; Wong, K.; Wang, C.; Wang, J. High stable CsPbBr3 quantum dots coated with alkyl phosphate for white light-emitting diodes. Nanoscale 2017, 9, 15286–15290. [Google Scholar] [CrossRef] [PubMed]
  4. Konstantakou, M.; Perganti, D.; Falaras, P.; Stergiopoulos, T. Anti-solvent crystallization strategies for highly efficient perovskite solar cells. Crystals 2017, 7, 291. [Google Scholar] [CrossRef]
  5. Dirin, D.N.; Protesescu, L.; Trummer, D.; Kochetygov, I.V.; Yakunin, S.; Krumeich, F.; Stadie, N.P.; Kovalenko, M.V. Harnessing defect-tolerance at the nanoscale: Highly luminescent lead halide perovskite nanocrystals in mesoporous silica matrixes. Nano Lett. 2016, 16, 5866–5874. [Google Scholar] [CrossRef] [PubMed]
  6. Huang, H.; Bodnarchuk, M.I.; Kershaw, S.V.; Kovalenko, M.V.; Rogach, A.L. Lead Halide perovskite nanocrystals in the research spotlight: Stability and defect tolerance. ACS Energy Lett. 2017, 2, 2071–2083. [Google Scholar] [CrossRef] [PubMed]
  7. Koscher, B.A.; Swabeck, J.K.; Bronstein, N.D.; Alivisatos, A.P. Essentially trap-free CsPbBr3 colloidal nanocrystals by postsynthetic thiocyanate surface treatment. J. Am. Chem. Soc. 2017, 139, 6566–6569. [Google Scholar] [CrossRef] [PubMed]
  8. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Nedelcu, G.; Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Grotevent, M.J.; Kovalenko, M.V. Fast anion-exchange in highly luminescent nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, I). Nano Lett. 2015, 15, 5635–5640. [Google Scholar] [CrossRef] [PubMed]
  10. Zhang, D.; Yang, Y.; Bekenstein, Y.; Yu, Y.; Gibson, N.A.; Wong, A.B.; Eaton, S.W.; Kornienko, N.; Kong, Q.; Lai, M.; et al. Synthesis of composition tunable and highly luminescent cesium lead halide nanowires through anion-exchange reactions. J. Am. Chem. Soc. 2016, 138, 7236–7239. [Google Scholar] [CrossRef] [PubMed]
  11. Vybornyi, O.; Yakunin, S.; Kovalenko, M.V. Polar-solvent-free colloidal synthesis of highly luminescent alkylammonium lead halide perovskite nanocrystals. Nanoscale 2016, 8, 6278–6283. [Google Scholar] [CrossRef] [PubMed]
  12. Huang, H.; Polavarapu, L.; Sichert, J.A.; Susha, A.S.; Urban, A.S.; Rogach, A.L. Colloidal lead halide perovskite nanocrystals: Synthesis, optical properties and applications. NPG Asia Mater. 2016, 8, e328. [Google Scholar] [CrossRef]
  13. Luo, B.; Pu, Y.-C.; Yang, Y.; Lindley, S.A.; Abdelmageed, G.; Ashry, H.; Li, Y.; Li, X.; Zhang, J.Z. Synthesis, optical properties, and exciton dynamics of organolead bromide perovskite nanocrystals. J. Phys. Chem. C 2015, 119, 26672–26682. [Google Scholar] [CrossRef]
  14. Tyagi, P.; Arveson, S.M.; Tisdale, W.A. Colloidal organohalide perovskite nanoplatelets exhibiting quantum confinement. J. Phys. Chem. Lett. 2015, 6, 1911–1916. [Google Scholar] [CrossRef] [PubMed]
  15. Akkerman, Q.A.; Park, S.; Radicchi, E.; Nunzi, F.; Mosconi, E.; De Angelis, F.; Brescia, R.; Rastogi, P.; Prato, M.; Manna, L. Nearly monodisperse insulator Cs4PbX6 (X = Cl, Br, I) nanocrystals, their mixed halide compositions, and their transformation into CsPbX3 nanocrystals. Nano Lett. 2017, 17, 1924–1930. [Google Scholar] [CrossRef] [PubMed]
  16. Noh, J.H.; Im, S.H.; Heo, J.H.; Mandal, T.N.; Seok, S.I. Chemical management for colorful, efficient, and stable inorganic-organic hybrid nanostructured solar cells. Nano Lett. 2013, 13, 1764–1769. [Google Scholar] [CrossRef] [PubMed]
  17. Johnston, M.B.; Herz, L.M. Hybrid perovskites for photovoltaics: charge-carrier recombination, diffusion, and radiative efficiencies. Acc. Chem. Res. 2016, 49, 146–154. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, Y.; Yang, Z.; Cui, D.; Ren, X.; Sun, J.; Liu, X.; Zhang, J.; Wei, Q.; Fan, H.; Yu, F.; et al. Two-inch-sized perovskite CH3NH3PbX3 (X = Cl, Br, I) crystals: Growth and characterization. Adv. Mater. 2015, 27, 5176–5183. [Google Scholar] [CrossRef] [PubMed]
  19. Fitzmorris, B.C.; Pu, Y.C.; Cooper, J.K.; Lin, Y.F.; Hsu, Y.J.; Li, Y.; Zhang, J.Z. Optical properties and exciton dynamics of alloyed core/shell/shell Cd(1-X)Zn(X)Se/ZnSe/ZnS quantum dots. ACS Appl. Mater. Interfaces 2013, 5, 2893–2900. [Google Scholar] [CrossRef] [PubMed]
  20. Zeng, R.; Zhang, T.; Dai, G.; Zou, B. Highly emissive, color-tunable, phosphine-free Mn:ZnSe/ZnS core/shell and Mn:ZnSeS shell-alloyed doped nanocrystals. J. Phys. Chem. C 2011, 115, 3005–3010. [Google Scholar] [CrossRef]
  21. 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, 731–737. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, M.; Zhong, G.; Yin, Y.; Miao, J.; Li, K.; Wang, C.; Xu, X.; Shen, C.; Meng, H. Aluminum-doped cesium lead bromide perovskite nanocrystals with stable blue photoluminescence used for display backlight. Adv. Sci. 2017, 1700335. [Google Scholar] [CrossRef] [PubMed]
  23. Van der Stam, W.; Geuchies, J.J.; Altantzis, T.; van den Bos, K.H.; Meeldijk, J.D.; Van Aert, S.; Bals, S.; Vanmaekelbergh, D.; de Mello Donega, C. Highly emissive divalent-ion-doped colloidal CsPb1−xMxBr3 perovskite nanocrystals through cation exchange. J. Am. Chem. Soc. 2017, 139, 4087–4097. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, C.C.; Xu, K.Y.; Wang, D.; Meijerink, A. Luminescent manganese-doped CsPbCl3 perovskite quantum dots. Sci. Rep. 2017, 7, 45906. [Google Scholar] [CrossRef] [PubMed]
  25. 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, 7376–7380. [Google Scholar] [CrossRef] [PubMed]
  26. Das Adhikari, S.; Dutta, S.K.; Dutta, A.; Guria, A.K.; Pradhan, N. Chemically tailoring the dopant emission in Manganese-doped CsPbCl3 perovskite nanocrystals. Angew. Chem. 2017, 56, 8746–8750. [Google Scholar] [CrossRef] [PubMed]
  27. Huang, G.; Wang, C.; Xu, S.; Zong, S.; Lu, J.; Wang, Z.; Lu, C.; Cui, Y. Postsynthetic doping of MnCl2 molecules into preformed CsPbBr3 perovskite nanocrystals via a halide exchange-driven cation exchange. Adv. Mater. 2017, 29. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, H.; Wu, Z.; Shao, J.; Yao, D.; Gao, H.; Liu, Y.; Yu, W.; Zhang, H.; Yang, B. CsPbxMn1-xCl3 perovskite quantum dots with high mn substitution ratio. ACS Nano 2017, 11, 2239–2247. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, Q.; Zhang, X.; Jin, Z.; Zhang, J.; Gao, Z.; Li, Y.; Liu, S.F. Energy-Down-Shift CsPbCl3: Mn quantum dots for boosting the efficiency and stability of perovskite solar cells. ACS Energy Lett. 2017, 2, 1479–1486. [Google Scholar] [CrossRef]
  30. Mir, W.J.; Jagadeeswararao, M.; Das, S.; Nag, A. Colloidal mn-doped cesium lead halide perovskite nanoplatelets. ACS Energy Lett. 2017, 537–543. [Google Scholar] [CrossRef]
  31. Liu, W.; Lin, Q.; Li, H.; Wu, K.; 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, 14954–14961. [Google Scholar] [CrossRef] [PubMed]
  32. Shamsi, J.; Dang, Z.; Bianchini, P.; Canale, C.; Stasio, F.D.; Brescia, R.; Prato, M.; Manna, L. colloidal synthesis of quantum confined single crystal CsPbBr3 nanosheets with lateral size control up to the micrometer range. J. Am. Chem. Soc. 2016, 138, 7240–7243. [Google Scholar] [CrossRef] [PubMed]
  33. Bekenstein, Y.; Koscher, B.A.; Eaton, S.W.; Yang, P.; Alivisatos, A.P. Highly luminescent colloidal nanoplates of perovskite cesium lead halide and their oriented assemblies. J. Am. Chem. Soc. 2015, 137, 16008–16011. [Google Scholar] [CrossRef] [PubMed]
  34. Luo, B.; Pu, Y.C.; Lindley, S.A.; Yang, Y.; Lu, L.; Li, Y.; Li, X.; Zhang, J.Z. Organolead halide perovskite nanocrystals: Branched capping ligands control crystal size and stability. Angew. Chem. 2016, 55, 8864–8868. [Google Scholar] [CrossRef] [PubMed]
  35. Sun, C.; Zhang, Y.; Ruan, C.; Yin, C.; Wang, X.; Wang, Y.; Yu, W.W. Efficient and stable white leds with silica-coated inorganic perovskite quantum dots. Adv. Mater. 2016, 28, 10088–10094. [Google Scholar] [CrossRef] [PubMed]
  36. Moller, C.K. A phase transition in cesium plumbochloride. Nature 1957, 180, 981–982. [Google Scholar] [CrossRef]
  37. Rossi, D.; Parobek, D.; Dong, Y.; Son, D.H. Dynamics of exciton-mn energy transfer in mn-doped CsPbCl3 perovskite nanocrystals. J. Phys. Chem. C 2017, 121, 17143–17149. [Google Scholar] [CrossRef]
  38. Guria, A.K.; Dutta, S.K.; Adhikari, S.D.; Pradhan, N. Doping Mn2+ in lead halide perovskite nanocrystals: Successes and challenges. ACS Energy Lett. 2017, 2, 1014–1021. [Google Scholar] [CrossRef]
  39. Zou, S.; Liu, Y.; Li, J.; Liu, C.; Feng, R.; Jiang, F.; Li, Y.; Song, J.; Zeng, H.; Hong, M.; 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]
Figure 1. XRD pattern of Mn2+-doped CH3NH3PbCl3 PNCs with different concentration of Mn2+ (nominal concentration). The asterisks and circles denote the characteristic XRD peaks of CH3NH3Cl and MnCl2, respectively.
Figure 1. XRD pattern of Mn2+-doped CH3NH3PbCl3 PNCs with different concentration of Mn2+ (nominal concentration). The asterisks and circles denote the characteristic XRD peaks of CH3NH3Cl and MnCl2, respectively.
Crystals 08 00004 g001
Figure 2. TEM images of Mn2+-doped CH3NH3PbCl3 PNCs doped with different Mn2+ concentrations. (a) 0 at. % Mn2+, (b) 25 at. % Mn2+, (c) 50 at. % Mn2+, (d) 75 at. % Mn2+, (e) 90 at. % Mn2+ and (f) HRTEM image of 0 at. % Mn2+.
Figure 2. TEM images of Mn2+-doped CH3NH3PbCl3 PNCs doped with different Mn2+ concentrations. (a) 0 at. % Mn2+, (b) 25 at. % Mn2+, (c) 50 at. % Mn2+, (d) 75 at. % Mn2+, (e) 90 at. % Mn2+ and (f) HRTEM image of 0 at. % Mn2+.
Crystals 08 00004 g002
Figure 3. Absorption and PL spectra of Mn2+-doped CH3NH3PbCl3 PNCs solution with different nominal Mn2+ doping concentration. Inset: Enlarged views of PL spectra located in 375–450 nm range and digital pictures of Mn2+-doped CH3NH3PbCl3 PNCs under 365 nm UV light.
Figure 3. Absorption and PL spectra of Mn2+-doped CH3NH3PbCl3 PNCs solution with different nominal Mn2+ doping concentration. Inset: Enlarged views of PL spectra located in 375–450 nm range and digital pictures of Mn2+-doped CH3NH3PbCl3 PNCs under 365 nm UV light.
Crystals 08 00004 g003
Figure 4. PL decay spectrum of Mn2+-doped CH3NH3PbCl3 PNCs films prepared with different MnCl2 concentration. (a) Monitored at exciton emission (~410 nm) and (b) Mn2+ dopant emission of 50 at.% Mn (~610 nm).
Figure 4. PL decay spectrum of Mn2+-doped CH3NH3PbCl3 PNCs films prepared with different MnCl2 concentration. (a) Monitored at exciton emission (~410 nm) and (b) Mn2+ dopant emission of 50 at.% Mn (~610 nm).
Crystals 08 00004 g004
Figure 5. Temperature-dependent Powder XRD test of (a) CH3NH3PbCl3 PNCs passivated with octylamine, (b) 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs passivated with octylamine, (c) CH3NH3PbCl3 PNCs passivated with APTES and (d) 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs passivated with APTES.
Figure 5. Temperature-dependent Powder XRD test of (a) CH3NH3PbCl3 PNCs passivated with octylamine, (b) 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs passivated with octylamine, (c) CH3NH3PbCl3 PNCs passivated with APTES and (d) 50 at.% Mn2+-doped CH3NH3PbCl3 PNCs passivated with APTES.
Crystals 08 00004 g005

Share and Cite

MDPI and ACS Style

Li, X.; Guo, Y.; Luo, B. Improved Stability and Photoluminescence Yield of Mn2+-Doped CH3NH3PbCl3 Perovskite Nanocrystals. Crystals 2018, 8, 4. https://doi.org/10.3390/cryst8010004

AMA Style

Li X, Guo Y, Luo B. Improved Stability and Photoluminescence Yield of Mn2+-Doped CH3NH3PbCl3 Perovskite Nanocrystals. Crystals. 2018; 8(1):4. https://doi.org/10.3390/cryst8010004

Chicago/Turabian Style

Li, Xianli, Yan Guo, and Binbin Luo. 2018. "Improved Stability and Photoluminescence Yield of Mn2+-Doped CH3NH3PbCl3 Perovskite Nanocrystals" Crystals 8, no. 1: 4. https://doi.org/10.3390/cryst8010004

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