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

Controllable one-step doping synthesis for the white-light emission of cesium copper iodide perovskites

Open Access Open Access

Abstract

In this paper, a controllable one-step doping method has been successfully adopted in the cesium copper iodide perovskite’s luminescence, a high-quality white-light emission with Commission Internationale de l´Eclairage (CIE) coordinates of (0.3397, 0.3325), and a color rendering index (CRI) reaching up to 90 was realized in a convenient way. Through adding impurities into the Cs3Cu2I5 system, high efficiency and stable CsCu2I3 was synthesized, and the coexistence of varied high luminescence phases realized the white lighting. Strikingly, blue-emitting Cs3Cu2I5 and yellow-emitting CsCu2I3 could coexist, and their respective luminescence was not interacted in the compound, which was beneficial for acquiring a single emission and highly efficient white lighting. This work carried out a deep exploration of the Cu-based metal halides, and would be favorable to the applications of lead-free perovskites.

© 2021 Chinese Laser Press

1. INTRODUCTION

Halide perovskites are prospective materials for optoelectronics because of the remarkable properties, such as low manufacturing cost, high luminescence efficiency, and tunable light emission characters due to their abundant variety of compositions [14]. Among the numerous applications, the white-light emission of organic-inorganic hybrids and all inorganic halide perovskites have received special attention [57]. Through flexible regulation of halogen elements in CsPbX3 (X = Cl, Br, I), the white emission with a diverse color temperature and index can be generated [8,9]. However, the anion exchange derived from mixtures of different halide perovskites, and the relatively narrow coverage of lead (Pb) in halide-based perovskites’ luminescent spectra are important factors hindering the development of white lighting [1012]. Meanwhile, the problems of thermal instability in organic systems and toxicity caused by the heavy Pb limit their further commercial applications [13,14]. In view of this situation, all-inorganic lead-free metal halide perovskites will have more promising prospects in an efficient, stable, and eco-friendly white lighting.

Control of material dimensionality enables them to form various structure types and light emission features [1517]. The typical 3D and 2D metal halides have been widely investigated because of their excellent optical and electronic characteristics. Nevertheless, low-dimensional 1D and 0D materials exhibit more unique photophysical properties, such as a large Stokes shift, broadband emission, and high photoluminescence quantum yield (PLQY) due to the self-trapped excitons or excited state structural reorganization [1820]. Recently, Cu-based metal halide perovskites are gradually coming into view because of their abundance, low environmental impact, high efficiency, and low-dimensional structures. For instance, green emission CsCuCl3 and Cs2CuCl4 nanocrystals [21], blue-green luminescence Cs2CuX4 (X = Cl, Br, Br/I) perovskite quantum dots [22], blue light emission Cs3Cu2Br5xIx (0x5) with near-unity PLQY [23], and CsCu2X3 (X =Cl, Br, I) with improved air and thermal stability [24] have been recently reported. Specifically, Cs3Cu2I5 with a 0D electronic structure, large Stokes shift, and strong blue emission has been fabricated and applied for LED devices [25,26]. Another phase of cesium copper iodide, CsCu2I3, shows broadband yellow emission [24], and the stable yellow LED based on it was successfully realized [27]. Additionally, CsCu2I3 single crystals with a high-PLQY due to strongly localized 1D excitonic recombination have been reported [28]. As both Cs3Cu2I5 and CsCu2I3 are pure iodide phases, the general mixed halide’s exchange existing in perovskite’s mixtures could be avoided, and pure white emission can be further achieved by appropriate mixing of these phases. Previously, a pure white-light emission has been successfully realized by adjusting the appropriate mixing ratio of these two phases [29,30]. Nevertheless, these methods always needed to fabricate two pure phases of blue emission Cs3Cu2I5 and yellow emission CsCu2I3, and then precisely control the mixing ratio of these individual luminescent materials, which was intricate and inconvenient for the practical application. Therefore, developing a simple intrinsic white-light emission of a cesium copper iodide system will become an effective approach. It has been reported that through a controllable CsI–CuI phase transformation by solvent treatment, stable CsCu2I3 was obtained from Cs3Cu2I5, and a single white-emission layer could be prepared [31]. In addition, Shi’s group has reported the electroluminescent white-light emitting diodes in terms of Cu-based halide materials [32].

In this paper, we have successfully prepared the white luminescent material by a one-step doping method, achieving full coverage of the visible spectrum. Through adding impurities into the Cs3Cu2I5 system, high efficiency and stable CsCu2I3 was successfully synthesized, and the coexistence of varied high luminescence phases realized the white lighting. By means of the disposable preparation of Cs3Cu2I5 and CsCu2I3 phases, a high-quality and more uniform white luminescence with CIE coordinates of (0.3397, 0.3325) and a CRI reaching 90 could be generated in a simple way.

2. MATERIALS AND METHODS

Cesium iodide (CsI, 99.9%), copper (I) iodide (CuI, 99.999%), neodymium iodide (NdI3, 99.9%), terbium iodide (TbI3, 99.99%), praseodymium iodide (PrI3, 99.9%), bismuth iodide (BiI3, 99.9%), N,N-dimethylformamide (DMF, 99.9%), and isopropanol (99.5%) were directly used without further purification.

The cesium copper iodide perovskites were synthesized via an antisolvent infiltration method, which was performed at room temperature by adding the precursor solution within a good solvent into a nonpolar poor solvent. The blend of two various solvents induced a transient supersaturation, leading to the nucleation and form of perovskites. Cesium iodide, copper (I) iodide, and impurity materials (in this paper, NdI3, TbI3, PrI3, and BiI3 were used as impurities, respectively) in different molar ratios were firstly dissolved in DMF to get a precursor solution. Then, the solution was rapidly dropped into the antisolvent of isopropanol to form a precipitate, and the resulting products were the mixture of Cs3Cu2I5 and CsCu2I3. With the increase in the molar ratio of doping materials, a mixture with different proportions of Cs3Cu2I5 and CsCu2I3 could be obtained. The specific synthesis procedures are as below.

Adding NdI3 into the Cs3Cu2I5 system. In the synthesis of molar ratios of 9 mol%, CsI (0.6 mmol), CuI (0.364 mmol), and NdI3 (0.036 mmol) were dissolved in DMF (4 mL). The mixture was stirred for 2 h at 80°C, then we let it cool naturally to room temperature. Next the precursor solution was rapidly added into isopropanol (20 mL) with vigorous stirring under air ambient at room temperature, and a precipitate was produced immediately during this process. Then, the resulting precipitate was filtered and washed with isopropanol; the yield of the product was about 75%. The other three concentrations were obtained by three corresponding molar doses: 3 mol% (0.6 mmol CsI, 0.388 mmol CuI, and 0.012 mmol NdI3), 5 mol% (0.6 mmol CsI, 0.38 mmol CuI, and 0.02 mmol NdI3), and 7 mol% (0.6 mmol CsI, 0.372 mmol CuI, and 0.028 mmol NdI3).

Adding TbI3, PrI3, and BiI3 into a Cs3Cu2I5 system. The impurities were chosen as 9 mol% (0.6 mmol CsI, 0.364 mmol CuI, and 0.036 mmol for TbI3, PrI3, and BiI3, respectively); the synthesis process was the same as above.

Adding NdI3 into the CsCu2I3 system. The impurities were chosen as 9 mol% (0.3 mmol CsI, 0.546 mmol CuI, and 0.054 mmol NdI3), and the synthesis process was the same as above.

Synthesis of pure Cs3Cu2I5: 0.6mmol CsI and 0.4 mmol CuI were used; the synthesis process was the same as above.

Synthesis of pure CsCu2I3: 0.3mmol CsI and 0.6 mmol CuI were used; the synthesis process was the same as above.

Photoluminescence (PL), photoluminescence excitation (PLE), and photoluminescence quantum yield (PLQY) measurements were performed at ambient temperature by an FS5 fluorescence spectrometer equipped with a xenon lamp and an integrating sphere. Powder X-ray diffraction (PXRD) measurements were performed by the Rigaku MiniFlex600 system equipped with a Cu-Kα radiation source (λ=1.5418μm). All scans were performed at room temperature with a step size of 0.02˚. X-ray photoelectron spectroscopy (XPS) analyses were conducted using an ESCALAB 250Xi spectrometer. Scanning electron microscope (SEM) measurements were performed by a ZEISS SUPRA 55 from Carl Zeiss, Germany. High resolution transmission electron microscopy (HRTEM) images were measured by the JEM-3200FS (JEOL).

3. RESULTS AND DISCUSSION

The Cs3Cu2I5 system with added NdI3 is first detected by the SEM and TEM techniques. The SEM image [Fig. 1(a)] implies the large quantity and good uniformity of the Cs3Cu2I5 and CsCu2I3 phases. Energy-dispersive X-ray spectroscopy (EDS) mapping [Fig. 1(b)] confirms that Cs, Cu, I, and Nd are evenly distributed in the material. The SEM and TEM [Fig. 1(c)] images of the as-synthesized Cs3Cu2I5 and CsCu2I3 composites show a dual-phase morphology characteristic composed of nanoparticles and nanorods, which correspond to Cs3Cu2I5 and CsCu2I3, respectively. Such morphology features can effectively avoid the color changes of the composites caused by anion exchange, because the two components exist individually without the formation of compact structure configuration, which will make it quite fit to emit stable white light. The TEM image [Fig. 1(c)] shows the micro morphologies of Cs3Cu2I5 and CsCu2I3, in which Cs3Cu2I5 crystallizes as nanoparticles with an average diameter of 300nm, whereas CsCu2I3 appears as nanorods with the average length of 2μm and width of 450nm. The TEM lattice fringes and fast Fourier transform (FFT) images of d-spacing 3.24Å and 2.1Å (1 Å = 0.1 nm) coincide with the lattice planes of (213) and (350) for Cs3Cu2I5 and CsCu2I3, respectively [Figs. 1(d) and 1(e)]. The X-ray photoelectron spectroscopy (XPS) measurements are conducted to validate the chemical states of ions [Figs. 1(f)–1(i)]. The binding energies of 951.8 eV and 932 eV coincide to Cu 2p1/2 and Cu 2p3/2, respectively, which demonstrates the monovalent state for copper in the material. Meanwhile, Cs 3d and I 3d correspond to the +1 and 1 states, which are consistent with published data for cesium copper iodides [25,29]. These results indicate that a system including both high quality Cs3Cu2I5 and CsCu2I3 has been successfully prepared through the doping method.

 figure: Fig. 1.

Fig. 1. (a) SEM, (b) EDS mapping, and (c) TEM images of the sample with NdI3. The lattice planes and fast Fourier transform (FFT) images of (d) Cs3Cu2I5 nanoparticles and (e) CsCu2I3 nanorods. (f) XPS analysis of the sample with 9 mol% NdI3, and the respective spectra of (g) Cs 3d, (h) Cu 2p, and (i) I 3d.

Download Full Size | PDF

Figure 2(a) demonstrates the PL (photoluminescence) spectra of pure Cs3Cu2I5 and CsCu2I3, in which Cs3Cu2I5 nanoparticles exhibit blue luminescence at 446 nm with a broad FWHM (full width at half maximum) of about 80 nm, while CsCu2I3 nanorods display yellow emitting at 576 nm with an FWHM of approximately 120 nm. Both samples show large Stokes shifts, and in this way self-absorption by the system can be efficiently prevented. The emissions are possible mostly owing to the self-trapped excitons, deriving from the Jahn-Teller distortion or strong exciton-phonon coupling [23,33]. In addition, broad emissions are usually helpful for enhancing the CRI of white lighting. The PLQYs are ∼69.94% for pure Cs3Cu2I5 and 9.91% for CsCu2I3, respectively, and the higher PLQY of Cs3Cu2I5 nanoparticles than CsCu2I3 nanorods may result from the stronger exciton binding energy in 0D Cs3Cu2I5. The PL spectra of the Cs3Cu2I5 system with various ratios of NdI3 under excitation of 316 nm are shown in Fig. 2(b). At the ratio of 3 mol%, a PL emission involving both Cs3Cu2I5 and CsCu2I3 has been detected in the material. With the increase of molar ratios, the intensity difference between the emission of CsCu2I3 and Cs3Cu2I5 diminishes, indicating that the proportion of CsCu2I3 gradually increases. As Cs3Cu2I5 and CsCu2I3 possess comprehensive red, green, and blue components, a mixture of them could form a white lighting. When the intensities of yellow emission from CsCu2I3 and blue emission from Cs3Cu2I5 reach an almost equal ratio, a normal white lighting can be produced. From Fig. 2(b), it can be seen that at the ratio of 9 mol%, a white emission has been successfully formed. Figure 2(d) shows that the CIE coordinates of it are (0.3397, 0.3325), approaching the standard white light of (0.33, 0.33). In addition, the CRI of this material can reach up to 90 and its PLQY is 16.9%. The PLE (photoluminescence excitation) spectrum for the sample with concentration of 9 mol% is shown in Fig. 2(c). It can be seen that under the excitation wavelength of 316 nm, the PLE efficiency for CsCu2I3 and Cs3Cu2I5 is almost the same, indicating that a white emission can be motivated at this wavelength, which is consistent with the 9 mol% white PL spectra in Fig. 2(b). Apparently, at this wavelength both of the PL efficiencies for CsCu2I3 and Cs3Cu2I5 do not reach the optimum, but the yellow emission from CsCu2I3 and blue emission from Cs3Cu2I5 reach a balance, and a normal white emission can be produced. Figure 2(c) also demonstrates that although the PLQY of Cs3Cu2I5 is much higher than that of CsCu2I3, only the PL efficiency of Cs3Cu2I5, which has the same amount as that of CsCu2I3, will be effective enough to make a contribution to the form of white lighting. As a result, the emitting intensities will nearly achieve a balance. The large amount of CsCu2I3 in the material with the ratio of 9 mol% will lead to the increase of the balanced effective PL efficiency and the overall powerful white luminescence. To further investigate how the doping ratios influence the formation of CsCu2I3 and the proportion between Cs3Cu2I5 and CsCu2I3, the XRD results are shown in Fig. 2(e). At the low ratio of 3 mol%, the corresponding diffraction peaks of Cs3Cu2I5 and CsCu2I3 appear simultaneously, indicating that both of the two phases are synthesized in the material. Meanwhile, the weak peak intensity of CsCu2I3 signifies that the proportion of it is tiny compared to another phase of Cs3Cu2I5. With the increase of dopant, the diffraction peak of CsCu2I3 is enhanced, while the diffraction from Cs3Cu2I5 is gradually weakened, which means that the amount of CsCu2I3 increases accordingly. This result is consistent with the PL spectra [Fig. 2(b)], along with the increase of the doping molar ratio, and the PL intensity difference between emission of CsCu2I3 and Cs3Cu2I5 gradually diminishes.

 figure: Fig. 2.

Fig. 2. (a) PL spectra of pure Cs3Cu2I5 and CsCu2I3. (b) PL spectra of samples with different concentration of NdI3 under excitation of 316 nm. (c) PLE spectra of emission peaks corresponding to Cs3Cu2I5 and CsCu2I3 of the sample with 9 mol% NdI3. (d) CIE coordinates of the sample with 9 mol% NdI3. (e) XRD diffraction patterns of samples with different concentration of NdI3 compared to the standard XRD patterns of Cs3Cu2I5 and CsCu2I3.

Download Full Size | PDF

It has been reported that Cs3Cu2I5 crystallizes in the Pnma space group of an orthorhombic crystal system. Two types of Cu+ appear as tetrahedral sites of Cu+I4 and triangular sites of Cu+I3 in the material, which are edge-connected coming into the isolated [Cu2I5]3 units [34,35]. In particular, these units are divided by the nearby Cs+ ions, forming a zero-dimensional structure [Fig. 3(a)]. In contrast, CsCu2I3 crystallizes in the Cmcm space group of an orthorhombic system, in which Cu+I4 tetrahedra sites share common faces and edges and form double chains of Cu2I3 along the c-axis separated by cesium, leading to a 1D structure [Fig. 3(b)]. In the research of lead bromine perovskite, it has been reported that CsPbBr3 can gradually transform into another phase of Cs4PbBr6 by adding ZnBr2 as a revulsive. The process of “survival of the fittest” occurs in CsPbBr3, in which the good quality and stable CsPbBr3 is retained, while the unstable CsPbBr3 is decomposed and likely ripened into the 416-structure by adding the ZnBr2 revulsive [36]. Like the role of ZnBr2 in the system of Cs-Pb-Br, the mechanism of the doping process in the transformation of CsCu2I3 and Cs3Cu2I5 can be attributed to the role the impurities played as a conversion revulsive. The 1D CsCu2I3 is more stable and easier to generate than the 0D Cs3Cu2I5. When the dopants are added into the Cs3Cu2I5 system, the balance of the whole system will be destroyed, in this process the stable Cs3Cu2I5 nanocrystals are retained, while the unstable particles with poor quality will be disintegrated and rapidly formed into the CsCu2I3 structure. The resulting mixture of these two phases can be obtained on this basis. However, when the circumstance is in reverse, namely materials are added into the CsCu2I3 system, another phase of the Cs3Cu2I5 does not appear simultaneously (Fig. 4). It can be seen that in every doping concentration, only PL and XRD peaks of CsCu2I3 and undissolved CuI (PL emission of 420 nm) exist in the material. The form of nanocrystals requires that the precursor concentration exceeds the critical concentration of nucleation, and a large number of nuclei occur in a very short period of time. When precursor concentration is lower, the nanoparticles will dissolve and disappear, and thus the 0D Cs3Cu2I5 cannot form concurrently in the CsCu2I3 system by the doping method.

 figure: Fig. 3.

Fig. 3. Crystal structures of (a) Cs3Cu2I5 and (b) CsCu2I3 with optimized lattice parameters from the top.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. (a) PL spectra under excitation wavelength of 320 nm. (b) XRD patterns of the samples with NdI3 added into the CsCu2I3 system with various ratios.

Download Full Size | PDF

Next, other materials like TbI3, PrI3, and BiI3 were added in turn to measure their respective effect on the white emission efficiency. As shown in Fig. 5(a), PL white spectra involving Cs3Cu2I5 and CsCu2I3 peaks have been detected in every material. The PLQYs for NdI3, TbI3, PrI3, and BiI3 are 16.9%, 14.5%, 9.32%, and 5.52%, respectively. The CIE coordinates in the system of NdI3 and TbI3 are (0.3397, 0.3325) and (0.3477, 0.3368), respectively [Fig. 5(c)], indicating that white emission has been successfully formed within them. As shown in Fig. 5(b), the corresponding XRD peaks of Cs3Cu2I5 and CsCu2I3 appear simultaneously but with different proportions in these samples. Compared to other impurities, the NdI3 sample has the strongest CsCu2I3 diffraction peak, indicating the largest amount of CsCu2I3 in it. In the reaction process, the solubility and the ionic activity of impurities in the Cs3Cu2I5 precursor solution will affect the amount and quality of CsCu2I3 to a large extent and will further influence the overall luminescence effect. The inset of Fig. 5(a) shows the NdI3, TbI3, and PrI3 doped Cs3Cu2I5 solution. It can be seen that the NdI3 sample is completely dissolved and more clarified than others, while the TbI3 and PrI3 samples are like a suspension similar to colloid, which means that these two impurities do not dissolve absolutely in the Cs3Cu2I5 solution. In addition, Nd is one of the liveliest rare earth metals; it will quickly react in hot solution which is effective to the fast formation of CsCu2I3 and makes positive influence on the overall emission efficiency. Therefore, compared to other dopants, the NdI3 sample has the strongest CsCu2I3 diffraction peak, indicating the largest amount of CsCu2I3 in it. As discussed previously in Fig. 2, the more amount of CsCu2I3 is beneficial to obtain stronger white emission, and thus the PLQY of the NdI3 sample is higher than that of other materials. Moreover, the uniformity and quality of the sample could also affect the whole luminescence intensity. The SEM images are shown in Figs. 5(d)–5(g); the uniformity and quality of Cs3Cu2I5 and CsCu2I3 in the NdI3 sample are much better compared to others, corresponding to its higher PL efficiency. In addition, the new generated material is another factor affecting the PLQY. From Fig. 5(b), it can be found that in the BiI3-doped sample new Cs3Bi2I9 appears. It consumes a great deal of Cs and I ions which will compose Cs3Cu2I5 and CsCu2I3, resulting in the whole material getting few effective luminous centers for white emission and a low PL intensity. Therefore, combined with the above factors, the PLQY of the NdI3 sample is higher than that of other doping materials.

 figure: Fig. 5.

Fig. 5. (a) PL spectra of samples with 9 mol% concentration of NdI3 (PL excitation of 316 nm), TbI3 (PL excitation of 317 nm), PrI3 (PL excitation of 319 nm), and BiI3 (PL excitation of 308 nm). Inset: the precursor solution of NdI3, TbI3, and PrI3 samples. (b) XRD patterns of samples with 9 mol% concentration of NdI3, TbI3, PrI3, and BiI3. (c) CIE coordinates of samples with 9 mol% NdI3 and TbI3. (d)–(g) SEM images of the samples with 9 mol% NdI3, TbI3, PrI3, and BiI3, respectively.

Download Full Size | PDF

4. CONCLUSIONS

In conclusion, a controllable one-step doping method was adopted in the cesium copper iodide perovskite’s luminescence, and the results indicated that a system including Cs3Cu2I5 and CsCu2I3 with high quality had been successfully prepared. Through comparing the PL efficiency of samples under various molar ratios and materials, it has been investigated that the amount of CsCu2I3 combining with the uniformity and quality of Cs3Cu2I5 and CsCu2I3 was the key factor affecting the white-lighting. Therefore, a high-quality white-emission with CIE coordinates of (0.3397, 0.3325) and CRI of 90 was obtained in a convenient way. This work provides a new approach for the investigation of Cu-based metal halide perovskites and will be helpful for the exploration of lead-free perovskites.

Funding

Natural Science Foundation of Guangdong Province (2020A1515010541); Science and Technology Project of Shenzhen (JCYJ20180305124930169, JCYJ20190808143419622, ZDSYS201707271014468).

Disclosures

The authors declare no conflicts of interest.

REFERENCES

1. A. Dutta, R. K. Behera, P. Pal, S. Baitalik, and N. Pradhan, “Near-unity photoluminescence quantum efficiency for all CsPbX3 (X=Cl, Br and I) perovskite nanocrystals: a generic synthesis approach,” Angew. Chem. 131, 5608–5612 (2019). [CrossRef]  

2. R. J. Sutton, G. E. Eperon, L. Miranda, E. S. Parrott, B. A. Kamino, J. B. Patel, M. T. Hörantner, M. B. Johnston, A. A. Haghighirad, D. T. Moore, and H. J. Snaith, “Bandgap-tunable cesium lead halide perovskites with high thermal stability for efficient solar cells,” Adv. Energy Mater. 6, 1502458 (2016). [CrossRef]  

3. Z.-K. Tan, R. S. Moghaddam, M. L. Lai, P. Docampo, R. Higler, F. Deschler, M. Price, A. Sadhanala, L. M. Pazos, D. Credgington, F. Hanusch, T. Bein, H. J. Snaith, and R. H. Friend, “Bright light-emitting diodes based on organometal halide perovskite,” Nat. Nanotechnol. 9, 687–692 (2014). [CrossRef]  

4. 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, “Highly efficient broadband yellow phosphor based on zero-dimensional tin mixed-halide perovskite,” ACS Appl. Mater. Interfaces 9, 44579–44583 (2017). [CrossRef]  

5. E. P. Yao, Z. Yang, L. Meng, P. Sun, S. Dong, Y. Yang, and Y. Yang, “High-brightness blue and white LEDs based on inorganic perovskite nanocrystals and their composites,” Adv. Mater. 29, 1606859 (2017). [CrossRef]  

6. 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, “One-dimensional lead halide perovskites with bluish white-light emission,” Nat. Commun. 8, 14051 (2017). [CrossRef]  

7. C.-Y. Chang, A. N. Solodukhin, S.-Y. Liao, K. P. O. Mahesh, C.-L. Hsu, S. A. Ponomarenko, Y. N. Luponosov, and Y.-C. Chao, “Perovskite white light-emitting diodes based on a molecular blend perovskite emissive layer,” J. Mater. Chem. C 7, 8634–8642 (2019). [CrossRef]  

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

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

10. P. Vashishtha and J. E. Halpert, “Field-driven ion migration and color instability in red-emitting mixed halide perovskite nanocrystal light-emitting diodes,” Chem. Mater. 29, 5965–5973 (2017). [CrossRef]  

11. M. C. Brennan, S. Draguta, P. V. Kamat, and M. Kuno, “Light-induced anion phase segregation in mixed halide perovskites,” ACS Energy Lett. 3, 204–213 (2018). [CrossRef]  

12. G. Nedelcu, L. Protesescu, S. Yakunin, M. I. Bodnarchuk, M. J. Grotevent, and M. V. Kovalenko, “Fast anion-exchange in highly luminescent nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, I),” Nano lett. 15, 5635–5640 (2015). [CrossRef]  

13. J. A. Sichert, Y. Tong, N. Mutz, M. Vollmer, S. Fischer, K. Z. Milowska, R. Garcia Cortadella, B. Nickel, C. Cardenas-Daw, J. K. Stolarczyk, A. S. Urban, and J. Feldmann, “Quantum size effect in organometal halide perovskite nanoplatelets,” Nano Lett. 15, 6521–6527 (2015). [CrossRef]  

14. A. Babayigit, D. D. Thanh, A. Ethirajan, J. Manca, M. Muller, H.-G. Boyen, and B. Conings, “Assessing the toxicity of Pb-and Sn-based perovskite solar cells in model organism Danio rerio,” Sci. Rep. 6, 18721 (2016). [CrossRef]  

15. H. Yang, Y. Zhang, J. Pan, J. Yin, O. M. Bakr, and O. F. Mohammed, “Room-temperature engineering of all-inorganic perovskite nanocrystals with different dimensionalities,” Chem. Mater. 29, 8978–8982 (2017). [CrossRef]  

16. J. Li, H. Dong, B. Xu, S. Zhang, Z. Cai, J. Wang, and L. Zhang, “CsPbBr3 perovskite quantum dots: saturable absorption properties and passively Q-switched visible lasers,” Photon. Res. 5, 457–460 (2017). [CrossRef]  

17. P. Cheng, L. Sun, L. Feng, S. Yang, Y. Yang, D. Zheng, Y. Zhao, Y. Sang, R. Zhang, D. Wei, W. Deng, and K. Han, “Colloidal synthesis and optical properties of all-inorganic low-dimensional cesium copper halide nanocrystals,” Angew. Chem. 58, 16087–16091 (2019). [CrossRef]  

18. K. Sim, T. Jun, J. Bang, H. Kamioka, J. Kim, H. Hiramatsu, and H. Hosono, “Performance boosting strategy for perovskite light-emitting diodes,” Appl. Phys. Rev. 6, 031402 (2019). [CrossRef]  

19. M. I. Saidaminov, J. Almutlaq, S. Sarmah, I. Dursun, A. A. Zhumekenov, R. Begum, J. Pan, N. Cho, F. Mohammed, and O. M. Bakr, “Pure Cs4PbBr6: highly luminescent zero-dimensional perovskite solids,” ACS Energy Lett. 1, 840–845 (2016). [CrossRef]  

20. X.-X. Feng, X.-D. Lv, Q. Liang, J. Cao, and Y. Tang, “Diammonium porphyrin-induced CsPbBr3 nanocrystals to stabilize perovskite films for efficient and stable solar cells,” ACS Appl. Mater. Interfaces 12, 16236–16242 (2020). [CrossRef]  

21. E. P. Booker, J. T. Griffiths, L. Eyre, C. Ducati, N. C. Greenham, and N. J. L. K. Davis, “Synthesis, characterization, and morphological control of Cs2CuCl4 nanocrystals,” J. Phys. Chem. C 123, 16951–16956 (2019). [CrossRef]  

22. P. Yang, G. Liu, B. Liu, X. Liu, Y. Lou, J. Chen, and Y. Zhao, “All-inorganic Cs2CuX4 (X = Cl, Br, and Br/I) perovskite quantum dots with blue-green luminescence,” Chem. Commun. 54, 11638–11641 (2018). [CrossRef]  

23. R. Roccanova, A. Yangui, H. Nhalil, H. Shi, M.-H. Du, and B. Saparov, “Near-unity photoluminescence quantum yield in blue-emitting Cs3Cu2Br5−xIx (0 ≤ x ≤ 5),” ACS Appl. Electron. Mater. 1, 269–274 (2019). [CrossRef]  

24. R. Roccanova, A. Yangui, G. Seo, T. D. Creason, Y. Wu, D. Y. Kim, M.-H. Du, and B. Saparov, “Bright luminescence from nontoxic CsCu2X3 (X= Cl, Br, I),” ACS Mater. Lett. 1, 459–465 (2019). [CrossRef]  

25. L. Xie, B. Chen, F. Zhang, Z. Zhao, X. Wang, L. Shi, Y. Liu, L. Huang, R. Liu, B. Zou, and Y. Wang, “Highly luminescent and stable lead-free cesium copper halide perovskite powders for UV-pumped phosphor-converted light-emitting diodes,” Photon. Res. 8, 768–775 (2020). [CrossRef]  

26. Y. Li, P. Vashishtha, Z. Zhou, Z. Li, S. B. Shivarudraiah, C. Ma, J. Liu, K. S. Wong, H. Su, and J. E. Halpert, “Room temperature synthesis of stable, printable Cs3Cu2X5 (X = I, Br/I, Br, Br/Cl, Cl) colloidal nanocrystals with near-unity quantum yield green emitters (X = Cl),” Chem. Mater. 32, 5515–5524 (2020). [CrossRef]  

27. Z. Ma, Z. Shi, C. Qin, M. Cui, D. Yang, X. Wang, L. Wang, X. Ji, X. Chen, J. Sun, D. Wu, Y. Zhang, X. J. Li, L. Zhang, and C. Shan, “Stable yellow light-emitting devices based on ternary copper halides with broadband emissive self-trapped excitons,” ACS Nano 14, 4475–4486 (2020). [CrossRef]  

28. R. Lin, Q. Guo, Q. Zhu, Y. Zhu, W. Zheng, and F. Huang, “All-inorganic CsCu2I3 single crystal with high-PLQY (≈15.7%) intrinsic white-light emission via strongly localized 1D excitonic recombination,” Adv. Mater. 31, 1905079 (2019). [CrossRef]  

29. P. Vashishtha, G. V. Nutan, B. E. Griffith, Y. Fang, D. Giovanni, M. Jagadeeswararao, T. C. Sum, N. Mathews, S. G. Mhaisalkar, J. V. Hanna, and T. Wu, “Cesium copper iodide tailored nanoplates and nanorods for blue, yellow, and white emission,” Chem. Mater. 31, 9003–9011 (2019). [CrossRef]  

30. S. Fang, Y. Wang, H. Li, F. Fang, K. Jiang, Z. Liu, H. Li, and Y. Shi, “Rapid synthesis and mechanochemical reactions of cesium copper halides for convenient chromaticity tuning and efficient white light emission,” J. Mater. Chem. C 8, 4895–4901 (2020). [CrossRef]  

31. S. Liu, Y. Yue, X. Zhang, C. Wang, G. Yang, and D. Zhu, “A controllable and reversible phase transformation between all-inorganic perovskites for white light emitting diodes,” J. Mater. Chem. C 8, 8374–8379 (2020). [CrossRef]  

32. Z. Ma, Z. Shi, D. Yang, Y. Li, F. Zhang, L. Wang, X. Chen, D. Wu, Y. Tian, Y. Zhang, L. Zhang, X. Li, and C. Shan, “High color-rendering index and stable white light-emitting diodes by assembling two broadband emissive self-trapped excitons,” Adv. Mater. 33, 2001367 (2021). [CrossRef]  

33. 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, “Efficient and stable emission of warm-white light from lead-free halide double perovskites,” Nature 563, 541–545 (2018). [CrossRef]  

34. S. Hull and P. Berastegui, “Crystal structures and ionic conductivities of ternary derivatives of the silver and copper monohalides—II: ordered phases within the (AgX)x–(MX)1-x and (CuX)x–(MX)1-x (M = K, Rb and Cs; X=Cl, Br and I) systems,” J. Solid State Chem. 177, 3156–3173 (2004). [CrossRef]  

35. T. Jun, K. Sim, S. Iimura, M. Sasase, H. Kamioka, J. Kim, and H. Hosono, “Lead-free highly efficient blue-emitting Cs3Cu2I5 with 0D electronic structure,” Adv. Mater. 30, 1804547 (2018). [CrossRef]  

36. Y. Su, Q. Zeng, X. Chen, W. Ye, L. She, X. Gao, Z. Rena, and X. Li, “Highly efficient CsPbBr3 perovskite nanocrystals induced by structure transformation between CsPbBr3 and Cs4PbBr6 phases,” J. Mater. Chem. C 7, 7548–7553 (2019). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) SEM, (b) EDS mapping, and (c) TEM images of the sample with NdI3. The lattice planes and fast Fourier transform (FFT) images of (d) Cs3Cu2I5 nanoparticles and (e) CsCu2I3 nanorods. (f) XPS analysis of the sample with 9 mol% NdI3, and the respective spectra of (g) Cs 3d, (h) Cu 2p, and (i) I 3d.
Fig. 2.
Fig. 2. (a) PL spectra of pure Cs3Cu2I5 and CsCu2I3. (b) PL spectra of samples with different concentration of NdI3 under excitation of 316 nm. (c) PLE spectra of emission peaks corresponding to Cs3Cu2I5 and CsCu2I3 of the sample with 9 mol% NdI3. (d) CIE coordinates of the sample with 9 mol% NdI3. (e) XRD diffraction patterns of samples with different concentration of NdI3 compared to the standard XRD patterns of Cs3Cu2I5 and CsCu2I3.
Fig. 3.
Fig. 3. Crystal structures of (a) Cs3Cu2I5 and (b) CsCu2I3 with optimized lattice parameters from the top.
Fig. 4.
Fig. 4. (a) PL spectra under excitation wavelength of 320 nm. (b) XRD patterns of the samples with NdI3 added into the CsCu2I3 system with various ratios.
Fig. 5.
Fig. 5. (a) PL spectra of samples with 9 mol% concentration of NdI3 (PL excitation of 316 nm), TbI3 (PL excitation of 317 nm), PrI3 (PL excitation of 319 nm), and BiI3 (PL excitation of 308 nm). Inset: the precursor solution of NdI3, TbI3, and PrI3 samples. (b) XRD patterns of samples with 9 mol% concentration of NdI3, TbI3, PrI3, and BiI3. (c) CIE coordinates of samples with 9 mol% NdI3 and TbI3. (d)–(g) SEM images of the samples with 9 mol% NdI3, TbI3, PrI3, and BiI3, respectively.
Select as filters


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