Next Article in Journal
Thermal Conductivity and Microstructure of Novel Flaxseed-Gum-Filled Epoxy Resin Biocomposite: Analytical Models and X-ray Computed Tomography
Previous Article in Journal
Temperature and E-Poling Evolution of Structural, Vibrational, Dielectric, and Ferroelectric Properties of Ba1−xSrxTiO3 Ceramics (x = 0, 0.1, 0.2, 0.3, 0.4 and 0.45)
Previous Article in Special Issue
Progress in the Synthesis and Application of Transparent Conducting Film of AZO (ZnO:Al)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lead-Free Halide Perovskite Nanocrystals for Light-Emitting Diodes

1
Department of Materials Science and Engineering, Soongsil University, 369 Sangdo-ro, Dongjak-gu, Seoul 06978, Republic of Korea
2
Department of Green Chemistry and Materials Engineering, Soongsil University, 369 Sangdo-ro, Dongjak-gu, Seoul 06978, Republic of Korea
3
Integrative Institute of Basic Science, Soongsil University, 369 Sangdo-ro, Dongjak-gu, Seoul 06978, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2023, 16(18), 6317; https://doi.org/10.3390/ma16186317
Submission received: 13 August 2023 / Revised: 11 September 2023 / Accepted: 18 September 2023 / Published: 20 September 2023

Abstract

:
Lead-based halide perovskite nanocrystals (PeNCs) have demonstrated remarkable potential for use in light-emitting diodes (LEDs). This is because of their high photoluminescence quantum yield, defect tolerance, tunable emission wavelength, color purity, and high device efficiency. However, the environmental toxicity of Pb has impeded their commercial viability owing to the restriction of hazardous substances directive. Therefore, Pb-free PeNCs have emerged as a promising solution for the development of eco-friendly LEDs. This review article presents a detailed analysis of the various compositions of Pb-free PeNCs, including tin-, bismuth-, antimony-, and copper-based perovskites and double perovskites, focusing on their stability, optoelectronic properties, and device performance in LEDs. Furthermore, we address the challenges encountered in using Pb-free PeNC-LEDs and discuss the prospects and potential of these Pb-free PeNCs as sustainable alternatives to lead-based PeLEDs. In this review, we aim to shed light on the current state of Pb-free PeNC LEDs and highlight their significance in driving the development of eco-friendly LED technologies.

1. Introduction

In recent years, metal halide perovskites (MHPs) have received considerable attention as highly promising materials for a range of optoelectronic applications, including solar cells, light-emitting diodes (LEDs), photodetectors, lasers, and photocatalysts [1,2,3,4,5,6,7,8,9,10,11,12]. Notably, lead (Pb)-based halide perovskites (LHPs) have emerged as particularly appealing. This is due to their exceptional optoelectronic properties, which are characterized by high photoluminescence quantum yield (PLQY), long carrier diffusion length, high charge carrier mobility, high defect tolerance, and narrow emission spectra [13,14,15,16,17,18,19,20]. These properties make these materials highly promising candidates for next-generation LEDs. Moreover, a reduction in the crystal size toward the nanometer scale has a substantial impact on the material properties, thereby enabling deliberate adjustment of the optoelectronic characteristics [21,22,23,24,25]. Applications of MHP nanocrystals (PeNCs) have already been established in various fields [26,27,28,29,30,31,32], and their device applications are expected to be commercialized.
However, the presence of Pb2+ in LHPs poses significant challenges to their widespread commercialization. Pb2+ is highly soluble in water, which introduces potential risks to the environment and human health throughout the product life cycle, including disposal. Exposure to Pb can result in various adverse health effects, including neurological disorders, gastrointestinal issues, insomnia, coma, and convulsions. The inherent toxicity of Pb raises concerns regarding its environmental impact and human safety, making LHPs incompatible with the restriction of hazardous substances (RoHS) directive and industrial reliability standards [33,34,35]. Therefore, there has been growing interest, and various research efforts have focused on exploring and developing Pb-free PeNCs as promising alternatives to address these intrinsic challenges.
The development of Pb-free perovskites is a crucial alternative strategy for the commercialization of MHP-based devices by substituting Pb cations with nontoxic metal cations, such as tin (Sn), bismuth (Bi), antimony (Sb), or transition metals such as copper (Cu). To achieve this goal, researchers have aimed to eliminate toxicity concerns associated with Pb while achieving comparable or enhanced optoelectronic performance [36,37,38,39,40].
In this review, we provide a comprehensive overview of recent advances in Pb-free PeNCs for LED (PeNC-LED) applications. We discuss the synthetic strategies employed for the development of Pb-free PeNCs, including the selection of nontoxic metal cations and the optimization of crystal structures. Additionally, we examine the optoelectronic properties of Pb-free PeNCs, such as their photoluminescence (PL) characteristics, band structure, and emission mechanism. Moreover, we explore the strategies employed to enhance the stability and operational lifetime of Pb-free PeNC-LEDs and discuss the challenges by addressing key research directions and prospects for the development of Pb-free PeNC-LEDs.

2. B-Site Metal Candidates for Pb-Free Perovskite Nanocrystals

In the context of traditional LHPs, the general chemical formula can be expressed as ABX3, wherein A represents an inorganic monocation such as Cs+, Rb+, or K+ and an organic monocation such as the methylammonium cation (MA+) and the formamidinium cation; B denotes a bivalent Pb2+ ion; and X represents halide ions such as Cl, Br, or I [41,42]. The stability of three-dimensional (3D) ABX3 perovskite structures can be evaluated using Goldschmidt’s tolerance factor t (1) and the octahedral factor μ (2):
t = r A + r X 2 ( r B + r X )
μ = r B r X
where rA, rB, and rX correspond to the ionic radii A, B, and X, respectively. Stable perovskite crystal structures are formed when these two factors are in the ranges of 0.8 ≤ t ≤ 1.0 and 0.44 ≤ μ ≤ 0.90 [41,43,44,45]. However, as the ionic radius of the A-site cation increases, which leads to a larger tolerance factor, lower-dimensional perovskite structures are formed [46,47,48]. Such dimensions and crystal structure variations result in MHPs having diverse band structures and photophysical and optoelectronic properties, making them promising candidates for various applications. Figure 1 shows the elements in the periodic table denoted by distinct colors (A-site: green, B-site: sky-blue, and X-site: orange), considered potentially viable or having been employed at the A, B, and X positions of the MHPs [49,50,51,52,53,54]. Extensive research is being conducted on MHPs with various compositions to progressively enhance their performance across diverse fields.
Furthermore, MHPs based on toxic metals, such as Pb and Cd, encounter challenges for commercialization owing to the RoHS directive. Thus, research on Pb-free MHPs is essential for their industrial applications. To substitute non-toxic metals for toxic Pb in LHPs, attention should be paid to the ionic radius and charge neutrality of the substituting metal elements. The selection of elements must satisfy the requirements of an appropriate ionic radius to maintain the structural integrity of the perovskite lattice. Moreover, substitutions should ensure charge neutrality within the crystal structure to maintain overall electrostatic balance [55]. By carefully considering these factors, it is possible to identify suitable elements for effective Pb substitution in LHPs. This leads to the development of Pb-free MHPs with the desired properties for optoelectronic device applications, including LEDs. Among the promising elemental candidates, we focus on Sn-, Bi-, Sb-, Cu-based, and double PeNCs and provide an overview of their applications in PeNC-LEDs.

2.1. Sn-Based Perovskite Nanocrystals

Sn belongs to group 14 of the periodic table, which is the same group as Pb, and possesses an equivalent valence configuration, resulting in the ability to retain electrical neutrality within the ABX3 crystal structure (Figure 2a). Moreover, Sn cations have an ideal cubic structure surrounded by an octahedron of anions. These characteristics have led to extensive investigations into Sn as a promising B-site metal replacement for Pb. However, ABX3-structured Sn-based perovskites have been observed to spontaneously undergo rapid decolorization and conversion to the A2BX6 formula after exposure to atmospheric conditions, implying a rapid degradation phenomenon [56]. Such structural changes adversely affect the chemical and luminescence properties of Sn-based PeNCs (Sn-PeNCs) [57]. Therefore, significant efforts have been made in recent years to investigate the structural and luminescent properties of Sn-PeNCs.
Precise control of the precursor ratios over various ranges is crucial for achieving a low density of structural defects within the lattice of the CsSnX3 NCs [58]. This is because the defect formation energy is intricately linked to the chemical potential of each constituent involved in the reaction. Figure 2b shows bright-field transmission electron microscopy (TEM) images of CsSnI3 NCs with different Cs:Sn ratios. As the Sn content increases, the average size of the cubic CsSnI3 NCs also increases. For instance, an average size of 26 nm was obtained at a Cs:Sn ratio of 0.25:3 and 37 nm at a ratio of 0.25:4.8. Additionally, byproducts, which are presumed to be unreacted species or reaction intermediates, were observed in CsSnX3 NCs with a ratio of 0.25:4.8. Furthermore, at the higher Sn ratio of 4.8, large crystals exceeding 200 nm were formed, larger than those at ratios of 3, 3.6, and 4.2. These results indicate that the precursor ratio plays a crucial role in determining the size distribution and purity of CsSnI3 NCs during synthesis.
By consistently addressing the challenge of preventing the oxidation of Sn(II) to Sn(IV), the performance of Sn-based MHPs has demonstrated consistent improvement. The low stability of CsSnBr3 NCs can be enhanced by the precise synthesis of CsSnBr3 cubic nanocages, which uses a simple hot-injection colloidal approach [59]. The fabrication of the nanocages can be carefully regulated by selecting suitable precursors, such as MgBr, and adjusting the reaction temperature, resulting in their formation through a self-assembly-driven process. MgBr was used as the bromide source because Mg2+ has a smaller ionic radius, which, according to tolerance factor theory, prevents its penetration into the perovskite structure. Uncoordinated 1-octadecene (ODE) was chosen as the solvent, and oleic acid (OA) and oleylamine (OAm) were selected as ligands. During this process, injection of the Cs-oleate precursor into the Sn-Br precursor solution resulted in a rapid color change, turning the solution dark red within a few seconds. This color transformation indicates the successful formation of CsSnBr3 nanocages.
A surface treatment with perfluorooctanoic acid (PFOA) also substantially enhanced the stability of CsSnBr3 cubic nanocages [59]. PFOA exhibits a stronger electron-withdrawing capability than Br, which passivates the surface of the CsSnBr3 nanocages. It forms a stronger interaction with Sn2+, effectively hindering oxidation. Additionally, PFOA induces steric hindrance, further enhancing the stability of the CsSnBr3 nanocages. As shown in Figure 2c, the pristine CsSnBr3 nanocage films show gradual disintegration, with approximately 30% decomposition after 48 h under various desiccation conditions, whereas the PFOA-functionalized CsSnBr3 nanocage film exhibits a relatively slow decomposition rate of 25% in a moisture-free environment. When the films were exposed to illumination, the pristine CsSnBr3 films degraded rapidly, with over 95% decomposition within 24 h. Conversely, the PFOA-functionalized CsSnBr3 film exhibited a slight degradation of approximately 10% under the same conditions. Therefore, water is a critical factor that directly influences the stability of CsSnBr3 nanocages, and exposure to light accelerates degradation by the decomposition process.
Surface passivation using polymers is another effective strategy for enhancing the stability of Sn-PeNCs [60,61,62]. The incorporation of gelatin, an amphoteric polymer obtained as a partial product of collagen hydrolysis, is crucial for providing resistance against water degradation [61]. The hygroscopic nature of the gelatin network effectively stabilizes and protects CsSnCl3 NCs from water-induced damage. Moreover, the high molecular weight of gelatin allows tight coating and isolation of the PeNCs, effectively suppressing self-aggregation and oxidation.
Figure 2. (a) Crystal structure of CsSnI3; (b) TEM images of CsSnI3 PeNCs with different ratios of Cs:Sn. Insets indicate histograms of the average particle size of each CsSnI3 NC; (b) Reprinted with permission from Ref. [58]. Copyright, 2021 American Chemical Society. (c) The variation of absorbance over time intensity at 620 nm under various conditions. The temperature and light intensity were 25 °C and 1200 mW/cm2 for all measurements; (c) Reprinted with permission from Ref. [59]. Copyright, 2017 American Chemical Society. (d) Remnant PL values of gelatin-NCs and oleic acid (OA)/oleylamine (OAm)-NCs over time under exposure to air, UV irradiation, and water. (Inset: remnant PL of gelatin-NCs and OA/OAm-NCs in 0.1 h). (e) Schematic image of gelatin coordinate with Sn in CsSnI3 NCs; (d,e) Reprinted with permission from Ref. [61]. Copyright 2020, Elsevier.
Figure 2. (a) Crystal structure of CsSnI3; (b) TEM images of CsSnI3 PeNCs with different ratios of Cs:Sn. Insets indicate histograms of the average particle size of each CsSnI3 NC; (b) Reprinted with permission from Ref. [58]. Copyright, 2021 American Chemical Society. (c) The variation of absorbance over time intensity at 620 nm under various conditions. The temperature and light intensity were 25 °C and 1200 mW/cm2 for all measurements; (c) Reprinted with permission from Ref. [59]. Copyright, 2017 American Chemical Society. (d) Remnant PL values of gelatin-NCs and oleic acid (OA)/oleylamine (OAm)-NCs over time under exposure to air, UV irradiation, and water. (Inset: remnant PL of gelatin-NCs and OA/OAm-NCs in 0.1 h). (e) Schematic image of gelatin coordinate with Sn in CsSnI3 NCs; (d,e) Reprinted with permission from Ref. [61]. Copyright 2020, Elsevier.
Materials 16 06317 g002
Accordingly, gelatin-coated CsSnI3 nanocrystals (gel-CsSnI3 NCs) have been shown to exhibit notably enhanced stability under harsh environmental conditions when compared to pristine OA/OAm-based CsSnI3 NCs (Figure 2d). The PL intensity of the gel-CsSnI3 NCs was maintained above 93.28% and 85% for 72 h under dark conditions and UV irradiation, respectively (pink and green lines). By contrast, the water resistance test revealed that the PL stability of pristine CsSnI3 NCs degraded by over 95% within 5 min. The low stability of pristine CsSnI3 NCs could originate from the addition of water to the PeNC solution in cyclohexane, resulting in delamination, ligand loss, self-aggregation, PL quenching, and even decomposition. On the contrary, the PL intensity of gel-CsSnI3 NCs exhibited minimal change after 5 min (Figure 2d, inset) and remained above 75% after 72 h (blue line).
Based on these results, Figure 2e presents a schematic of the proposed interaction mechanism between the gelatin and CsSnCl3 NCs. The interaction between gelatin and the CsSnCl3 NCs involves the coordination of the carboxylate and amino groups of gelatin with the Sn atoms in the NCs. Additionally, the protonated amino groups form hydrogen bonds with the Cl atoms. The numerous functional groups in gelatin lead to the occupation of active sites on the NC surface, creating a “rich ligand” state. As a result, comprehensive passivation of the NC surface by gelatin effectively inhibits the formation of surface defects. Therefore, the surface passivation strategy enhances the stability of Sn-PeNCs under atmospheric conditions, highlighting their potential for various applications.
Sn-PeNCs have been extensively studied as environmentally friendly Pb-free PeNCs due to their promising characteristics, including the same ns2 valence electron configuration and similar ion radii (Sn:1.35 Å, Pb:1.49 Å). They enable the substitution of Pb cations within the ABX3 structure while maintaining charge neutrality [63]. To enhance their structural stability, numerous studies have been conducted on controlling precursor ratios and surface engineering through the introduction of additives such as perfluoroalkyl acids and polymers. However, their practical applications have been limited owing to a lack of research on suitable additives for preventing Sn oxidation. Consequently, current research efforts have focused on enhancing the stability of Sn-PeNCs by exploring additives that can effectively inhibit Sn oxidation.

2.2. Bi-Based Perovskite Nanocrystals

The electronic properties of the LHPs can be attributed to the electronic configuration of Pb2+, which is characterized as 6s26p0. This configuration enables the hybridization of the Pb 6s orbital with the halogen p orbital for valence band formation and of the Pb 6p orbitals with halogen p orbitals for conduction bands. This leads to the observed electronic properties. The presence of shallow defect states in LHPs is closely linked to the strong antibonding interactions that occur between the Pb 6s and halide 5p orbitals within the valence band [64,65].
Bi, which belongs to the same period as Pb, possesses a cationic electron configuration similar to that of Pb2+, featuring a [Xe] 4f145d106s2 electron arrangement. The presence of a 6s2 lone pair is a distinctive characteristic of Bi that enables it to displace Pb by a heterogeneous substitution mechanism, resulting in the replacement of Pb2+ with Bi3+ [66]. Moreover, Bi is an eco-friendly alternative to Pb, offering low toxicity and high stability [67,68,69,70].
When Bi replaces Pb in LHPs, maintaining electrical neutrality requires the substitution of three Pb2+ ions with two Bi3+ ions, leading to the formation of A3B2X9 structures within the ABX3 framework. Leng et al. first reported the synthesis of MA3Bi2Br9 NCs that exhibited an emission wavelength of 430 nm and a maximum PLQY of 12% [71]. Subsequently, numerous studies have been conducted on the synthesis of Bi-PeNCs [69,70,71,72,73,74]. However, the inconsistent optical properties observed for various colloidal Bi-PeNC synthesis methods indicate the presence of various crystal structures. The combination of Cs+ and Bi3+ cations with halide anions can generate Cs3BiX6 and Cs3Bi2X9 (Figure 3a) [75]. Cs3Bi2Br9 possesses a two-dimensional (2D) layered structure in the P 3 ¯ m1 space group, and one-third of its B-site cation positions are replaced by vacancies. By contrast, dimer structures such as Cs3Bi2I9 are formed by Cs+ or MA+ cations in the face-sharing octahedra and belong to the P63/mmc space group. The bandgap calculated by density functional theory (DFT) using the Vienna ab initio simulation package indicates that the layered structure has a smaller bandgap than the dimer structure, which can be attributed to its smaller lattice constant and volume [76,77]. By changing the A-site in Cs3Bi2I9 from Cs+ to Rb+ or K+, the transformation from a dimer to a layered structure can be observed, leading to a transition from an indirect bandgap to a direct bandgap. However, when Rb+ or K+ cations are used at the A-site of the A3Bi2I9 structure, the smaller size of these ions results in a lower-symmetry distorted lattice with the P21/n space group [78].
To investigate the bandgap characteristics associated with the Cs3Bi2I9 crystal structure, colloidal Cs3BiX6 and Cs3Bi2X9 NCs were synthesized using the hot-injection method and anion exchange with trimethylsilyl halides (TMS-X) [75]. The direct synthesis of Cs3Bi2I9 leads to the formation of a dimer rather than a layered structure [79,80,81]. Because the direct synthesis of layered Cs3Bi2I9 is challenging, layered Cs3Bi2I9 can be achieved through anion exchange with TMS-I. When comparing the optical properties of dimeric Cs3Bi2I9 with those of layered Cs3Bi2I9, it was observed that the first absorption peak in dimeric Cs3Bi2I9 appeared at 492 nm, while a similar transition in layered Cs3Bi2I9 appeared to be red-shifted to 550 nm (Figure 3b). Layered Cs3Bi2I9 exhibits distinctive peaks that differ from dimeric Cs3Bi2I9, although the powder X-ray diffraction (XRD) patterns do not entirely exclude the presence of dimeric structures in the layered phase (Figure 3c). Additionally, band structure calculations using plane-wave DFT revealed that the band gap of the layered structure is lower than that of the dimeric structure. This observation is attributed to the lower conduction band minimum (CBM) in the layered structure compared to that in the dimer structure, resulting in a reduction in the bandgap. The crystal structure of Bi-PeNCs can be influenced by the heating-up process, which is commonly employed for the synthesis of colloidal inorganic quantum dots (QDs). This is because it induces the nucleation, growth, and crystallization of inorganic QDs at high temperatures. The heating-up synthesis method provides relative safety owing to the absence of injection at high temperatures and the controllability of the crystal size by varying the reaction temperature and time, making it potentially advantageous for large-scale production [82,83,84,85,86]. Building on this approach, heating-up synthesis without injection can be used to synthesize cesium bismuth bromide NCs [87]. A mixture of CsBr, BiBr3, OA, and OAm was stirred in ODE under vacuum, followed by heat treatment, resulting in the synthesis of Cs3BiBr6 or Cs3Bi2Br9 NCs. The structure of cesium bismuth bromide NCs can transform from Cs3BiBr6 to Cs3Bi2Br9, depending on the temperature and amount of OA. The PL and PL excitation (PLE) spectra of Cs3BiBr6 and Cs3Bi2Br9 are shown in Figure 3d and 3e, respectively. The emission peak of Cs3BiBr6 is observed at 435 nm, whereas that of Cs3Bi2Br9 is red-shifted to 461 nm. The absorption spectra corresponded to a bandgap of 3.05 and 2.67 eV, respectively.
Figure 3. (a) Crystal structures of Cs3BiCl6, Cs3Bi2Br9, and Cs3Bi2I9 in bulk; (b) extinction coefficient (solid lines) and PL (dashed lines) spectra for Cs3Bi2I9 polymorphs (blue: layered, green: dimer); (c) Powder X-ray diffraction (XRD) patterns for Cs3Bi2I9 polymorphs; (ac) reprinted with permission from Ref. [75]. Copyright 2019, American Chemical Society. (d) PL excitation (PLE) and PL spectrum of Cs3BiBr6 NCs. PL at an excitation wavelength of 360 nm and PLE at an emission wavelength of 440 nm; (e) PL and PLE spectra of Cs3Bi2Br9 NCs. PL under an excitation wavelength of 358 nm and PLE at an emission wavelength of 461 nm; (d,e) reprinted with permission from Ref. [87]. Copyright 2020, Royal Society of Chemistry. (f) Absorption (solid lines) and PL (dashed lines) spectra of different levels of Mn-doped Cs3Bi2I9 NCs, respectively. Inset: the appearance of different Mn-doped Cs3Bi2I9 NCs; (g) emission mechanisms of Cs3Bi2I9 with Mn2+ doping; (f,g) reprinted with permission from Ref. [88]. Copyright 2021, Wiley-VCH.
Figure 3. (a) Crystal structures of Cs3BiCl6, Cs3Bi2Br9, and Cs3Bi2I9 in bulk; (b) extinction coefficient (solid lines) and PL (dashed lines) spectra for Cs3Bi2I9 polymorphs (blue: layered, green: dimer); (c) Powder X-ray diffraction (XRD) patterns for Cs3Bi2I9 polymorphs; (ac) reprinted with permission from Ref. [75]. Copyright 2019, American Chemical Society. (d) PL excitation (PLE) and PL spectrum of Cs3BiBr6 NCs. PL at an excitation wavelength of 360 nm and PLE at an emission wavelength of 440 nm; (e) PL and PLE spectra of Cs3Bi2Br9 NCs. PL under an excitation wavelength of 358 nm and PLE at an emission wavelength of 461 nm; (d,e) reprinted with permission from Ref. [87]. Copyright 2020, Royal Society of Chemistry. (f) Absorption (solid lines) and PL (dashed lines) spectra of different levels of Mn-doped Cs3Bi2I9 NCs, respectively. Inset: the appearance of different Mn-doped Cs3Bi2I9 NCs; (g) emission mechanisms of Cs3Bi2I9 with Mn2+ doping; (f,g) reprinted with permission from Ref. [88]. Copyright 2021, Wiley-VCH.
Materials 16 06317 g003
Metal-cation doping is an efficient approach for enhancing the optoelectronic properties of PeNCs. Liu et al. synthesized Mn2+-doped Cs3Bi2I9 NCs to improve these properties [88]. In this method, MnI2 was added to the Bi-oleate solution, and the addition of Mn2+ ions suppressed and eliminated the formation of CsI impurities arising from the decomposition of the intermediate Cs3BiI6 species during synthesis. Figure 3f shows the absorption and PL spectra as a function of the Mn2+ doping concentration. Compared with the absorption spectrum of the undoped Cs3Bi2I9 NCs, that of Mn2+-doped Cs3Bi2I9 was red-shifted, accompanied by double peaks at 530–540 and 610–620 nm. This can be attributed to emission from a spin-forbidden 4T1-6A1 Mn d-d transition resulting from the transfer of exciton energy from the host Cs3Bi2I9 to the Mn2+ dopants, as commonly observed in Mn2+-doped LHP NCs [89,90,91]. Additionally, with increasing Mn2+ concentration from 1 to 5%, the intensity of the PL spectrum increased with a small spectral blue shift from 620 to 610 nm and an improved PLQY from 0.54% (undoped) to 1.57% (5% doped). Figure 3g illustrates the emission mechanism of the Mn2+-doped Cs3Bi2I9 NCs. When the Mn2+-doped Cs3Bi2I9 NCs were excited at ~510 nm, energy was transferred from the Cs3Bi2I9 host to the Mn2+ dopants, resulting in emission from the Mn2+. Exciton relaxation occurs through three pathways: (i) recombination within the host material assisted by phonons; (ii) non-radiative recombination involving trap sites; and (iii) radiative emission through the Mn d-d transition (4T1-6A1), facilitated by energy transfer to Mn2+. Through these exciton relaxation pathways, the Mn2+-doped Cs3Bi2I9 NCs exhibited two emission peaks attributed to radiative recombination within the Cs3Bi2I9 host and Mn2+ dopants, and an increase in radiative recombination leads to an enhanced PLQY.
Despite significant progress in the field of Bi-PeNCs, their PLQY remains generally low [69,70,71,72,75,87,92]. This is primarily attributed to residual surface states, strong photon-phonon coupling, and dangling bonds present on the PeNC surfaces [93,94]. To address this issue, the synthesis of MA3Bi2(Cl, Br)9 NCs was proposed through a collaborative ligand-assisted reprecipitation (Co-LARP) method, which leads to an improvement in the PLQY through surface passivation [95]. The added Cl anions serve as ligands on the surface of the PeNCs, reducing surface defects and inducing a blue-shifted emission spectrum. Notably, when the Cl/(Cl + Br) ratio reached 33%, the MA3Bi2(Cl, Br)9 NCs exhibited a maximum PLQY of 54.1%, and the PL intensity remained at 88% after 12 h under 365 nm UV irradiation.
Bi-PeNCs exhibit various crystal structures depending on their constituent elements. Their optical, photophysical, and electronic properties can be altered using different synthesis methods and doping strategies. Various approaches offer the potential for advancement by tuning and enhancing their optoelectronic characteristics. Although there are no reported cases of Bi-PeNC-based LEDs and their optical properties are inferior to those of LHP NCs, the low cost, non-toxicity, and high stability of Bi suggest that Bi-PeNCs are promising candidates for Pb replacement.

2.3. Sb-Based Perovskite Nanocrystals

Cs3Sb2X9 (X = Cl, Br, and I), which has a trigonal crystal structure with space group P 3 ¯ m1 (no. 164), consists of two [Sb2X9]3− polyhedra within a single unit [96]. In the crystal structure of Cs3Sb2X9 shown in Figure 4a, two Sb3+ ions are located on the body diagonal of the unit cell, whereas the eight top corners are occupied by Cs+ ions. The connected [Sb2X9]3− polyhedra form a bilayer structure that is stacked together. Cs3Sb2X9 has a layered structure and exhibits triangular symmetry.
A simple reprecipitation method did not yield Sb-PeNCs without the assistance of long-chain ligands or excessive ligand usage. Cs3Sb2Br9 NCs can be synthesized using the ligand-assisted reprecipitation (LARP) method. This enables room-temperature PeNC synthesis without traditional heating, an inert gas atmosphere, or injection conditions [97]. LARP involves the direct addition of a reaction precursor solution to a vigorously stirred, poor solvent to rapidly achieve a supersaturated state (Figure 4b). To initiate the LARP synthesis process, a clear precursor solution was prepared by dissolving a mixture of SbBr3, CsBr, and OAm in a good solvent such as N,N-dimethylformamide (DMF) or dimethylsulfoxide (DMSO). Subsequently, a fixed quantity of the prepared precursor solution was carefully dropped into a vigorously stirred solution of octane and OA, which serves as a poor solvent for the precursors. Supersaturation induces the immediate and efficient recrystallization of Cs3Sb2Br9 NCs, leading to their successful synthesis. By utilizing a mixture of octane and OA as ligands, Cs3Sb2Br9 NCs with PLQYs of ~20% were synthesized. Moreover, by further optimizing the synthesis conditions of a precursor concentration of 0.033 mM, an octane/OA ratio of 10:1, and a room-temperature reaction, monodispersed Cs3Sb2Br9 NCs were obtained with a high PLQY of 46% and a uniform size of 3.07 nm [98].
Figure 4. (a) Crystal structure of Cs3Sb2X9 (X = Cl, Br, and I), with a specific view of its xy plane illustrating the spatial distribution of Cs, Sb, and X atoms within the crystal lattice and layered form; (a) reprinted with permission from Ref. [96]. Copyright 2022, MDPI. (b) Schematic image of the Cs3Sb2Br9 NC precursor reaction system for the ligand-assisted reprecipitation (LARP) technique; (b) reprinted with permission from Ref. [98]. Copyright 2017, American Chemical Society. (c) The PL mechanism of Mn2+-doped K3SbCl6 NCs; (d) XRD patterns of as-prepared Mn2+-doped K3SbCl6 NCs at different doping concentrations; (e) PL spectra of Mn2+-doped K3SbCl6 NCs at different doping concentrations under excitation at 365 nm; (ce) Reprinted with permission from Ref. [99]. Copyright 2020, Elsevier.
Figure 4. (a) Crystal structure of Cs3Sb2X9 (X = Cl, Br, and I), with a specific view of its xy plane illustrating the spatial distribution of Cs, Sb, and X atoms within the crystal lattice and layered form; (a) reprinted with permission from Ref. [96]. Copyright 2022, MDPI. (b) Schematic image of the Cs3Sb2Br9 NC precursor reaction system for the ligand-assisted reprecipitation (LARP) technique; (b) reprinted with permission from Ref. [98]. Copyright 2017, American Chemical Society. (c) The PL mechanism of Mn2+-doped K3SbCl6 NCs; (d) XRD patterns of as-prepared Mn2+-doped K3SbCl6 NCs at different doping concentrations; (e) PL spectra of Mn2+-doped K3SbCl6 NCs at different doping concentrations under excitation at 365 nm; (ce) Reprinted with permission from Ref. [99]. Copyright 2020, Elsevier.
Materials 16 06317 g004
This doping strategy was proven to be highly effective in enhancing the optical properties of Sb-PeNCs synthesized by the hot injection method. Pristine K3SbCl6 NCs displayed a broad PL spectrum associated with self-trapped excitons (STE) and achieved a moderate PLQY of 22.3% [99]. However, through the incorporation of Mn2+ ions as dopants, the Mn2+-doped K3SbCl6 NCs exhibited a remarkable PLQY improvement to 37.2%, attaining white light emission. Furthermore, Mn2+-doped K3SbCl6 NCs introduce additional red emission associated with the intrinsic transition (4T16A1) of Mn2+ ions [100,101]. A primary dynamic process for the Mn2+-doped K3SbCl6 NCs was proposed (Figure 4c). Initially, upon excitation, the excited charge carriers cause rapid elastic distortions, leading to the formation of STE states owing to strong electron-phonon interactions within the K3SbCl6 host. Simultaneously, an energy transfer process occurs between the Mn2+ dopant ions and the surrounding STE states, resulting in the excitation of the Mn2+ dopants.
This excitation leads to an efficient red emission spectrum through the intrinsic transitions (4T16A1) of the Mn2+ ions, contributing to an observed additional red emission in the Mn2+-doped K3SbCl6 NCs. The interplay of these dynamic processes not only enhances the overall PLQY but also induces the versatile luminescent properties of Mn2+-doped K3SbCl6 NCs, offering possibilities for optoelectronic applications. The energy transfer process was further validated using excitation spectra obtained by selectively monitoring the energy transfer from the STE at 440 nm and the intrinsic transition of Mn2+ ions at 600 nm. Both emission spectra exhibited similar broad excitation bands, confirming the occurrence of energy transfer between these two entities. The emission decay was significantly faster when the STE was monitored at 440 nm with increasing concentrations of Mn2+. This indicates that the excited energy level (4T16A1) of the Mn2+ ions primarily originates from the energy transferred from the STE of the K3SbCl6 NCs host. The faster emission decay suggests an efficient energy transfer from the perovskite host to the Mn2+ ions, confirming the occurrence of energy transfer between them. By controlling the doping concentration of Mn2+ ions, it was possible to achieve white-light emission with a PLQY of 37.2%. XRD analysis of the 4.2% Mn2+-doped K3SbCl6 NCs revealed that the diffraction peaks at approximately 15.1°, 24.4°, 29.3°, 34.0°, and 46.0° remained similar to those of the undoped K3SbCl6 NCs (Figure 4d). These diffraction peaks are associated with the zero-dimensional (0D) monoclinic K3Sb2Cl6 perovskite structure (PDF#24-0833). Interestingly, the diffraction peaks of the Mn2+-doped K3SbCl6 NCs showed minimal changes at Mn2+ ion doping concentrations of 0.12%, 1.5%, and 3.7%, and a clear red shift in the diffraction peaks was observed when the Mn2+ ion doping concentration exceeded 4.2%. This observation provides further evidence that Mn2+ ions were successfully incorporated into the lattice of the 0D K3SbCl6 host. Furthermore, the relative intensity of the STE emission, compared to the intrinsic transition of Mn2+ ions, decreased progressively as the Mn2+ ion doping concentration increased from 0.12% to 4.2% (Figure 4e). This decrease signifies the occurrence of energy transfer from the K3SbCl6 host to Mn2+.
Sb-PeNCs are environmentally friendly B-site substitutes and are capable of forming an A3B(III)2X9 structure by the heterovalent replacement of Pb cations. One of the advantages of this structure is the feasibility of low-temperature synthesis and greater thermodynamic stability compared with the 2D perovskite phase. However, the lack of a suitable synthesis method for producing high-quality Sb-PeNCs remains a limitation, leading to poor morphology and hindering their application. A doping strategy shows great potential for solid-state lighting applications but may require different approaches for high-purity displays.

2.4. Cu-Based Perovskite Nanocrystals

Transition metals such as Cu have gained considerable attention because of their low toxicity, abundance, excellent stability, and high PLQYs compared to other transition metal-based MHPs [102,103,104]. The small ionic radius of Cu, measuring 0.77 nm for Cu+ and 0.73 nm for Cu2+, prevents it from conforming to a tolerance factor of 1 and maintaining a 3D structure. Therefore, Cu mainly has electronic 0D or 1D structures, which have special optical properties owing to their large exciton binding energies, the AxCuyXx+y structure in monovalent form, and the AxCuyXx+2y structure in divalent form [102,105].
In the synthesis of Cu(II)-PeNCs, such as Cs2CuX4, Cs2CuI4 could not be obtained because pure Cu(II) diiodide is unstable and decomposes easily into iodine and Cu(I) iodide. Cs2CuCl4 exhibited a broad PL spectral range of 400–650 nm, with a PLQY of 51.8%. Therefore, extensive research has focused on Cu (I)-PeNCs owing to the instability and broad emission spectra of Cu(II)-PeNCs [106].
Cu(I)-PeNCs exhibit structural variations depending on the A-site. When Cs is located at the A-site of Cu-PeNC, Cs3Cu2X5 and CsCu2X3 structures are observed (Figure 5a) [107]. The space groups of the structures were identified from their XRD patterns (Figure 5b) [108]. Cs3Cu2I5 belongs to the Pnma orthorhombic space group, and its crystal structure consists of [Cu2I5]3− cluster units composed of tetrahedral [CuI4] and trigonal planar [CuI3] units that are separated from the Cs+ ions. This arrangement gives rise to an electronic 0D structure [109]. CsCu2I3 belongs to the Cmcm space group of orthorhombic crystal systems and exhibits a 1D chain-like structure. This structure is formed by tetrahedral [CuI4]3− units that share a common edge, connecting them in a linear structure. By contrast, Cu(I)-PeNCs with Rb or K at the A-site follow the A2CuX3 structure, which belongs to the Pnma space group. They exhibit a 1D shape where tetrahedral [CuX4]3− units share only the common vertex of the halides, resulting in a chain-like structure [110]. Overall, Cu(I)-PeNCs exhibit diverse crystal structures, including 0D and 1D forms, depending on the specific combination of the A-site and X-site compositions.
In the excited state, Cu-PeNCs show reorganization owing to the Jahn–Teller distortion that forms STE energy states and induces a Stokes shift, representing the difference between the absorption and emission spectra. In the case of Cs3Cu2X5, the absorption of photon energy converts Cu(I) 3d10 to Cu(II) 3d9, which changes the electronic configuration and induces a Jahn–Teller distortion (Figure 5c) [111]. Similarly, the conduction and valence bands of CsCu2I3 are also dispersed along the direction parallel to the tetrahedral [Cu2I6]4− chain, but when it is excited, the electrons are localized due to strong Coulombic coupling and large structural distortion of the [Cu2I6]4− tetrahedron, and the structure is reorganized [112,113,114]. When the lattice distortion energy is high, charges are trapped because the trapped state is more stable than the strained state. Therefore, the Jahn–Teller distortion, which forms STE states, effectively confines charges to localized regions. Simultaneously, as STEs are formed, the ground state energy increases owing to the lattice strain energy. Consequently, STE emission exhibits a large Stokes shift [115,116]. The emission energy (Eem) from the STE states can be expressed as follows [116]:
Eem = Eg − Eb − Est − Ed
where Eg, Eb, Est, and Ed represent the bandgap energy, the exciton binding energy, the self-trapping energy, and the lattice strain energy, respectively. The smaller distortion in CsCu2I3 compared to the octahedral distortion in Cs3Cu2I5 indicates a smaller bandgap of CsCu2I3 (Figure 5d,e) [117]. Moreover, the extended PL decay lifetime of the time-resolved photoluminescence (TrPL) in Cs3Cu2I5 compared to that in CsCu2I3 can be attributed to two key factors: the augmented exciton binding energy and significant lattice distortion. Cs3Cu2I5 exhibits an enhanced exciton binding energy of 224 meV, while CsCu2I3 has a relatively low binding energy of 123 meV. The higher exciton binding energy in Cs3Cu2I5 decelerates the exciton decay rate, resulting in an extended PL lifetime [118,119].
Figure 5. (a) Crystal structures of 0D Cs3Cu2I5 and 1D CsCu2I3; (a) Reprinted with permission from Ref. [107]. Copyright 2019, American Chemical Society. (b) XRD patterns of Cs3Cu2I5 and CsCu2I3; (b) reprinted with permission from Ref. [108]. Copyright 2022, American Chemical Society. (c) Schematic illustration of the principle of STE. (i) A self-trapped state; (ii) a ground state; (iii) charge density for conduction and valence bands; (d) PL and PLE spectra of Cs3Cu2I5; (e) PL and PLE spectra of CsCu2I3; (d,e) Reprinted with permission from Ref. [117]. Copyright 2021, American Chemical Society (f) PLE (Cl = 320 nm, Br = 290 nm, I = 305 nm) and PL (Cl = 515 nm, Br = 461 nm, I = 445 nm) spectra of Cs3Cu2X5 at room temperature; (c,f) Reprinted with permission from Ref. [111]. Copyright 2020, American Chemical Society.
Figure 5. (a) Crystal structures of 0D Cs3Cu2I5 and 1D CsCu2I3; (a) Reprinted with permission from Ref. [107]. Copyright 2019, American Chemical Society. (b) XRD patterns of Cs3Cu2I5 and CsCu2I3; (b) reprinted with permission from Ref. [108]. Copyright 2022, American Chemical Society. (c) Schematic illustration of the principle of STE. (i) A self-trapped state; (ii) a ground state; (iii) charge density for conduction and valence bands; (d) PL and PLE spectra of Cs3Cu2I5; (e) PL and PLE spectra of CsCu2I3; (d,e) Reprinted with permission from Ref. [117]. Copyright 2021, American Chemical Society (f) PLE (Cl = 320 nm, Br = 290 nm, I = 305 nm) and PL (Cl = 515 nm, Br = 461 nm, I = 445 nm) spectra of Cs3Cu2X5 at room temperature; (c,f) Reprinted with permission from Ref. [111]. Copyright 2020, American Chemical Society.
Materials 16 06317 g005
By tuning the halogen anions (X = Cl, Br, and I), the emission wavelength of Cs3Cu2X5 can be controlled, ranging from the green wavelength region (515 nm, Cs3Cu2Cl5) to the blue wavelength region (445 nm, Cs3Cu2I5) (Figure 5f) [111]. This spectral tunability of Cs3Cu2X5 is attributed to the effect of the halogen size and electronegativity on the electronic structure. During electron excitation, as the halogen radius increases and electronegativity decreases, stronger hybridization between the halogen p-orbital and the Cu 3d orbital is induced [120,121]. This enhanced hybridization leads to a decrease in the valence band maximum (VBM) of Cs3Cu2X5. Contrary to the general trend, this decrease in the VBM increases the band gap, leading to a blue shift in the emission wavelength. This phenomenon can be attributed to electronic interactions between the halogen and Cu atoms, which modulate the energy levels and affect the emission properties.
A hot-injection method was used to synthesize 0D Cs3Cu2X5 and 1D CsCu2X3 by controlling the reaction temperature [118]. In this synthesis method, nucleation was induced within a short time by injecting a Cs-oleate into a Cu-oleate containing OAm and OA ligand molecules as quickly as possible and then cooling the solution. When the temperature of Cu-oleate reached 70 °C and the Cs-oleate solution was rapidly injected, 0D Cs3Cu2I5 NCs were synthesized, and when injected at a high temperature of 110 °C, 1D CsCu2I3 was synthesized as the final product. The synthesized Cs3Cu2I5 NCs exhibited blue PL at 441 nm and a PLQY of 67%, whereas the 1D CsCu2I3 NCs exhibited yellow PL at 553 nm and a PLQY of ~5%.
Using an anti-solvent precipitation method, 0D Cs3Cu2X5 and 1D CsCu2X3 can also be synthesized by adjusting the concentration of the precursor solution and the amount of anti-solvent [122]. The CsI and CuI precursors were dissolved in a good solvent, such as DMF or DMSO, and then an anti-solvent such as ethanol was added to precipitate the Cu-PeNC. 1D CsCu2I3 can be obtained in a low-concentration precursor solution because Cu+ precipitates rapidly when an antisolvent is added to a low-concentration precursor. As the concentration of the precipitated Cu+ ions increases, the number of ions participating in the crystallization process increases, and 1D CsCu2I3 is synthesized. By contrast, 0D Cs3Cu2I5 can be obtained from a high-concentration precursor solution (0.4 M) without a significant change in the Cs:Cu ratio when an antisolvent is added. The synthesized Cs3Cu2I5 NCs exhibited a PL emission peak at 440 nm with a full width at half maximum (FWHM) of 73 nm and a PLQY of 67%, whereas 1D CsCu2I3 exhibited a PL emission peak at 580 nm with a broad FWHM of 125 nm and a low PLQY of 7.4%.
During the hot-injection process, metal halide doping, such as with a ZnI2 precursor, can lead to a surface passivation effect in Cs3Cu2X5 NCs [123]. Additionally, Zn cation doping resulted in a tetrahedral coordination arrangement, and Cu+ ions were heterosubstituted with Zn2+ ions. This doping enriched the excited electron density, thereby enhancing the STE emission. Moreover, the incorporation of ZnI2 results in enriched I anions, which can effectively reduce the number of non-radiative centers caused by the iodine vacancies. Zn-doped Cs3Cu2I5 NCs exhibited a blue emission peak at ~440 nm with an FWHM of 73 nm. The PLQY increased from 72.2% for Cs3Cu2I5 NCs to 92.8% for the Zn-doped Cs3Cu2I5 NCs.
Cu-PeNCs can be synthesized in both the monovalent Cu+ and divalent Cu2+ forms using hot injection and antisolvent precipitation methods. In Cu(I)-PeNCs, lattice distortion arises from the structural characteristics of the PeNCs, resulting in STEs. This phenomenon contributes to a high PLQY and a prolonged PL lifetime resulting from highly localized electrons and holes. Moreover, in contrast to the typical red shift observed during the change in halide composition from Cl to I, the observed blue shift in the emission spectrum can be attributed to structural transformations and significant exciton effects. The emission wavelength and crystal structure can be controlled by adjusting the synthesis conditions, precursor ratios, and halide composition. Additionally, defect passivation can be facilitated through metal halide doping, which enhances the optical properties of the Cu-PeNCs. However, the broad FWHM of the emission spectrum stemming from the presence of STEs remains a major challenge to be solved using a novel approach in materials engineering.

2.5. Double Perovskite Nanocrystals

Pb-free double perovskites consist of two different cations occupying the B-sites of an A2BB’X6 structure, providing high flexibility in the selection of B-site cations. The double perovskite structure maintains electrical neutrality by substituting two B2+ ions with a B+ cation and a B3+ cation, forming an A2B(I)B(III)X6 structure (Figure 6a, left). The B-site can accommodate monovalent cations such as K+, Na+, and Ag+ and trivalent cations such as Bi3+, In3+, and Sb3+. Another structural variant is vacancy-ordered double perovskites, where two B2+ ions are replaced by a single tetravalent cation, such as Sn4+ or Ti4+, and B-site vacancies, forming the A2B(IV)X6 structure (Figure 6a, right). Additionally, a 3D-layered structure with one vacancy can be achieved when four B2+ ions are substituted by one B2+ ion and two B3+ ions, as represented by vacancy-ordered quadruple perovskites A4B(II)B(III)2X12. The incorporation of one Cu2+ ion and two Sb3+ ions in place of four Pb2+ ions generates a single vacancy, resulting in the formation of a monoclinic layered double perovskite structure such as Cs4CuSb2Cl12 (Figure 6b) [124]. Thus, the combination of B-site cations and vacancies results in a wide range of structural variations.
Double PeNCs have gained significant attention as promising candidates to replace Pb because of their tunable optoelectronic properties; high stability against heat, humidity, and light; and versatility in selecting B-site cations [125,126,127,128]. However, their commercial application is hindered by poor PL characteristics, which are attributed to intrinsic and surface defects, indirect bandgaps, and forbidden transitions [129,130,131,132]. To address these challenges, various strategies have been explored to improve their optoelectronic properties.
Ag+Bi3+-PeNCs exhibit low PLQYs, mainly owing to their indirect bandgap characteristics [126,129,133]. The absorption tail observed in the long-wavelength region of the absorption spectrum indicates sub-bandgap absorption [70,132]. Yang et al. introduced changes in the indirect bandgap characteristics by doping In3+ ions in Cs2AgBiCl6 NCs [134]. Optical analysis of the Cs2AgInxBi1−xCl6 NCs revealed blue-shifted absorption peaks with increasing In3+ doping concentration, and sharp absorption peaks were observed at x values of 0.75 and 0.9, indicating possible direct bandgap characteristics. The intensity of the PL emission peak observed at ~400 nm increased with increasing In3+ doping concentration, and a second peak appeared at ~570 nm at x > 0.75.
To gain a deeper understanding of the optical mechanism and bandgap characteristics, DFT calculations were performed using the generalized gradient approximation of the Perdew–Burke–Ernzerhof functional. Figure 6c,d depict the band structures of Cs2AgInxBi1−xCl6 NCs with In ratios of 0% and 75%, respectively. The band structure of Cs2AgBiCl6 NC exhibits an indirect bandgap, whereas that of Cs2AgIn0.75Bi0.25Cl6 NC shows a nearly direct bandgap. In the Cs2AgBiCl6 NC, the VBM at the Γ point has Cl-p, Ag-d, and Bi-s characteristics, whereas the CBM at L has Bi-p, Cl-p, and Ag-s characteristics. Conversely, in the Cs2AgIn0.75Bi0.25Cl6 NC, the VBM at the Γ point has Cl-p, Ag-d, and Bi-s characteristics, similar to those of the Cs2AgBiCl6 NC, whereas the CBM comprises In-s, Cl-p, and Ag-s characteristics, similar to those of the Cs2AgInCl6 NC. This indicates the possibility of direct transitions, as observed in the band structure of the Cs2AgIn0.75Bi0.25Cl6 NC. Similar to Cs2AgInCl6 NC, where a parity-forbidden transition from the VBM to the CBM occurs at the Γ point, Cs2AgIn0.75Bi0.25Cl6 NC provides a pathway for relaxation through parity-forbidden transitions, leading to emission at approximately 570 nm [135]. These results suggest that the dual PL emission peaks can be attributed to the emissions resulting from both band- and parity-forbidden transitions. The PLQYs showed a pronounced enhancement above 0.75 in the doping ratio, resulting in a PLQY of 36.6% in Cs2AgIn0.75Bi0.25Cl6 NC, but a slight increase below 0.5 in the doping ratio, resulting in a PLQY of ~10% in Cs2AgIn0.5Bi0.5Cl6 NC. This substantial improvement in the PLQY above 0.75 in the ratio originates from the transition from an indirect bandgap to a direct bandgap (Figure 6e). Band structure diagrams in Figure 6f,g depict the photoinduced charge carrier transportation mechanism and provide an intuitive understanding of the absorption and emission processes based on the indirect and direct bandgap characteristics of the double PeNCs. In the case of an indirect bandgap, both direct and indirect transitions can occur. These indirect bandgap transitions require phonon participation, resulting in a low PLQY [136,137]. However, for a direct bandgap, absorption occurs through direct transitions, and the excited carriers can undergo non-radiative relaxation processes to the forbidden states, resulting in two forms of luminescence: direct bandgap relaxation and relaxation to the forbidden states. Radiative relaxation of the band edge and forbidden states improved the PLQY of Cs2AgInxBi1−xCl6 when x was greater than 0.75.
Figure 6. (a) Crystal structures of double perovskites (A2B(I)B(III)X6) and vacancy-ordered double perovskites (A2B(IV)X6); (b) crystal structures of Cs4Cu(II)Sb(III)2Cl12 and Cs2Ag(I)Sb(III)Cl6; (b) reprinted with permission from Ref. [124]. Copyright 2020, American Chemical Society. Band structures of (c) Cs2AgBiCl6 and (d) Cs2AgIn0.75Bi0.25Cl6; (e) PLQE value of Cs2AgInxBi1−xCl6 (x = 0, 0.25, 0.5, 0.75, and 0.9) NCs; (ce) reprinted with permission from Ref. [134]. Copyright 2018, American Chemical Society. Charge carrier transportation mechanism models of (f) indirect bandgap Cs2AgInxBi1−xCl6 NCs and (g) direct bandgap Cs2AgInxBi1−xCl6 NCs.
Figure 6. (a) Crystal structures of double perovskites (A2B(I)B(III)X6) and vacancy-ordered double perovskites (A2B(IV)X6); (b) crystal structures of Cs4Cu(II)Sb(III)2Cl12 and Cs2Ag(I)Sb(III)Cl6; (b) reprinted with permission from Ref. [124]. Copyright 2020, American Chemical Society. Band structures of (c) Cs2AgBiCl6 and (d) Cs2AgIn0.75Bi0.25Cl6; (e) PLQE value of Cs2AgInxBi1−xCl6 (x = 0, 0.25, 0.5, 0.75, and 0.9) NCs; (ce) reprinted with permission from Ref. [134]. Copyright 2018, American Chemical Society. Charge carrier transportation mechanism models of (f) indirect bandgap Cs2AgInxBi1−xCl6 NCs and (g) direct bandgap Cs2AgInxBi1−xCl6 NCs.
Materials 16 06317 g006
Cs2SnX6-based vacancy-ordered double perovskites utilizing Sn4+ have attracted significant research interest owing to their air and moisture stability, high absorption coefficients, and carrier mobility [138,139,140]. Wang et al. reported a synthesis method for Cs2SnI6 perovskites that could be adjusted from spherical NCs, nanorods, nanowires, and nanobelts to nanoplatelets by varying the reaction time during the hot-injection process [141]. However, the PLQY of the Cs2SnI6 NCs was found to be as low as 0.48%. Cs2SnI6 NCs can also be synthesized by a ligand-mediated approach using the hot-injection method, similar to conventional LHP NC synthesis [142,143,144]. By adjusting the ratio of OA to octylamine, the structure of the Cs2SnI6 NCs could be controlled between the 3D and 2D forms. When only OA was used during the synthesis, 3D Cs2SnI6 NCs were formed, whereas the combination of OA and octylamine resulted in the formation of 2D Cs2SnI6 nanoplatelets. The one-layer 2D Cs2SnI6 exhibited a PL emission peak at 643 nm, while the two-layer 2D Cs2SnI6 displayed a PL emission peak at 742 nm. The PLQYs were measured as 28% for the one-layer Cs2SnI6 and 16% for the two-layer Cs2SnI6. These results can be attributed to the quantum confinement effect of the quantum-well structure, which enhances the exciton binding energy and improves the PL characteristics [145,146].
Similar to the behavior of other PeNCs, doping of metal ions, such as Sb3+, in Cs2SnCl6 NCs is an effective strategy for enhancing their luminescent properties [147]. Pristine Cs2SnCl6 NCs synthesized by the hot-injection process exhibited a sharp absorption edge at 310 nm with an absorption tail at longer wavelengths (Figure 7a). Under 365 nm excitation, the pristine Cs2SnCl6 NCs showed a PL spectrum peak at 438 nm with a PLQY of 4.37%. Sb3+ doping of the Cs2SnCl6 NCs was achieved with Sb/Sn precursor ratios of 0.05 and 0.1, where SbCl3 was mixed concurrently with SnCl2. The Sb3+-doped Cs2SnCl6 NCs exhibited double PL peaks at 438 and 615 nm, and the PL intensity at 615 nm increased with increasing Sb3+ concentration from 0.05 to 0.1 (Figure 7b). Further investigations revealed that the 615 nm emission could be attributed to the triplet states of STE induced by Sb3+ doping. The emission at 615 nm originates from recombination with the ground states through transitions from the singlet states of STE with emission at ~450 nm to the triplet states of STE with emission at 615 nm, resulting in broad emission wavelengths. Through Sb3+ doping at a ratio of 0.1, the PLQY of Sb3+-doped Cs2SnCl6 was enhanced to 8.25% compared to that of pristine Cs2SnCl6 (4.37%).
Vacancy-ordered quadruple perovskite A4B(I)B(III)2X12 has attracted significant research interest because of its large compositional space, direct bandgap characteristics, and excellent structural stability [148,149,150,151]. It consists of [M(II)X6]4− octahedra layers sandwiched between [M(III)X6]3− octahedra layers, crystallizing in the R 3 ¯ m space group symmetry perpendicular to the <111> direction of the cubic perovskite structure. However, their practical applications have been hindered by their poor luminescence properties. Therefore, research has focused on enhancing the luminescent characteristics of vacancy-ordered quadruple perovskites through metal cation doping [124,152,153,154,155,156,157]. The synthesis of Cs4Cd1−xMnxBi2Cl12 NCs was proposed using a hot-injection method with benzoyl chloride [152]. The [Mn]/([Mn] + [Cd]) ratios in the final products were quantified as 0%, 1.7%, 10.1%, 34.9%, 69.0%, and 100% using inductively coupled plasma atomic emission spectroscopy. Despite an increase in the Mn ratio, the absorption and PLE spectra showed minimal changes, whereas there was a distinct red shift in the PL emission spectra (Figure 7c). The Cs4Cd1−xMnxBi2Cl12 NCs with x = 0.349 exhibit the highest PLQY of 4.6%. However, as the value of x increased further, the PLQY decreased. In Figure 7c, the Stokes shift originated from a difference between absorption and emission peaks. The influences of Stokes shift and self-absorption in PeLEDs have been commonly explained in two ways. First, self-absorption can cause photon loss due to a difference between the numbers of internally generated and externally emitted photons that can be measured as PLQYs [158]. However, the large Stokes shift could prevent self-absorption loss because the absorption spectrum does not overlap with the emission wavelength. Therefore, reduced self-absorption resulting from the Stokes shift could enhance emission efficiencies in LEDs. Secondly, however, the Stokes shift could be observed along with the increase in Auger recombination affected by the large exciton binding energy, which can lead to increased efficiency roll-off in LEDs [159,160,161]. Therefore, further systematic analysis should be studied to clearly establish the effects of the Stokes shift on LED performance.
Figure 7. (a) Absorption (gray line) excitation (dashed line, emission at 438 nm) and PL (excitation at 365 nm) spectra of undoped Cs2SnCl6 NCs (inset: UV irradiation under 365 nm UV lamp); (b) PLE (dashed line, emission at 615 nm; gray line, emission at 438 nm) and PL (excitation at 365 nm) spectra of Sb-doped Cs2SnCl6 NCs (inset: UV irradiation of Sb/Sn = 0.1 NCs under 365 nm UV lamp); (a,b) reprinted with permission from Ref. [147]. Copyright 2019, American Chemical Society. (c) Absorption, PLE, and PL spectra of Cs4Cd1−xMnxCl12 NCs (inset: UV irradiation under UV light); (c) reprinted with permission from Ref. [152]. Copyright 2020, Royal Society of Chemistry. (d) Room temperature PL spectra of Cs4CdBi2Cl12 (pristine) and Cs4Cd0.6Mn0.4Bi2Cl12 (Mn-substituted) nanosheet (NS); (d) reprinted with permission from Ref. [157]. Copyright 2023, Royal Society of Chemistry.
Figure 7. (a) Absorption (gray line) excitation (dashed line, emission at 438 nm) and PL (excitation at 365 nm) spectra of undoped Cs2SnCl6 NCs (inset: UV irradiation under 365 nm UV lamp); (b) PLE (dashed line, emission at 615 nm; gray line, emission at 438 nm) and PL (excitation at 365 nm) spectra of Sb-doped Cs2SnCl6 NCs (inset: UV irradiation of Sb/Sn = 0.1 NCs under 365 nm UV lamp); (a,b) reprinted with permission from Ref. [147]. Copyright 2019, American Chemical Society. (c) Absorption, PLE, and PL spectra of Cs4Cd1−xMnxCl12 NCs (inset: UV irradiation under UV light); (c) reprinted with permission from Ref. [152]. Copyright 2020, Royal Society of Chemistry. (d) Room temperature PL spectra of Cs4CdBi2Cl12 (pristine) and Cs4Cd0.6Mn0.4Bi2Cl12 (Mn-substituted) nanosheet (NS); (d) reprinted with permission from Ref. [157]. Copyright 2023, Royal Society of Chemistry.
Materials 16 06317 g007
Similar PL characteristics were observed for Cs4Cd1−xMnxBi2Cl12 nanosheets (NS) with the same constituent elements. Bhardwaj et al. synthesized Cs4Cd0.6Mn0.4Bi2Cl12 and 2- to 3-layer-thick 2D NS [157]. Optical characterization of pristine Cs4CdBi2Cl12 and Cs4Cd0.6Mn0.4Bi2Cl12 NS revealed that a weak PL emission appeared at ~605 nm for the pristine Cs4CdBi2Cl12 NS, whereas the Cs4Cd0.6Mn0.4Bi2Cl12 NS showed a gradual increase in the PL emission peak intensity at ~601 nm with increasing Mn doping concentration, unaffected by different excitation wavelengths (Figure 7d). These red-shifted PL spectra due to Mn doping were attributed to an inter-Mn-distance-induced Mn–Mn coupling interaction, leading to a larger dd splitting of the Mn2+ ion centers, consequently reducing the gap in the 4T1-6A1 electronic transition of Mn2+ ions [152,162,163]. In Cs4CdBi2Cl12, the emission arises from the STE state, whereas the emission in Cs4Cd1−xMnxBi2Cl12 arises from energy transfer to Mn2+ in the STE states [153].
Double PeNCs also have limitations in practical applications owing to issues concerning instability due to metal ion oxidation, the formation of deep trap states, and a broad FWHM of the emission spectrum. To overcome these issues, ongoing research aims to enhance the photoelectric performance of double PeNCs through various doping strategies, synthesis methods, and control of variables such as ligands [164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182]. The characterization of changes induced by process variables will lead to diverse applications in various optoelectronic fields.
Based on the above discussion, the optical and photophysical properties of the diverse PeNCs are summarized in Table 1. It is worth noting that certain instances of Pb-free PeNCs exhibit superior PL properties in comparison to LHP NCs. In particular, for remarkable PLQYs of Cu-based PeNCs, it could be necessary to focus on mainly developing device engineering such as device physics, light out-coupling, and charge balance as future promising eco-friendly LEDs. On the other hand, for relatively low PLQYs of Sn-, Bi-, and double PeNCs, research on further developing their intrinsic material properties might be more crucial and would be helpful to further enhance PL properties through promising strategies such as defect passivation, ligand modification, surface engineering, and improved synthesis techniques.

3. Pb-Free Perovskite Nanocrystal-Based LEDs

Pb-free perovskite LEDs (PeLEDs) exhibit great potential as next-generation LEDs. This is due to their nontoxicity, cost-effectiveness, solution processability, facile emission wavelength tunability, and defect tolerance. Research on colloidal PeNC-based LEDs has recently gained attention in parallel with the rapid advancement of solution-based synthesis of bulk MHPs [183,184,185,186,187,188,189]. Optimizing the dimensionality of MHPs provides another avenue for tuning their electronic and optical properties [190,191,192]. PeNCs possess distinct electrical and optical properties attributable to their large exciton binding energy, tunable NC size, and reduced dimensionality [19,193].
Most Sn-based MHPs are solution-processed for fabricating LEDs because of their simplicity and low-cost processability. However, the rapid crystallization of Sn-MHPs can lead to poor morphology and low film quality, which degrades the performance of LEDs [194,195,196]. To overcome the intrinsic limitations of Sn-based PeLEDs, high-vacuum vapor deposition is an alternative approach for the fabrication of Sn-PeNC films [197,198]. The fabrication of PeLEDs based on CsSnBr3 PeNCs can be achieved by thermal evaporation with in-situ annealing [199]. To promote the formation of CsSnBr3 NCs films, a dual source was used for the thermal evaporation of CsBr and SnBr2. Subsequently, in the post-synthetic annealing stage, the initially stacked CsBr and SnBr2 layers interact at the interface, resulting in the formation of a CsSnBr3 layer through interdiffusion of the constituent elements facilitated by thermal annealing [183]. When testing various post-synthetic annealing temperatures, the CsSnBr3 films annealed at 85 °C showed a significant enhancement in crystallinity. This is evident from the increased diffraction intensity of the (100) and (200) planes and a reduced FWHM. The improved crystallinity can be attributed to an enhanced reaction between the CsBr and SnBr2 precursors, which occurs as a result of interdiffusion. In addition, the CsSnBr3 NCs obtained by the annealing-temperature process exhibited an average crystal size of 39.9 nm. The devices were fabricated using the structure of ITO/MoO3/4,4′-cyclohexylidenebis[N,N-bis(p-tolyl)aniline]/tris(4-carbazoyl-9-ylphenyl)amine (TCTA)/CsSnBr3 NCs/1,3,5-tri(m-pyrid-3-yl-phenyl)benzene (TmPyPB)/LiF/Al (Figure 8a). The CsSnBr3 NC-based LED demonstrated a turn-on voltage of 5.5 V (at 1 cd/m2) and a maximum luminance of 43 cd/m2 at an applied voltage of 10 V. Notably, it achieved a high current efficiency (CE) of 0.34 cd/A and an external quantum efficiency (EQE) of 0.16% (Figure 8b). At an applied voltage from 4.5 to 10 V, the CsSnBr3 NC-based LED exhibited consistent electroluminescence (EL) emission peaks at approximately 675 nm, indicating exceptional EL stability.
Bi-based PeNC LEDs have not been reported yet. This may be caused by the low PLQY resulting from surface defects, which are due to the intrinsic self-trapping-induced phonon-mediated nonradiative process [93,95]. It could also be attributed to strong photon-phonon coupling and dangling bonds present on the PeNC surfaces [70,93,94]. As there are no reports on Bi-based PeNCs LED yet, further research is necessary to improve PLQY and engineering for LED applications.
Sb3+, possessing an electronegativity similar to that of Pb2+ ions, demonstrates exceptional structural stability and outstanding luminescent properties, making it a promising candidate for incorporation as an emitting layer (EML) in LEDs. Stable violet-emitting Cs3Sb2Br9 NCs, synthesized using the LARP method, exhibited a high PLQY of 51.2% [186]. Moreover, the Cs3Sb2Br9 NCs exhibited a moderate PL stability change of ~17.5% after 73 h of continuous UV irradiation (365 nm, 30 W). Furthermore, the PL intensity of the Cs3Sb2Br9 NCs was maintained at approximately 80% of the initial emission intensity when 0.5 mL of deionized water was added to the Cs3Sb2Br9 NC solution for 45 h, while the CsPbBr3 NCs experienced rapid fluorescence quenching, retaining only ~9% of the initial emission intensity for only 4 h. Moreover, the fabricated PeLED with the structure ITO/ZnO/PEI/Cs3Sb2Br9 NCs/TCTA/MoO3/Al exhibited a maximum EQE of ~0.206% and luminance of 29.6 cd/m2 at an operating voltage of 8.0 V (Figure 8c,d). The EL intensity peak was maintained at a wavelength of 408 nm after 6 h of device operation, and the PeLED demonstrated excellent operational EL stability, with only ~10% emission decay.
Cu-based PeLEDs are attracting attention as blue-light emitters owing to their nontoxicity, high stability, and excellent optoelectronic properties. Wang et al. reported a Cu(I)-based PeLED with the structure ITO/NiO/Cs3Cu2I5 NCs/2,2′,2″-(1,3,5-Benzinetriyl)-tris(1-phenyl-1-H-benzimidazole) (TPBi)/LiF/Al [187]. Cs3Cu2I5 NCs synthesized by the hot-injection method were employed as an EML fabricated by the spin-coating process. The energy band diagram of ITO/NiO/Cs3Cu2I5 NCs/TPBi/LiF/Al was obtained using ultraviolet photoelectron spectroscopy (UPS) (Figure 8e). The device structure illustrates that p-NiO possesses a relatively high ionization potential energy, making it a suitable hole-donating and electron-blocking layer because of its compatible electron affinity with the Cs3Cu2I5 NCs [200]. Similarly, TPBi served the dual function of electron donation and a hole-blocking layer owing to its low, lowest unoccupied molecular orbital and highest occupied molecular orbital energy levels. Electrons were injected from the upper TPBi layer into the Cs3Cu2I5 NC EML, facilitating EL emission through the recombination of confined electrons and holes.
Cs3Cu2I5 NC-LED exhibited a Commission International de I’Eclairage (CIE) coordinate of (0.16, 0.07) at 7.0 V, in compliance with the blue coordinate of the National Television Standards Committee (NTSC) standard, reaching ~99% of the NTSC standard and contributing to a wide color gamut for Pb-free PeLEDs. Additionally, the maximum EQE of the Cs3Cu2I5 NC-LEDs was ~1.12% (Figure 8f), with high reproducibility and a small relative deviation of 15.7%. Furthermore, Cs3Cu2I5 NC-LED demonstrated continuous, relatively stable device operation for 170 h at an applied voltage of 6.7 V while maintaining an EL wavelength of 445 nm. These results indicate the excellent device performance and operational stability of Cs3Cu2I5 NC-LEDs.
Double PeNCs face challenges in LED implementation, owing to their intrinsic and surface defects, indirect bandgaps, and suppressed emission characteristics [129,130,131,132]. However, attempts have been made to improve the optical properties of LED implementation through bandgap modification by metal cation doping and changes in carrier dynamics. Zhang et al. first developed electrically excited white LEDs based on Cs2AgIn0.9Bi0.1Cl6 NCs without a phosphor [188]. Cs2AgIn0.9Bi0.1Cl6 NCs were synthesized using a hot-injection method with TMS-Cl injection. The Cs2AgIn0.9Bi0.1Cl6 NCs exhibited two absorption peaks at 333 and 367 nm, a broad dual-color emission from STE emission, and a PLQY of 31.4%. The energy band structure of Cs2AgIn0.9Bi0.1Cl6 NCs was determined by UPS measurements, revealing a CBM of −7.69 eV and a VBM of −2.98 eV obtained by a Tauc plot. DFT calculations demonstrated a direct bandgap for Cs2AgIn0.9Bi0.1Cl6, and band symmetry analysis indicated that the parity-forbidden transition is broken at the deep-energy level of the conduction band because of the contribution of the Bi atom [135]. A device was fabricated with the structure ITO/Poly(9-vinylcarbazole) (PVK)/Cs2AgIn0.9Bi0.1Cl6 NCs/TPBi/LiF/Al (Figure 8g). The EL emission spectrum remained unchanged with increasing applied voltage, and the turn-on voltage and maximum luminance of 34.7 cd/m2 were measured at 10 V. The white Cs2AgIn0.9Bi0.1Cl6 NCs LED exhibited a maximum CE and EQE of 0.058 cd/A and 0.064%, respectively, and emitted white light with CIE chromaticity coordinates (0.32, 0.32) (Figure 8h).
A PeLED utilizing 2D Cs2AgInxBi1−xCl6 alloyed double perovskite nanoplatelets (NPLs) was also fabricated [189]. A PeLED with the structure ITO/PEDOT: PSS/poly[(9,9-dioctylfluorenyl-2,7-diyl)-co-(4,40-(N-(4-sec-butylphenyl)) diphenylamine)] (TFB)/PVK:NPLs/ZnO/Al was fabricated. To enhance the dispersity of the NPLs and prevent electron leakage into the TFB layer, PVK was employed as a buffer component in the emissive layer, whereas ethanol-based ZnO was spin-coated as the solution-processed ETL. The Cs2AgIn0.9Bi0.1Cl6 NPL-based PeLED exhibited a single EL peak at 557 nm with a maximum luminance of 58 cd/m2 and an EQE of 0.01%.

4. Conclusions

In this review, we presented a comprehensive overview of recent advances in Pb-free PeNCs and their LED applications. PeNCs exhibit size-tunable properties, high PL efficiency, stability, and large exciton binding energy, making them promising candidates for various optoelectronic applications. However, the toxicity of Pb in LHP NCs has limited their commercialization and driven research on Pb-free MHPs.
We systematically reviewed the crystal structure, synthesis methods, efficient strategies for optoelectronic performance, and LED applications of Pb-free PeNCs based on promising B-site metal candidates as substitutes for Pb. We also discussed the synthetic strategies employed in the development of Pb-free NCs involving the optimization of crystal structures that tune their optoelectronic properties, including the PL characteristics and charge-carrier dynamics mechanism. However, the lower performance and stability, along with a broad emission wavelength of Pb-free PeNCs in comparison to LHP NCs, remain a challenge, and LED applications of Pb-free PeNCs are currently limited. These issues present difficulties for their commercialization.
Therefore, further investigations are required to explore strategies aimed at improving the material stability, optoelectronic properties, and device performance of Pb-free PeNCs. Active research and enhancements need to be undertaken with respect to electrical transport capabilities, optical properties, and stability to contribute to the realization and performance enhancement of LEDs. As mentioned earlier, research endeavors aimed at improving these properties necessitate the further development and utilization of techniques such as doping strategy, ligand exchange, and polymer composites (Figure 9). The directions of this material research could be a promising starting point for improving the performance of Pb-free PeNC-based LEDs. Furthermore, as research progresses on performance, stability improvement, and narrow emission spectra, Pb-free PeNCs will demonstrate the huge potential for LED development, offering prospects for a more sustainable and eco-friendly future in the field of display-related industries.

Author Contributions

D.-Y.K., J.-G.J., Y.-J.L. and M.-H.P. conceived the outlines and contents; Writing—Original Draft Preparation, D.-Y.K., J.-G.J., and Y.-J.L.; Writing—Review and Editing, D.-Y.K. and M.-H.P.; Supervision, M.-H.P.; Project administration, M.-H.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was mainly supported by Samsung Research Funding, the Incubation Center of Samsung Electronics under Project Number SRFC-TC2103-03, and the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (2021R1A6A1A10044154).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stranks, S.D.; Snaith, H.J. Metal-Halide Perovskites for Photovoltaic and Light-Emitting Devices. Nat. Nanotechnol. 2015, 10, 391–402. [Google Scholar] [CrossRef] [PubMed]
  2. Xiao, Z.; Kerner, R.A.; Zhao, L.; Tran, N.L.; Lee, K.M.; Koh, T.-W.; Scholes, G.D.; Rand, B.P. Efficient Perovskite Light-Emitting Diodes Featuring Nanometre-Sized Crystallites. Nat. Photonics 2017, 11, 108–115. [Google Scholar] [CrossRef]
  3. Liu, X.K.; Xu, W.; Bai, S.; Jin, Y.; Wang, J.; Friend, R.H.; Gao, F. Metal Halide Perovskites for Light-Emitting Diodes. Nat. Mater. 2021, 20, 10–21. [Google Scholar] [CrossRef]
  4. Park, M.H. 3D and 2D Metal Halide Perovskites for Blue Light-Emitting Diodes. Materials 2022, 15, 4571. [Google Scholar] [CrossRef] [PubMed]
  5. Lee, M.M.; Teuscher, J.; Miyasaka, T.; Murakami, T.N.; Snaith, H.J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643–647. [Google Scholar] [CrossRef]
  6. Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Sequential Deposition as a Route to High-Performance Perovskite-Sensitized Solar Cells. Nature 2013, 499, 316–319. [Google Scholar] [CrossRef] [PubMed]
  7. Ono, L.K.; Qi, Y.; Liu, S.F. Progress toward Stable Lead Halide Perovskite Solar Cells. Joule 2018, 2, 1961–1990. [Google Scholar] [CrossRef]
  8. Yakunin, S.; Sytnyk, M.; Kriegner, D.; Shrestha, S.; Richter, M.; Matt, G.J.; Azimi, H.; Brabec, C.J.; Stangl, J.; Kovalenko, M.V.; et al. Detection of X-ray Photons by Solution-Processed Lead Halide Perovskites. Nat. Photonics 2015, 9, 444–449. [Google Scholar] [CrossRef]
  9. Tian, W.; Zhou, H.; Li, L. Hybrid Organic–Inorganic Perovskite Photodetectors. Small 2017, 13, 1702107. [Google Scholar] [CrossRef]
  10. Wang, Y.; Li, X.; Nalla, V.; Zeng, H.; Sun, H. Solution-Processed Low Threshold Vertical Cavity Surface Emitting Lasers from All-Inorganic Perovskite Nanocrystals. Adv. Funct. Mater. 2017, 27, 1605088. [Google Scholar] [CrossRef]
  11. Kanhere, P.; Chen, Z. A Review on Visible Light Active Perovskite-Based Photocatalysts. Molecules 2014, 19, 19995–20022. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, T.H.; Park, I.; Lee, K.H.; Sim, J.H.; Park, M.H.; Han, T.H.; Paik, U.; Jang, J.; Park, H.B.; Kim, Y.H. Investigating the Interfacial Properties of Halide Perovskite/TiOx Heterostructures for Versatile Photocatalytic Reactions under Sunlight. Nanoscale 2023, 15, 7710–7714. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, H.; Liu, Z.; Xu, W.; Yang, L.; Liu, Y.; Yao, D.; Zhang, D.; Zhang, H.; Yang, B. Engineering the Photoluminescence of CsPbX3 (X = Cl, Br, and I) Perovskite Nanocrystals Across the Full Visible Spectra with the Interval of 1 Nm. ACS Appl. Mater. Interfaces 2019, 11, 14256–14265. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Akkerman, Q.A.; Rainò, G.; Kovalenko, M.V.; Manna, L. Genesis, Challenges and Opportunities for Colloidal Lead Halide Perovskite Nanocrystals. Nat. Mater. 2018, 17, 394–405. [Google Scholar] [CrossRef] [PubMed]
  16. Huang, H.; Susha, A.S.; Kershaw, S.V.; Hung, T.F.; Rogach, A.L. Control of Emission Color of High Quantum Yield CH3NH3PbBr3 Perovskite Quantum Dots by Precipitation Temperature. Adv. Sci. 2015, 2, 1500194. [Google Scholar] [CrossRef]
  17. Mir, W.J.; Warankar, A.; Acharya, A.; Das, S.; Mandal, P.; Nag, A. Colloidal Thallium Halide Nanocrystals with Reasonable Luminescence, Carrier Mobility and Diffusion Length. Chem. Sci. 2017, 8, 4602–4611. [Google Scholar] [CrossRef]
  18. 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]
  19. 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]
  20. Kovalenko, M.V.; Protesescu, L.; Bodnarchuk, M.I. Properties and Potential Optoelectronic Applications of Lead Halide Perovskite Nanocrystals. Science 2017, 358, 745–750. [Google Scholar] [CrossRef]
  21. Talapin, D.V.; Lee, J.S.; Kovalenko, M.V.; Shevchenko, E.V. Prospects of Colloidal Nanocrystals for Electronic and Optoelectronic Applications. Chem. Rev. 2010, 110, 389–458. [Google Scholar] [CrossRef] [PubMed]
  22. Kambhampati, P. Multiexcitons in Semiconductor Nanocrystals: A Platform for Optoelectronics at High Carrier Concentration. J. Phys. Chem. Lett. 2012, 3, 1182–1190. [Google Scholar] [CrossRef]
  23. Polavarapu, L.; Nickel, B.; Feldmann, J.; Urban, A.S. Advances in Quantum-Confined Perovskite Nanocrystals for Optoelectronics. Adv. Energy Mater. 2017, 7, 1700267. [Google Scholar] [CrossRef]
  24. Imran, M.; Caligiuri, V.; Wang, M.; Goldoni, L.; Prato, M.; Krahne, R.; De Trizio, L.; Manna, L. Benzoyl Halides as Alternative Precursors for the Colloidal Synthesis of Lead-Based Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2018, 140, 2656–2664. [Google Scholar] [CrossRef] [PubMed]
  25. Maddalena, F.; Xie, A.; Chin, X.Y.; Begum, R.; Witkowski, M.E.; Makowski, M.; Mahler, B.; Drozdowski, W.; Springham, S.V.; Rawat, R.S.; et al. Deterministic Light Yield, Fast Scintillation, and Microcolumn Structures in Lead Halide Perovskite Nanocrystals. J. Phys. Chem. C 2021, 125, 14082–14088. [Google Scholar] [CrossRef]
  26. Wang, Y.; Singh, K.; Li, J.; Dong, Y.; Wang, X.; Pina, J.M.; Yu, Y.; Sabatini, R.; Liu, Y.; Ma, D.; et al. In Situ Inorganic Ligand Replenishment Enables Bandgap Stability in Mixed-Halide Perovskite Quantum Dot Solids. Adv. Mater. 2022, 34, 2200854. [Google Scholar] [CrossRef]
  27. Kim, J.S.; Heo, J.-M.; Park, G.-S.; Woo, S.-J.; Cho, C.; Yun, H.J.; Kim, D.-H.; Park, J.; Lee, S.-C.; Park, S.-H.; et al. Ultra-Bright, Efficient and Stable Perovskite Light-Emitting Diodes. Nature 2022, 611, 688–694. [Google Scholar] [CrossRef]
  28. Hao, M.; Bai, Y.; Zeiske, S.; Ren, L.; Liu, J.; Yuan, Y.; Zarrabi, N.; Cheng, N.; Ghasemi, M.; Chen, P.; et al. Ligand-Assisted Cation-Exchange Engineering for High-Efficiency Colloidal Cs1−xFAxPbI3 Quantum Dot Solar Cells with Reduced Phase Segregation. Nat. Energy 2020, 5, 79–88. [Google Scholar] [CrossRef]
  29. Xie, Z.; Park, H.Y.; Kim, H.; Park, J.; Song, D.; Lee, J.; Chetan, L.; Jin, S.H. Stable Dual Cations Perovskite Nanocrystals as Absorbers for Perovskite Solar Cells with Efficiencies Exceeding 24%. Bull. Korean Chem. Soc. 2023, 44, 658–664. [Google Scholar] [CrossRef]
  30. Kang, C.H.; Dursun, I.; Liu, G.; Sinatra, L.; Sun, X.; Kong, M.; Pan, J.; Maity, P.; Ooi, E.N.; Ng, T.K.; et al. High-Speed Colour-Converting Photodetector with All-Inorganic CsPbBr3 Perovskite Nanocrystals for Ultraviolet Light Communication. Light. Sci. Appl. 2019, 8, 94. [Google Scholar] [CrossRef]
  31. Shyamal, S.; Dutta, S.K.; Pradhan, N. Doping Iron in CsPbBr3 Perovskite Nanocrystals for Efficient and Product Selective CO2 Reduction. J. Phys. Chem. Lett. 2019, 10, 7965–7969. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, X.; Sun, C.; Zhang, Y.; Wu, H.; Ji, C.; Chuai, Y.; Wang, P.; Wen, S.; Zhang, C.; Yu, W.W. Bright Perovskite Nanocrystal Films for Efficient Light-Emitting Devices. J. Phys. Chem. Lett. 2016, 7, 4602–4610. [Google Scholar] [CrossRef] [PubMed]
  33. Jia, C.; Das, P.; Zeng, Q.; Gabriel, J.C.P.; Tay, C.Y.; Lee, J.M. Activated Recovery of PVC from Contaminated Waste Extension Cord-Cable Using a Weak Acid. Chemosphere 2022, 303, 134878. [Google Scholar] [CrossRef]
  34. Li, J.; Cao, H.-L.; Jiao, W.-B.; Wang, Q.; Wei, M.; Cantone, I.; Lü, J.; Abate, A. Biological Impact of Lead from Halide Perovskites Reveals the Risk of Introducing a Safe Threshold. Nat. Commun. 2020, 11, 310. [Google Scholar] [CrossRef] [PubMed]
  35. Wani, A.L.; Ara, A.; Usmani, J.A. Lead Toxicity: A Review. Interdiscip. Toxicol. 2015, 8, 55–64. [Google Scholar] [CrossRef] [PubMed]
  36. Sun, J.; Yang, J.; Lee, J.I.; Cho, J.H.; Kang, M.S. Lead-Free Perovskite Nanocrystals for Light-Emitting Devices. J. Phys. Chem. Lett. 2018, 9, 1573–1583. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, F.; Ma, Z.; Shi, Z.; Chen, X.; Wu, D.; Li, X.; Shan, C. Recent Advances and Opportunities of Lead-Free Perovskite Nanocrystal for Optoelectronic Application. Energy Mater. Adv. 2021, 2021, 5198145. [Google Scholar] [CrossRef]
  38. Tang, H.; Xu, Y.; Hu, X.; Hu, Q.; Chen, T.; Jiang, W.; Wang, L.; Jiang, W. Lead-Free Halide Double Perovskite Nanocrystals for Light-Emitting Applications: Strategies for Boosting Efficiency and Stability. Adv. Sci. 2021, 8, 2004118. [Google Scholar] [CrossRef]
  39. Li, X.; Gao, X.; Zhang, X.; Shen, X.; Lu, M.; Wu, J.; Shi, Z.; Colvin, V.L.; Hu, J.; Bai, X.; et al. Lead-Free Halide Perovskites for Light Emission: Recent Advances and Perspectives. Adv. Sci. 2021, 8, 2003334. [Google Scholar] [CrossRef]
  40. De Souza Carvalho, T.A.; Magalhaes, L.F.; do Livramento Santos, C.I.; de Freitas, T.A.Z.; Carvalho Vale, B.R.; Vale da Fonseca, A.F.; Schiavon, M.A. Lead-Free Metal Halide Perovskite Nanocrystals: From Fundamentals to Applications. Chem. Eur. J. 2023, 29, e202202518. [Google Scholar] [CrossRef]
  41. Li, C.; Lu, X.; Ding, W.; Feng, L.; Gao, Y.; Guo, Z. Formability of ABX3 (X = F, Cl, Br, I) Halide Perovskites. Acta Crystallogr. Sect. B Struct. Sci. 2008, 64, 702–707. [Google Scholar] [CrossRef] [PubMed]
  42. Mao, X.; Sun, L.; Wu, T.; Chu, T.; Deng, W.; Han, K. First-Principles Screening of All-Inorganic Lead-Free ABX3 Perovskites. J. Phys. Chem. C 2018, 122, 7670–7675. [Google Scholar] [CrossRef]
  43. Li, C.; Soh, K.C.K.; Wu, P. Formability of ABO3 Perovskites. J. Alloys Compd. 2004, 372, 40–48. [Google Scholar] [CrossRef]
  44. Li, W.; Ionescu, E.; Riedel, R.; Gurlo, A. Can We Predict the Formability of Perovskite Oxynitrides from Tolerance and Octahedral Factors? J. Mater. Chem. A 2013, 1, 12239–12245. [Google Scholar] [CrossRef]
  45. Sun, Q.; Yin, W.J. Thermodynamic Stability Trend of Cubic Perovskites. J. Am. Chem. Soc. 2017, 139, 14905–14908. [Google Scholar] [CrossRef]
  46. Lin, H.; Zhou, C.; Tian, Y.; Siegrist, T.; Ma, B. Low-Dimensional Organometal Halide Perovskites. ACS Energy Lett. 2018, 3, 54–62. [Google Scholar] [CrossRef]
  47. Zhou, C.; Lin, H.; He, Q.; Xu, L.; Worku, M.; Chaaban, M.; Lee, S.; Shi, X.; Du, M.H.; Ma, B. Low Dimensional Metal Halide Perovskites and Hybrids. Mater. Sci. Eng. R Rep. 2019, 137, 38–65. [Google Scholar] [CrossRef]
  48. Luo, S.-Q.; Wang, J.-F.; Yang, B.; Yuan, Y.-B. Recent Advances in Controlling the Crystallization of Two-Dimensional Perovskites for Optoelectronic Device. Front. Phys. 2019, 14, 53401. [Google Scholar] [CrossRef]
  49. Morss, L.R.; Siegal, M.; Stenger, L.; Edelstein, N. Preparation of Cubic Chloro Complex Compounds of Trivalent Metals: Cs2NaMCl6. Inorg. Chem. 1970, 9, 1771–1775. [Google Scholar] [CrossRef]
  50. Kieslich, G.; Sun, S.; Cheetham, A.K. An Extended Tolerance Factor Approach for Organic-Inorganic Perovskites. Chem. Sci. 2015, 6, 3430–3433. [Google Scholar] [CrossRef]
  51. Filip, M.R.; Giustino, F. Computational Screening of Homovalent Lead Substitution in Organic-Inorganic Halide Perovskites. J. Phys. Chem. C 2016, 120, 166–173. [Google Scholar] [CrossRef]
  52. Hoefler, S.F.; Trimmel, G.; Rath, T. Progress on Lead-Free Metal Halide Perovskites for Photovoltaic Applications: A Review. Monatshefte Chem. 2017, 148, 795–826. [Google Scholar] [CrossRef] [PubMed]
  53. Xiao, Z.; Song, Z.; Yan, Y. From Lead Halide Perovskites to Lead-Free Metal Halide Perovskites and Perovskite Derivatives. Adv. Mater. 2019, 31, e1803792. [Google Scholar] [CrossRef] [PubMed]
  54. Yadav, R.; Srivastava, A.; Abraham, J.A.; Sharma, R.; Dar, S.A. First-Principles Calculations to Investigate Structural, Electronic, Thermoelectric, and Optical Properties of Heavy Thallium Perovskite TlPbX3 (X = Cl, Br, I). Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2022, 283, 115781. [Google Scholar] [CrossRef]
  55. Ning, W.; Gao, F. Structural and Functional Diversity in Lead-Free Halide Perovskite Materials. Adv. Mater. 2019, 31, e1900326. [Google Scholar] [CrossRef] [PubMed]
  56. Cheng, Y.-H.; Moriyama, R.; Ebe, H.; Mizuguchi, K.; Yamakado, R.; Nishitsuji, S.; Chiba, T.; Kido, J. Two-Step Crystallization for Low-Oxidation Tin-Based Perovskite Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2022, 14, 22941–22949. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, M.Y.; Lin, J.T.; Hsu, C.S.; Chang, C.K.; Chiu, C.W.; Chen, H.M.; Chou, P.T. Strongly Coupled Tin-Halide Perovskites to Modulate Light Emission: Tunable 550–640 nm Light Emission (FWHM 36–80 nm) with a Quantum Yield of up to 6.4%. Adv. Mater. 2018, 30, e1706592. [Google Scholar] [CrossRef]
  58. Liu, Q.; Yin, J.; Zhang, B.B.; Chen, J.K.; Zhou, Y.; Zhang, L.M.; Wang, L.M.; Zhao, Q.; Hou, J.; Shu, J.; et al. Theory-Guided Synthesis of Highly Luminescent Colloidal Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2021, 143, 5470–5480. [Google Scholar] [CrossRef]
  59. Wang, A.; Guo, Y.; Muhammad, F.; Deng, Z. Controlled Synthesis of Lead-Free Cesium Tin Halide Perovskite Cubic Nanocages with High Stability. Chem. Mater. 2017, 29, 6493–6501. [Google Scholar] [CrossRef]
  60. Giuri, A.; Masi, S.; Listorti, A.; Gigli, G.; Colella, S.; Esposito Corcione, C.; Rizzo, A. Polymeric Rheology Modifier Allows Single-Step Coating of Perovskite Ink for Highly Efficient and Stable Solar Cells. Nano Energy 2018, 54, 400–408. [Google Scholar] [CrossRef]
  61. Lyu, B.; Guo, X.; Gao, D.; Kou, M.; Yu, Y.; Ma, J.; Chen, S.; Wang, H.; Zhang, Y.; Bao, X. Highly-Stable Tin-Based Perovskite Nanocrystals Produced by Passivation and Coating of Gelatin. J. Hazard. Mater. 2021, 403, 123967. [Google Scholar] [CrossRef] [PubMed]
  62. Sulaman, M.; Yang, S.Y.; Zhang, Z.H.; Imran, A.; Bukhtiar, A.; Ge, Z.H.; Tang, Y.; Jiang, Y.R.; Tang, L.B.; Zou, B.S. Lead-Free Tin-Based Perovskites Nanocrystals for High-Performance Self-Driven Bulk-Heterojunction Photodetectors. Mater. Today Phys. 2022, 27, 100829. [Google Scholar] [CrossRef]
  63. Lv, L.; Xu, Y.; Fang, H.; Luo, W.; Xu, F.; Liu, L.; Wang, B.; Zhang, X.; Yang, D.; Hu, W.; et al. Generalized Colloidal Synthesis of High-Quality, Two-Dimensional Cesium Lead Halide Perovskite Nanosheets and Their Applications in Photodetectors. Nanoscale 2016, 8, 13589–13596. [Google Scholar] [CrossRef] [PubMed]
  64. Zakutayev, A.; Caskey, C.M.; Fioretti, A.N.; Ginley, D.S.; Vidal, J.; Stevanovic, V.; Tea, E.; Lany, S. Defect Tolerant Semiconductors for Solar Energy Conversion. J. Phys. Chem. Lett. 2014, 5, 1117–1125. [Google Scholar] [CrossRef] [PubMed]
  65. Wu, C.; Zhang, Q.; Liu, G.; Zhang, Z.; Wang, D.; Qu, B.; Chen, Z.; Xiao, L. From Pb to Bi: A Promising Family of Pb-Free Optoelectronic Materials and Devices. Adv. Energy Mater. 2020, 10, 1902496. [Google Scholar] [CrossRef]
  66. Brandt, R.E.; Stevanović, V.; Ginley, D.S.; Buonassisi, T. Identifying Defect-Tolerant Semiconductors with High Minority-Carrier Lifetimes: Beyond Hybrid Lead Halide Perovskites. MRS Commun. 2015, 5, 265–275. [Google Scholar] [CrossRef]
  67. Lyu, M.; Yun, J.-H.; Cai, M.; Jiao, Y.; Bernhardt, P.V.; Zhang, M.; Wang, Q.; Du, A.; Wang, H.; Liu, G.; et al. Organic–Inorganic Bismuth (III)-Based Material: A Lead-Free, Air-Stable and Solution-Processable Light-Absorber beyond Organolead Perovskites. Nano Res. 2016, 9, 692–702. [Google Scholar] [CrossRef]
  68. Hoye, R.L.Z.; Brandt, R.E.; Osherov, A.; Stevanović, V.; Stranks, S.D.; Wilson, M.W.B.; Kim, H.; Akey, A.J.; Perkins, J.D.; Kurchin, R.C.; et al. Methylammonium Bismuth Iodide as a Lead-Free, Stable Hybrid Organic–Inorganic Solar Absorber. Chem. Eur. J. 2016, 22, 2605–2610. [Google Scholar] [CrossRef]
  69. Zhang, Y.; Yin, J.; Parida, M.R.; Ahmed, G.H.; Pan, J.; Bakr, O.M.; Brédas, J.-L.; Mohammed, O.F. Direct-Indirect Nature of the Bandgap in Lead-Free Perovskite Nanocrystals. J. Phys. Chem. Lett. 2017, 8, 3173–3177. [Google Scholar] [CrossRef]
  70. Yang, B.; Chen, J.; Hong, F.; Mao, X.; Zheng, K.; Yang, S.; Li, Y.; Pullerits, T.; Deng, W.; Han, K. Lead-Free, Air-Stable All-Inorganic Cesium Bismuth Halide Perovskite Nanocrystals. Angew. Chem. Int. Ed. 2017, 56, 12471–12475. [Google Scholar] [CrossRef]
  71. Leng, M.; Chen, Z.; Yang, Y.; Li, Z.; Zeng, K.; Li, K.; Niu, G.; He, Y.; Zhou, Q.; Tang, J. Lead-Free, Blue Emitting Bismuth Halide Perovskite Quantum Dots. Angew. Chem. Int. Ed. 2016, 55, 15012–15016. [Google Scholar] [CrossRef] [PubMed]
  72. Pal, J.; Bhunia, A.; Chakraborty, S.; Manna, S.; Das, S.; Dewan, A.; Datta, S.; Nag, A. Synthesis and Optical Properties of Colloidal M3Bi2I9 (M = Cs, Rb) Perovskite Nanocrystals. J. Phys. Chem. C 2018, 122, 10643–10649. [Google Scholar] [CrossRef]
  73. Lou, Y.; Fang, M.; Chen, J.; Zhao, Y. Formation of Highly Luminescent Cesium Bismuth Halide Perovskite Quantum Dots Tuned by Anion Exchange. Chem. Commun. 2018, 54, 3779–3782. [Google Scholar] [CrossRef] [PubMed]
  74. Leng, M.; Yang, Y.; Zeng, K.; Chen, Z.; Tan, Z.; Li, S.; Li, J.; Xu, B.; Li, D.; Hautzinger, M.P.; et al. All-Inorganic Bismuth-Based Perovskite Quantum Dots with Bright Blue Photoluminescence and Excellent Stability. Adv. Funct. Mater. 2018, 28, 1704446. [Google Scholar] [CrossRef]
  75. Creutz, S.E.; Liu, H.; Kaiser, M.E.; Li, X.; Gamelin, D.R. Structural Diversity in Cesium Bismuth Halide Nanocrystals. Chem. Mater. 2019, 31, 4685–4697. [Google Scholar] [CrossRef]
  76. Hong, K.-H.; Kim, J.; Debbichi, L.; Kim, H.; Im, S.H. Band Gap Engineering of Cs3Bi2I9 Perovskites with Trivalent Atoms Using a Dual Metal Cation. J. Phys. Chem. C 2017, 121, 969–974. [Google Scholar] [CrossRef]
  77. Kim, J.; Lee, S.C.; Lee, S.H.; Hong, K.H. Importance of Orbital Interactions in Determining Electronic Band Structures of Organo-Lead Iodide. J. Phys. Chem. C 2015, 119, 4627–4634. [Google Scholar] [CrossRef]
  78. Lehner, A.J.; Fabini, D.H.; Evans, H.A.; Hébert, C.A.; Smock, S.R.; Hu, J.; Wang, H.; Zwanziger, J.W.; Chabinyc, M.L.; Seshadri, R. Crystal and Electronic Structures of Complex Bismuth Iodides A3Bi2I9 (A = K, Rb, Cs) Related to Perovskite: Aiding the Rational Design of Photovoltaics. Chem. Mater. 2015, 27, 7137–7148. [Google Scholar] [CrossRef]
  79. McCall, K.M.; Stoumpos, C.C.; Kontsevoi, O.Y.; Alexander, G.C.B.; Wessels, B.W.; Kanatzidis, M.G. From 0D Cs3Bi2I9 to 2D Cs3Bi2I6Cl3: Dimensional Expansion Induces a Direct Band Gap but Enhances Electron-Phonon Coupling. Chem. Mater. 2019, 31, 2644–2650. [Google Scholar] [CrossRef]
  80. Yu, B.B.; Liao, M.; Yang, J.; Chen, W.; Zhu, Y.; Zhang, X.; Duan, T.; Yao, W.; Wei, S.H.; He, Z. Alloy-Induced Phase Transition and Enhanced Photovoltaic Performance: The Case of Cs3Bi2I9-xBrx Perovskite Solar Cells. J. Mater. Chem. A 2019, 7, 8818–8825. [Google Scholar] [CrossRef]
  81. Hodgkins, T.L.; Savory, C.N.; Bass, K.K.; Seckman, B.L.; Scanlon, D.O.; Djurovich, P.I.; Thompson, M.E.; Melot, B.C. Anionic Order and Band Gap Engineering in Vacancy Ordered Triple Perovskites. Chem. Commun. 2019, 55, 3164–3167. [Google Scholar] [CrossRef] [PubMed]
  82. Murray, C.B.; Sun, S.; Gaschler, W.; Doyle, H.; Betley, T.A.; Kagan, C.R. Colloidal Synthesis of Nanocrystals and Nanocrystal Superlattices. IBM J. Res. Dev. 2001, 45, 47–56. [Google Scholar] [CrossRef]
  83. Park, J.; An, K.; Hwang, Y.; Park, J.E.G.; Noh, H.J.; Kim, J.Y.; Park, J.H.; Hwang, N.M.; Hyeon, T. Ultra-Large-Scale Syntheses of Monodisperse Nanocrystals. Nat. Mater. 2004, 3, 891–895. [Google Scholar] [CrossRef] [PubMed]
  84. Jun, Y.W.; Choi, J.S.; Cheon, J. Shape Control of Semiconductor and Metal Oxide Nanocrystals through Nonhydrolytic Colloidal Routes. Angew. Chem. Int. Ed. 2006, 45, 3414–3439. [Google Scholar] [CrossRef]
  85. Wang, F.; Han, Y.; Lim, C.S.; Lu, Y.; Wang, J.; Xu, J.; Chen, H.; Zhang, C.; Hong, M.; Liu, X. Simultaneous Phase and Size Control of Upconversion Nanocrystals through Lanthanide Doping. Nature 2010, 463, 1061–1065. [Google Scholar] [CrossRef] [PubMed]
  86. Wu, L.; Chen, S.Y.; Fan, F.J.; Zhuang, T.T.; Dai, C.M.; Yu, S.H. Polytypic Nanocrystals of Cu-Based Ternary Chalcogenides: Colloidal Synthesis and Photoelectrochemical Properties. J. Am. Chem. Soc. 2016, 138, 5576–5584. [Google Scholar] [CrossRef] [PubMed]
  87. Lee, D.; Kim, M.H.; Woo, H.Y.; Chae, J.; Lee, D.; Jeon, S.; Oh, S.J.; Paik, T. Heating-up Synthesis of Cesium Bismuth Bromide Perovskite Nanocrystals with Tailored Composition, Morphology, and Optical Properties. RSC Adv. 2020, 10, 7126–7133. [Google Scholar] [CrossRef] [PubMed]
  88. Liu, M.; Ali-Löytty, H.; Hiltunen, A.; Sarlin, E.; Qudsia, S.; Smått, J.H.; Valden, M.; Vivo, P. Manganese Doping Promotes the Synthesis of Bismuth-Based Perovskite Nanocrystals While Tuning Their Band Structures. Small 2021, 17, e2100101. [Google Scholar] [CrossRef]
  89. Meinardi, F.; Akkerman, Q.A.; Bruni, F.; Park, S.; Mauri, M.; Dang, Z.; Manna, L.; Brovelli, S. Doped Halide Perovskite Nanocrystals for Reabsorption-Free Luminescent Solar Concentrators. ACS Energy Lett. 2017, 2, 2368–2377. [Google Scholar] [CrossRef]
  90. Qiao, T.; Parobek, D.; Dong, Y.; Ha, E.; Son, D.H. Photoinduced Mn Doping in Cesium Lead Halide Perovskite Nanocrystals. Nanoscale 2019, 11, 5247–5253. [Google Scholar] [CrossRef]
  91. 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]
  92. Ali, R.F.; Andreu, I.; Gates, B.D. Green Solvent Assisted Synthesis of Cesium Bismuth Halide Perovskite Nanocrystals and the Influences of Slow and Fast Anion Exchange Rates. Nanoscale Adv. 2019, 1, 4442–4449. [Google Scholar] [CrossRef]
  93. McCall, K.M.; Stoumpos, C.C.; Kostina, S.S.; Kanatzidis, M.G.; Wessels, B.W. Strong Electron–Phonon Coupling and Self-Trapped Excitons in the Defect Halide Perovskites A3M2I9 (A = Cs, Rb; M = Bi, Sb). Chem. Mater. 2017, 29, 4129–4145. [Google Scholar] [CrossRef]
  94. 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]
  95. Leng, M.; Yang, Y.; Chen, Z.; Gao, W.; Zhang, J.; Niu, G.; Li, D.; Song, H.; Zhang, J.; Jin, S.; et al. Surface Passivation of Bismuth-Based Perovskite Variant Quantum Dots To Achieve Efficient Blue Emission. Nano Lett. 2018, 18, 6076–6083. [Google Scholar] [CrossRef] [PubMed]
  96. Wu, Y.; Xiang, G.; Zhang, M.; Wei, D.; Cheng, C.; Leng, J.; Ma, H. Electronic Structures and Photoelectric Properties in Cs3Sb2X9 (X = Cl, Br, or I) under High Pressure: A First Principles Study. Nanomaterials 2022, 12, 2982. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, F.; Zhong, H.; Chen, C.; Wu, X.G.; 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]
  98. Zhang, J.; Yang, Y.; Deng, H.; Farooq, U.; Yang, X.; Khan, J.; Tang, J.; Song, H. High Quantum Yield Blue Emission from Lead-Free Inorganic Antimony Halide Perovskite Colloidal Quantum Dots. ACS Nano 2017, 11, 9294–9302. [Google Scholar] [CrossRef]
  99. Shao, H.; Wu, X.; Zhu, J.; Xu, W.; Xu, L.; Dong, B.; Hu, J.; Dong, B.; Bai, X.; Cui, H.; et al. Mn2+ Ions Doped Lead-Free Zero-Dimensional K3SbCl6 Perovskite Nanocrystals towards White Light Emitting Diodes. Chem. Eng. J. 2021, 413, 127415. [Google Scholar] [CrossRef]
  100. Yang, H.; Santra, S.; Holloway, P.H. Syntheses and Applications of Mn-Doped II-VI Semiconductor Nanocrystals. J. Nanosci. Nanotechnol. 2005, 5, 1364–1375. [Google Scholar] [CrossRef] [PubMed]
  101. Beaulac, R.; Archer, P.I.; Gamelin, D.R. Luminescence in Colloidal Mn2+-Doped Semiconductor Nanocrystals. J. Solid. State Chem. 2008, 181, 1582–1589. [Google Scholar] [CrossRef]
  102. Gao, F.; Zhu, X.; Feng, Q.; Zhong, W.; Liu, W.; Xu, H.; Liu, Y. Deep-Blue Emissive Cs3Cu2I5 Perovskites Nanocrystals with 96.6% Quantum Yield via InI3-Assisted Synthesis for Light-Emitting Device and Fluorescent Ink Applications. Nano Energy 2022, 98, 107270. [Google Scholar] [CrossRef]
  103. Qu, J.; Xu, S.; Shao, H.; Xia, P.; Lu, C.; Wang, C.; Ban, D. Recent Progress of Copper Halide Perovskites: Properties, Synthesis and Applications. J. Mater. Chem. C 2023, 11, 6260–6275. [Google Scholar] [CrossRef]
  104. Wang, M.; Wang, W.; Ma, B.; Shen, W.; Liu, L.; Cao, K.; Chen, S.; Huang, W. Lead-Free Perovskite Materials for Solar Cells. Nano-Micro Lett. 2021, 13, 62. [Google Scholar] [CrossRef] [PubMed]
  105. Hui, Y.; Chen, S.; Lin, R.; Zheng, W.; Huang, F. Photophysics in Cs3Cu2I5 and CsCu2I3. Mater. Chem. Front. 2021, 5, 7088–7107. [Google Scholar] [CrossRef]
  106. Yang, P.; Liu, G.; Liu, B.; Liu, X.; Lou, Y.; Chen, J.; Zhao, Y. All-Inorganic Cs2CuX4 (X = Cl, Br, and Br/I) Perovskite Quantum Dots with Blue-Green Luminescence. Chem. Commun. 2018, 54, 11638–11641. [Google Scholar] [CrossRef] [PubMed]
  107. Vashishtha, P.; Nutan, G.V.; Griffith, B.E.; Fang, Y.; Giovanni, D.; Jagadeeswararao, M.; Sum, T.C.; Mathews, N.; Mhaisalkar, S.G.; Hanna, J.V.; et al. Cesium Copper Iodide Tailored Nanoplates and Nanorods for Blue, Yellow, and White Emission. Chem. Mater. 2019, 31, 9003–9011. [Google Scholar] [CrossRef]
  108. Cui, W.; Zhao, J.; Wang, L.; Lv, P.; Li, X.; Yin, Z.; Yang, C.; Tang, A. Unraveling the Phase Transition and Luminescence Tuning of Pb-Free Cs-Cu-I Perovskites Enabled by Reaction Temperature and Polar Solvent. J. Phys. Chem. Lett. 2022, 13, 4856–4863. [Google Scholar] [CrossRef] [PubMed]
  109. Hull, S.; Berastegui, P. 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. 2004, 177, 3156–3173. [Google Scholar] [CrossRef]
  110. Vashishtha, P.; Hooper, T.J.N.; Fang, Y.; Kathleen, D.; Giovanni, D.; Klein, M.; Sum, T.C.; Mhaisalkar, S.G.; Mathews, N.; White, T. Room Temperature Synthesis of Low-Dimensional Rubidium Copper Halide Colloidal Nanocrystals with near Unity Photoluminescence Quantum Yield. Nanoscale 2021, 13, 59–65. [Google Scholar] [CrossRef]
  111. Lian, L.; Zheng, M.; Zhang, P.; Zheng, Z.; Du, K.; Lei, W.; Gao, J.; Niu, G.; Zhang, D.; Zhai, T.; et al. Photophysics in Cs3Cu2X5 (X = Cl, Br, or I): Highly Luminescent Self-Trapped Excitons from Local Structure Symmetrization. Chem. Mater. 2020, 32, 3462–3468. [Google Scholar] [CrossRef]
  112. Xiao, Z.; Du, K.; Meng, W.; Mitzi, D.B.; Yan, Y. Chemical Origin of the Stability Difference between Copper(I)- and Silver(I)-Based Halide Double Perovskites. Angew. Chem. 2017, 129, 12275–12279. [Google Scholar] [CrossRef]
  113. Jun, T.; Sim, K.; Iimura, S.; Sasase, M.; Kamioka, H.; Kim, J.; Hosono, H. Lead-Free Highly Efficient Blue-Emitting Cs3Cu2I5 with 0D Electronic Structure. Adv. Mater. 2018, 30, e1804547. [Google Scholar] [CrossRef]
  114. Ma, Z.; Shi, Z.; Qin, C.C.; Cui, M.; Yang, D.; Wang, X.; Wang, L.; Ji, X.; Chen, X.; Sun, J.; et al. Stable Yellow Light-Emitting Devices Based on Ternary Copper Halides with Broadband Emissive Self-Trapped Excitons. ACS Nano 2020, 14, 4475–4486. [Google Scholar] [CrossRef] [PubMed]
  115. Du, M.H. Emission Trend of Multiple Self-Trapped Excitons in Luminescent 1D Copper Halides. ACS Energy Lett. 2020, 5, 464–469. [Google Scholar] [CrossRef]
  116. Li, S.; Luo, J.; Liu, J.; Tang, J. Self-Trapped Excitons in All-Inorganic Halide Perovskites: Fundamentals, Status, and Potential Applications. J. Phys. Chem. Lett. 2019, 10, 1999–2007. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, X.; Zhou, B.; Chen, X.; Yu, W.W. Reversible Transformation between Cs3Cu2I5and CsCu2I3 Perovskite Derivatives and Its Anticounterfeiting Application. Inorg. Chem. 2022, 61, 399–405. [Google Scholar] [CrossRef]
  118. Cheng, P.; Sun, L.; Feng, L.; Yang, S.; Yang, Y.; Zheng, D.; Zhao, Y.; Sang, Y.; Zhang, R.; Wei, D.; et al. Colloidal Synthesis and Optical Properties of All-Inorganic Low-Dimensional Cesium Copper Halide Nanocrystals. Angew. Chem. 2019, 131, 16233–16237. [Google Scholar] [CrossRef]
  119. Liu, S.; Liu, H.; Zhou, G.; Li, X.; Wang, S. Water-Induced Crystal Phase Transformation of Stable Lead-Free Cu-Based Perovskite Nanocrystals Prepared by One-Pot Method. Chem. Eng. J. 2022, 427, 131430. [Google Scholar] [CrossRef]
  120. Creason, T.D.; McWhorter, T.M.; Bell, Z.; Du, M.H.; Saparov, B. K2CuX3 (X = Cl, Br): All-Inorganic Lead-Free Blue Emitters with Near-Unity Photoluminescence Quantum Yield. Chem. Mater. 2020, 32, 6197–6205. [Google Scholar] [CrossRef]
  121. Li, Y.; Vashishtha, P.; Zhou, Z.; Li, Z.; Shivarudraiah, S.B.; Ma, C.; Liu, J.; Wong, K.S.; Su, H.; Halpert, J.E. 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. 2020, 32, 5515–5524. [Google Scholar] [CrossRef]
  122. Zhao, Z.; Li, X.; Xie, L.; Chen, B.; Jiang, T.; Cao, J.; Zhang, F.; Wang, M.; Wu, Y.; Zheng, H.; et al. Phase Control in the Synthesis of Cesium Copper Iodide Compounds for Their Photoluminescence and Radioluminescence Study. J. Lumin. 2022, 241, 118482. [Google Scholar] [CrossRef]
  123. Qu, K.; Lu, Y.; Ran, P.; Wang, K.; Zhang, N.; Xia, K.; Zhang, H.; Pi, X.; Hu, H.; Yang, Y.; et al. Zn (II)-Doped Cesium Copper Halide Nanocrystals with High Quantum Yield and Colloidal Stability for High-Resolution X-Ray Imaging. Adv. Opt. Mater. 2023, 11, 2202883. [Google Scholar] [CrossRef]
  124. Cai, T.; Shi, W.; Hwang, S.; Kobbekaduwa, K.; Nagaoka, Y.; Yang, H.; Hills-Kimball, K.; Zhu, H.; Wang, J.; Wang, Z.; et al. Lead-Free Cs4CuSb2Cl12 Layered Double Perovskite Nanocrystals. J. Am. Chem. Soc. 2020, 142, 11927–11936. [Google Scholar] [CrossRef] [PubMed]
  125. Slavney, A.H.; Hu, T.; Lindenberg, A.M.; Karunadasa, H.I. A Bismuth-Halide Double Perovskite with Long Carrier Recombination Lifetime for Photovoltaic Applications. J. Am. Chem. Soc. 2016, 138, 2138–2141. [Google Scholar] [CrossRef] [PubMed]
  126. McClure, E.T.; Ball, M.R.; Windl, W.; Woodward, P.M. Cs2AgBiX6 (X = Br, Cl): New Visible Light Absorbing, Lead-Free Halide Perovskite Semiconductors. Chem. Mater. 2016, 28, 1348–1354. [Google Scholar] [CrossRef]
  127. Zhou, J.; Xia, Z.; Molokeev, M.S.; Zhang, X.; Peng, D.; Liu, Q. Composition Design, Optical Gap and Stability Investigations of Lead-Free Halide Double Perovskite Cs2AgInCl6. J. Mater. Chem. A 2017, 5, 15031–15037. [Google Scholar] [CrossRef]
  128. Zhou, L.; Xu, Y.F.; Chen, B.X.; Kuang, D.B.; Su, C.Y. Synthesis and Photocatalytic Application of Stable Lead-Free Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14, e1703762. [Google Scholar] [CrossRef]
  129. Xiao, Z.; Meng, W.; Wang, J.; Yan, Y. Thermodynamic Stability and Defect Chemistry of Bismuth-Based Lead-Free Double Perovskites. ChemSusChem 2016, 9, 2628–2633. [Google Scholar] [CrossRef]
  130. Meng, W.; Wang, X.; Xiao, Z.; Wang, J.; Mitzi, D.B.; Yan, Y. Parity-Forbidden Transitions and Their Impact on the Optical Absorption Properties of Lead-Free Metal Halide Perovskites and Double Perovskites. J. Phys. Chem. Lett. 2017, 8, 2999–3007. [Google Scholar] [CrossRef]
  131. Volonakis, G.; Haghighirad, A.A.; Milot, R.L.; Sio, W.H.; Filip, M.R.; Wenger, B.; Johnston, M.B.; Herz, L.M.; Snaith, H.J.; Giustino, F. Cs2InAgCl6: A New Lead-Free Halide Double Perovskite with Direct Band Gap. J. Phys. Chem. Lett. 2017, 8, 772–778. [Google Scholar] [CrossRef]
  132. Yang, B.; Chen, J.; Yang, S.; Hong, F.; Sun, L.; Han, P.; Pullerits, T.; Deng, W.; Han, K. Lead-Free Silver-Bismuth Halide Double Perovskite Nanocrystals. Angew. Chem. Int. Ed. 2018, 57, 5359–5363. [Google Scholar] [CrossRef] [PubMed]
  133. Creutz, S.E.; Crites, E.N.; De Siena, M.C.; Gamelin, D.R. Colloidal Nanocrystals of Lead-Free Double-Perovskite (Elpasolite) Semiconductors: Synthesis and Anion Exchange to Access New Materials. Nano Lett. 2018, 18, 1118–1123. [Google Scholar] [CrossRef]
  134. Yang, B.; Mao, X.; Hong, F.; Meng, W.; Tang, Y.; Xia, X.; Yang, S.; Deng, W.; Han, K. Lead-Free Direct Band Gap Double-Perovskite Nanocrystals with Bright Dual-Color Emission. J. Am. Chem. Soc. 2018, 140, 17001–17006. [Google Scholar] [CrossRef] [PubMed]
  135. Bekenstein, Y.; Dahl, J.C.; Huang, J.; Osowiecki, W.T.; Swabeck, J.K.; Chan, E.M.; Yang, P.; Alivisatos, A.P. The Making and Breaking of Lead-Free Double Perovskite Nanocrystals of Cesium Silver-Bismuth Halide Compositions. Nano Lett. 2018, 18, 3502–3508. [Google Scholar] [CrossRef] [PubMed]
  136. Hutter, E.M.; Gélvez-Rueda, M.C.; Osherov, A.; Bulović, V.; Grozema, F.C.; Stranks, S.D.; Savenije, T.J. Direct-Indirect Character of the Bandgap in Methylammonium Lead Iodide Perovskite. Nat. Mater. 2017, 16, 115–120. [Google Scholar] [CrossRef] [PubMed]
  137. Manna, D.; Das, T.K.; Yella, A. Tunable and Stable White Light Emission in Bi3+-Alloyed Cs2AgInCl6 Double Perovskite Nanocrystals. Chem. Mater. 2019, 31, 10063–10070. [Google Scholar] [CrossRef]
  138. Guo, F.; Lu, Z.; Mohanty, D.; Wang, T.; Bhat, I.B.; Zhang, S.; Shi, S.; Washington, M.A.; Wang, G.C.; Lu, T.M. A Two-Step Dry Process for Cs2SnI6 Perovskite Thin Film. Mater. Res. Lett. 2017, 5, 540–546. [Google Scholar] [CrossRef]
  139. Ke, J.C.R.; Lewis, D.J.; Walton, A.S.; Spencer, B.F.; O’Brien, P.; Thomas, A.G.; Flavell, W.R. Ambient-Air-Stable Inorganic Cs2SnI6 Double Perovskite Thin Films: Via Aerosol-Assisted Chemical Vapour Deposition. J. Mater. Chem. A 2018, 6, 11205–11214. [Google Scholar] [CrossRef]
  140. Shin, H.O.; Kim, B.M.; Jang, T.; Kim, K.M.; Roh, D.H.; Nam, J.S.; Kim, J.S.; Kim, U.Y.; Lee, B.; Pang, Y.; et al. Surface State-Mediated Charge Transfer of Cs2SnI6 and Its Application in Dye-Sensitized Solar Cells. Adv. Energy Mater. 2019, 9, 1803243. [Google Scholar] [CrossRef]
  141. Wang, A.; Yan, X.; Zhang, M.; Sun, S.; Yang, M.; Shen, W.; Pan, X.; Wang, P.; Deng, Z. Controlled Synthesis of Lead-Free and Stable Perovskite Derivative Cs2SnI6 Nanocrystals via a Facile Hot-Injection Process. Chem. Mater. 2016, 28, 8132–8140. [Google Scholar] [CrossRef]
  142. Pan, A.; He, B.; Fan, X.; Liu, Z.; Urban, J.J.; Alivisatos, A.P.; He, L.; Liu, Y. Insight into the Ligand-Mediated Synthesis of Colloidal CsPbBr3 Perovskite Nanocrystals: The Role of Organic Acid, Base, and Cesium Precursors. ACS Nano 2016, 10, 7943–7954. [Google Scholar] [CrossRef] [PubMed]
  143. Haydous, F.; Gardner, J.M.; Cappel, U.B. The Impact of Ligands on the Synthesis and Application of Metal Halide Perovskite Nanocrystals. J. Mater. Chem. A 2021, 9, 23419–23443. [Google Scholar] [CrossRef]
  144. Xu, Y.; Li, S.; Zhang, Z.; Hu, Y.; Yuan, L.; Chen, W.; Chen, Z.; Patterson, R.; Huang, S. Ligand-Mediated Synthesis of Colloidal Cs2SnI6 Three-Dimensional Nanocrystals and Two-Dimensional Nanoplatelets. Nanotechnology 2019, 30, 295601. [Google Scholar] [CrossRef] [PubMed]
  145. Muljarov, E.A.; Tikhodeev, S.G.; Gippius, N.A.; Ishihara, T. Excitons in Self-Organized Semiconductor/Insulator Superlattices: PbI-Based Perovskite Compounds. Phys. Rev. B 1995, 51, 14370–14378. [Google Scholar] [CrossRef] [PubMed]
  146. Goto, T.; Makino, H.; Yao, T.; Chia, C.H.; Makino, T.; Segawa, Y.; Mousdis, G.A.; Papavassiliou, G.C. Localization of Triplet Excitons and Biexcitons in the Two-Dimensional Semiconductor (CH3C6H4CH2NH3)2PbBr4. Phys. Rev. B 2006, 73, 115206. [Google Scholar] [CrossRef]
  147. Jing, Y.; Liu, Y.; Zhao, J.; Xia, Z. Sb3+ Doping-Induced Triplet Self-Trapped Excitons Emission in Lead-Free Cs2SnCl6 Nanocrystals. J. Phys. Chem. Lett. 2019, 10, 7439–7444. [Google Scholar] [CrossRef]
  148. Vargas, B.; Ramos, E.; Pérez-Gutiérrez, E.; Alonso, J.C.; Solis-Ibarra, D. A Direct Bandgap Copper-Antimony Halide Perovskite. J. Am. Chem. Soc. 2017, 139, 9116–9119. [Google Scholar] [CrossRef]
  149. Vargas, B.; Torres-Cadena, R.; Rodríguez-Hernández, J.; Gembicky, M.; Xie, H.; Jiménez-Mier, J.; Liu, Y.S.; Menéndez-Proupin, E.; Dunbar, K.R.; Lopez, N.; et al. Optical, Electronic, and Magnetic Engineering of «111» Layered Halide Perovskites. Chem. Mater. 2018, 30, 5315–5321. [Google Scholar] [CrossRef]
  150. Singhal, N.; Chakraborty, R.; Ghosh, P.; Nag, A. Low-Bandgap Cs4CuSb2Cl12 Layered Double Perovskite: Synthesis, Reversible Thermal Changes, and Magnetic Interaction. Chem. Asian J. 2018, 13, 2085–2092. [Google Scholar] [CrossRef]
  151. Vargas, B.; Torres-Cadena, R.; Reyes-Castillo, D.T.; Rodríguez-Hernández, J.; Gembicky, M.; Menéndez-Proupin, E.; Solis-Ibarra, D. Chemical Diversity in Lead-Free, Layered Double Perovskites: A Combined Experimental and Computational Approach. Chem. Mater. 2020, 32, 424–429. [Google Scholar] [CrossRef]
  152. Yang, H.; Shi, W.; Cai, T.; Hills-Kimball, K.; Liu, Z.; Dube, L.; Chen, O. Synthesis of Lead-Free Cs4(Cd1−xMnx)Bi2Cl12 (0 ≤ x ≤ 1) Layered Double Perovskite Nanocrystals with Controlled Mn-Mn Coupling Interaction. Nanoscale 2020, 12, 23191–23199. [Google Scholar] [CrossRef] [PubMed]
  153. Peng, C.; Wei, Q.; Chen, L.; Zeng, R.; Zhang, Q.; Hu, Q.; Zou, B. Efficient Energy Transfer in Cs4MnxCd1−xSb2Cl12 layered Perovskites and Anomalously Responsive Photodetectors. J. Mater. Chem. C 2021, 9, 15522–15529. [Google Scholar] [CrossRef]
  154. Cai, T.; Shi, W.; Gosztola, D.J.; Kobbekaduwa, K.; Yang, H.; Jin, N.; Nagaoka, Y.; Dube, L.; Schneider, J.; Hwang, S.; et al. Colloidal Synthesis and Charge Carrier Dynamics of Cs4Cd1−xCuxSb2Cl12 (0 ≤ x ≤ 1) Layered Double Perovskite Nanocrystals. Matter 2021, 4, 2936–2952. [Google Scholar] [CrossRef]
  155. Sartori, E.; Campolucci, M.; Baranov, D.; Zeng, M.; Toso, S.; Divitini, G.; Ferretti, M.; Hens, Z.; Manna, L.; Locardi, F. Red-Emissive Nanocrystals of Cs4MnxCd1−xSb2Cl12 layered Perovskites. Nanoscale 2022, 14, 305–311. [Google Scholar] [CrossRef]
  156. Sawahreh, A.; Binyamin, T.; Jiang, J.; Millo, O.; Goldberg, O.; Azulay, D.; Pachter, R.; Etgar, L. Electrical and Chemical Properties of Vacancy-Ordered Lead Free Layered Double Perovskite Nanoparticles. Nanoscale 2022, 14, 3487–3495. [Google Scholar] [CrossRef] [PubMed]
  157. Bhardwaj, A.; Kundu, K.; Sasmal, R.; Acharyya, P.; Pradhan, J.; Kalita, S.; Agasti, S.S.; Biswas, K. 2D Nanosheets of Layered Double Perovskites: Synthesis, Photostable Bright Orange Emission and Photoluminescence Blinking. Chem. Sci. 2023, 14, 7161–7169. [Google Scholar] [CrossRef] [PubMed]
  158. Richter, J.M.; Abdi-Jalebi, M.; Sadhanala, A.; Tabachnyk, M.; Rivett, J.P.H.; Pazos-Outón, L.M.; Gödel, K.C.; Price, M.; Deschler, F.; Friend, R.H. Enhancing Photoluminescence Yields in Lead Halide Perovskites by Photon Recycling and Light Out-Coupling. Nat. Commun. 2016, 7, 13941. [Google Scholar] [CrossRef]
  159. Klimov, V.I.; Mikhailovsky, A.A.; McBranch, D.W.; Leatherdale, C.A.; Bawendi, M.G. Quantization of Multiparticle Auger Rates in Semiconductor Quantum Dots. Science 2000, 287, 1011–1013. [Google Scholar] [CrossRef]
  160. Milot, R.L.; Sutton, R.J.; Eperon, G.E.; Haghighirad, A.A.; Martinez Hardigree, J.; Miranda, L.; Snaith, H.J.; Johnston, M.B.; Herz, L.M. Charge-Carrier Dynamics in 2D Hybrid Metal–Halide Perovskites. Nano Lett. 2016, 16, 7001–7007. [Google Scholar] [CrossRef]
  161. Zou, W.; Li, R.; Zhang, S.; Liu, Y.; Wang, N.; Cao, Y.; Miao, Y.; Xu, M.; Guo, Q.; Di, D.; et al. Minimising Efficiency Roll-off in High-Brightness Perovskite Light-Emitting Diodes. Nat. Commun. 2018, 9, 608. [Google Scholar] [CrossRef] [PubMed]
  162. Suyver, J.F.; Wuister, S.F.; Kelly, J.J.; Meijerink, A. Luminescence of Nanocrystalline ZnSe:Mn2+. Phys. Chem. Chem. Phys. 2000, 2, 5445–5448. [Google Scholar] [CrossRef]
  163. 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]
  164. Liu, Y.; Jing, Y.; Zhao, J.; Liu, Q.; Xia, Z. Design Optimization of Lead-Free Perovskite Cs2AgInCl6:Bi Nanocrystals with 11.4% Photoluminescence Quantum Yield. Chem. Mater. 2019, 31, 3333–3339. [Google Scholar] [CrossRef]
  165. Zhang, H.; Zhu, L.; Cheng, J.; Chen, L.; Liu, C.; Yuan, S. Photoluminescence Characteristics of Sn2+ and Ce3+-Doped Cs2SnCl6 Double-Perovskite Crystals. Materials 2019, 12, 1501. [Google Scholar] [CrossRef] [PubMed]
  166. Li, J.; Tan, Z.; Hu, M.; Chen, C.; Luo, J.; Li, S.; Gao, L.; Xiao, Z.; Niu, G.; Tang, J. Antimony Doped Cs2SnCl6 with Bright and Stable Emission. Front. Optoelectron. 2019, 12, 352–364. [Google Scholar] [CrossRef]
  167. Zhou, J.; Rong, X.; Zhang, P.; Molokeev, M.S.; Wei, P.; Liu, Q.; Zhang, X.; Xia, Z. Manipulation of Bi3+/In3+ Transmutation and Mn2+-Doping Effect on the Structure and Optical Properties of Double Perovskite Cs2NaBi1−xInxCl6. Adv. Opt. Mater. 2019, 7, 1801435. [Google Scholar] [CrossRef]
  168. Lin, T.W.; Su, C.; Lin, C.C. Phase Transition and Energy Transfer of Lead-Free Cs2SnCl6 Perovskite Nanocrystals by Controlling the Precursors and Doping Manganese Ions. J. Inf. Disp. 2019, 20, 209–216. [Google Scholar] [CrossRef]
  169. Noculak, A.; Noculak, A.; Morad, V.; Morad, V.; McCall, K.M.; Yakunin, S.; Shynkarenko, Y.; Yakunin, S.; Shynkarenko, Y.; Wörle, M.; et al. Bright Blue and Green Luminescence of Sb(III) in Double Perovskite Cs2MInCl6 (M = Na, K) Matrices. Chem. Mater. 2020, 32, 5118–5124. [Google Scholar] [CrossRef]
  170. Tan, Z.; Chu, Y.; Chen, J.; Li, J.; Ji, G.; Niu, G.; Gao, L.; Xiao, Z.; Tang, J. Lead-Free Perovskite Variant Solid Solutions Cs2Sn1–xTexCl6: Bright Luminescence and High Anti-Water Stability. Adv. Mater. 2020, 32, e2002443. [Google Scholar] [CrossRef]
  171. Koyanagi, T.; Kapil, G.; Ogomi, Y.; Yoshino, K.; Shen, Q.; Toyoda, T.; Murakami, T.N.; Segawa, H.; Hayase, S. Hot-Injection and Ultrasonic Irradiation Syntheses of Cs2SnI6 Quantum Dot Using Sn Long-Chain Amino-Complex. J. Nanoparticle Res. 2020, 22, 69. [Google Scholar] [CrossRef]
  172. Holzapfel, N.P.; Majher, J.D.; Strom, T.A.; Moore, C.E.; Woodward, P.M. Cs4Cd1−xMnxBi2Cl12- A Vacancy-Ordered Halide Perovskite Phosphor with High-Efficiency Orange-Red Emission. Chem. Mater. 2020, 32, 3510–3516. [Google Scholar] [CrossRef]
  173. Vargas, B.; Reyes-Castillo, D.T.; Coutino-Gonzalez, E.; Sánchez-Aké, C.; Ramos, C.; Falcony, C.; Solis-Ibarra, D. Enhanced Luminescence and Mechanistic Studies on Layered Double-Perovskite Phosphors: Cs4Cd1−xMnxBi2Cl12. Chem. Mater. 2020, 32, 9307–9315. [Google Scholar] [CrossRef]
  174. Zheng, W.; Sun, R.; Liu, Y.; Wang, X.; Liu, N.; Ji, Y.; Wang, L.; Liu, H.; Zhang, Y. Excitation Management of Lead-Free Perovskite Nanocrystals through Doping. ACS Appl. Mater. Interfaces 2021, 13, 6404–6410. [Google Scholar] [CrossRef] [PubMed]
  175. Levy, S.; Khalfin, S.; Pavlopoulos, N.G.; Kauffmann, Y.; Atiya, G.; Shaek, S.; Dror, S.; Shechter, R.; Bekenstein, Y. The Role Silver Nanoparticles Plays in Silver-Based Double-Perovskite Nanocrystals. Chem. Mater. 2021, 33, 2370–2377. [Google Scholar] [CrossRef] [PubMed]
  176. Ahmad, R.; Zdražil, L.; Kalytchuk, S.; Naldoni, A.; Rogach, A.L.; Schmuki, P.; Zboril, R.; Kment, Š. Uncovering the Role of Trioctylphosphine on Colloidal and Emission Stability of Sb-Alloyed Cs2NaInCl6 Double Perovskite Nanocrystals. ACS Appl. Mater. Interfaces 2021, 13, 47845–47859. [Google Scholar] [CrossRef] [PubMed]
  177. Shen, J.; Zhu, W.; Lian, Z.; Lin, A.; Shi, S.F.; Yang, K.; Li, M.; Zhao, D.; Sun, Y.Y.; Lian, J. Metal Ion-Incorporated Lead-Free Perovskites toward Broadband Photodetectors. ACS Appl. Electron. Mater. 2023. [Google Scholar] [CrossRef]
  178. Liu, M.; Matta, S.K.; Ali-Löytty, H.; Matuhina, A.; Grandhi, G.K.; Lahtonen, K.; Russo, S.P.; Vivo, P. Moisture-Assisted near-UV Emission Enhancement of Lead-Free Cs4CuIn2Cl12 Double Perovskite Nanocrystals. Nano Lett. 2022, 22, 311–318. [Google Scholar] [CrossRef]
  179. Cui, E.; Yuan, X.; Tang, L.; Yang, L.; Yang, X.; Liao, X.; Tang, J.; Zhao, Y.; Sun, W.; Liu, K.; et al. Eu3+, Tb3+ Doping Induced Tunable Luminescence of Cs2AgInCl6 Double Perovskite Nanocrystals and Its Mechanism. Appl. Surf. Sci. 2023, 609, 155472. [Google Scholar] [CrossRef]
  180. Dolzhnikov, D.S.; Wang, C.; Xu, Y.; Kanatzidis, M.G.; Weiss, E.A. Ligand-Free, Quantum-Confined Cs2SnI6 Perovskite Nanocrystals. Chem. Mater. 2017, 29, 7901–7907. [Google Scholar] [CrossRef]
  181. Tang, G.; Xiao, Z.; Hosono, H.; Kamiya, T.; Fang, D.; Hong, J. Layered Halide Double Perovskites Cs3+nM(II)nSb2X9+3n (M = Sn, Ge) for Photovoltaic Applications. J. Phys. Chem. Lett. 2018, 9, 43–48. [Google Scholar] [CrossRef] [PubMed]
  182. Tan, Z.; Li, J.; Zhang, C.; Li, Z.; Hu, Q.; Xiao, Z.; Kamiya, T.; Hosono, H.; Niu, G.; Lifshitz, E.; et al. Highly Efficient Blue-Emitting Bi-Doped Cs2SnCl6 Perovskite Variant: Photoluminescence Induced by Impurity Doping. Adv. Funct. Mater. 2018, 28, 1801131. [Google Scholar] [CrossRef]
  183. Mu, H.; Hu, F.; Wang, R.; Jia, J.; Xiao, S. Effects of In-Situ Annealing on the Electroluminescence Performance of the Sn-Based Perovskite Light-Emitting Diodes Prepared by Thermal Evaporation. J. Lumin. 2020, 226, 117493. [Google Scholar] [CrossRef]
  184. Mahesh, K.P.O.; Chang, C.-Y.; Hong, W.-L.; Wen, T.-H.; Lo, P.-H.; Chiu, H.-Z.; Hsu, C.-L.; Horng, S.-F.; Chao, Y.-C. Lead-Free Cesium Tin Halide Nanocrystals for Light-Emitting Diodes and Color down Conversion. RSC Adv. 2020, 10, 37161–37167. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, X.; Wang, C.; Zhang, Y.; Zhang, X.; Wang, S.; Lu, M.; Cui, H.; Kershaw, S.V.; Yu, W.W.; Rogach, A.L. Bright Orange Electroluminescence from Lead-Free Two-Dimensional Perovskites. ACS Energy Lett. 2019, 4, 242–248. [Google Scholar] [CrossRef]
  186. Ma, Z.; Shi, Z.; Yang, D.; Zhang, F.; Li, S.; Wang, L.; Wu, D.; Zhang, Y.; Na, G.; Zhang, L.; et al. Electrically-Driven Violet Light-Emitting Devices Based on Highly Stable Lead-Free Perovskite Cs3Sb2Br9 Quantum Dots. ACS Energy Lett. 2020, 5, 385–394. [Google Scholar] [CrossRef]
  187. Wang, L.; Shi, Z.; Ma, Z.; Yang, D.; Zhang, F.; Ji, X.; Wang, M.; Chen, X.; Na, G.; Chen, S.; et al. Colloidal Synthesis of Ternary Copper Halide Nanocrystals for High-Efficiency Deep-Blue Light-Emitting Diodes with a Half-Lifetime above 100 H. Nano Lett. 2020, 20, 3568–3576. [Google Scholar] [CrossRef]
  188. Zhang, Y.; Zhang, Z.; Yu, W.; He, Y.; Chen, Z.; Xiao, L.; Shi, J.; Guo, X.; Wang, S.; Qu, B. Lead-free Double Perovskite Cs2AgIn0.9Bi0.1Cl6 Quantum Dots for White Light-Emitting Diodes. Adv. Sci. 2022, 9, 2102895. [Google Scholar] [CrossRef]
  189. Liu, Z.; Sun, Y.; Cai, T.; Yang, H.; Zhao, J.; Yin, T.; Hao, C.; Chen, M.; Shi, W.; Li, X.; et al. Two-Dimensional Cs2AgInxBi1−xCl6 Alloyed Double Perovskite Nanoplatelets for Solution-Processed Light-Emitting Diodes. Adv. Mater. 2023, 35, 2211235. [Google Scholar] [CrossRef]
  190. Dohner, E.R.; Hoke, E.T.; Karunadasa, H.I. Self-Assembly of Broadband White-Light Emitters. J. Am. Chem. Soc. 2014, 136, 1718–1721. [Google Scholar] [CrossRef]
  191. Dohner, E.R.; Jaffe, A.; Bradshaw, L.R.; Karunadasa, H.I. Intrinsic White-Light Emission from Layered Hybrid Perovskites. J. Am. Chem. Soc. 2014, 136, 13154–13157. [Google Scholar] [CrossRef] [PubMed]
  192. Even, J.; Pedesseau, L.; Katan, C. Understanding Quantum Confinement of Charge Carriers in Layered 2D Hybrid Perovskites. ChemPhysChem 2014, 15, 3733–3741. [Google Scholar] [CrossRef] [PubMed]
  193. 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]
  194. Jung, M.-C.; Raga, S.R.; Qi, Y. Properties and Solar Cell Applications of Pb-Free Perovskite Films Formed by Vapor Deposition. RSC Adv. 2016, 6, 2819–2825. [Google Scholar] [CrossRef]
  195. Liao, Y.; Liu, H.; Zhou, W.; Yang, D.; Shang, Y.; Shi, Z.; Li, B.; Jiang, X.; Zhang, L.; Quan, L.N.; et al. Highly Oriented Low-Dimensional Tin Halide Perovskites with Enhanced Stability and Photovoltaic Performance. J. Am. Chem. Soc. 2017, 139, 6693–6699. [Google Scholar] [CrossRef] [PubMed]
  196. Wang, S.; Li, X.; Wu, J.; Wen, W.; Qi, Y. Fabrication of Efficient Metal Halide Perovskite Solar Cells by Vacuum Thermal Evaporation: A Progress Review. Curr. Opin. Electrochem. 2018, 11, 130–140. [Google Scholar] [CrossRef]
  197. Moghe, D.; Wang, L.; Traverse, C.J.; Redoute, A.; Sponseller, M.; Brown, P.R.; Bulović, V.; Lunt, R.R. All Vapor-Deposited Lead-Free Doped CsSnBr3 Planar Solar Cells. Nano Energy 2016, 28, 469–474. [Google Scholar] [CrossRef]
  198. Liu, M.; Johnston, M.B.; Snaith, H.J. Efficient Planar Heterojunction Perovskite Solar Cells by Vapour Deposition. Nature 2013, 501, 395–398. [Google Scholar] [CrossRef]
  199. Yuan, F.; Xi, J.; Dong, H.; Xi, K.; Zhang, W.; Ran, C.; Jiao, B.; Hou, X.; Jen, A.K.Y.; Wu, Z. All-Inorganic Hetero-Structured Cesium Tin Halide Perovskite Light-Emitting Diodes with Current Density Over 900 A Cm−2 and Its Amplified Spontaneous Emission Behaviors. Phys. Status Solidi-Rapid Res. Lett. 2018, 12, e2207454. [Google Scholar] [CrossRef]
  200. Woo, J.Y.; Park, M.; Jeong, S.; Kim, Y.; Kim, B.; Lee, T.; Han, T. Advances in Solution-Processed OLEDs and Their Prospects for Use in Displays. Adv. Mater. 2023. early view. [Google Scholar] [CrossRef]
Figure 1. Periodic table of the elements that have the potential to constitute perovskite compositions (A-site: green; B-site: sky-blue; and X-site: orange).
Figure 1. Periodic table of the elements that have the potential to constitute perovskite compositions (A-site: green; B-site: sky-blue; and X-site: orange).
Materials 16 06317 g001
Figure 8. (a) Energy diagram of the PeLED with CsSnBr3 PeNCs; (b) current efficiency (CE) and external quantum efficiency (EQE) versus voltage curves of the CsSnBr3 NC-based PeLED at various in-situ annealing temperatures; (a,b) Reprinted with permission from Ref. [183]. Copyright 2020, Elsevier. (c) Cross-sectional scanning electron microscopy (SEM) image of ITO/ZnO nanoparticles (NP)/PEI/Cs3Sb2Br9 NCs/TCTA/MoO3/Al; (d) EQE-voltage curve of Cs3Sb2Br9 NC-based PeLED; (c,d) reprinted with permission from Ref. [186]. Copyright 2020, American Chemical Society. (e) Energy band of the Cs3Cu2I5 NC-based LED; (f) EQE-voltage curve of Cs3Cu2I5 NCs PeLED; (e,f) reprinted with permission from Ref. [187]. Copyright 2020, American Chemical Society. (g) Energy band diagram of Cs2AgIn0.9Bi0.1Cl6 NC-based PeLED; (h) current efficiency-voltage curve (black line) and EQE-Voltage curve of Cs2AgIn0.9Bi0.1Cl6 NC-based PeLED; (g,h) reprinted with permission from Ref. [188]. Copyright 2021, Wiley-VCH.
Figure 8. (a) Energy diagram of the PeLED with CsSnBr3 PeNCs; (b) current efficiency (CE) and external quantum efficiency (EQE) versus voltage curves of the CsSnBr3 NC-based PeLED at various in-situ annealing temperatures; (a,b) Reprinted with permission from Ref. [183]. Copyright 2020, Elsevier. (c) Cross-sectional scanning electron microscopy (SEM) image of ITO/ZnO nanoparticles (NP)/PEI/Cs3Sb2Br9 NCs/TCTA/MoO3/Al; (d) EQE-voltage curve of Cs3Sb2Br9 NC-based PeLED; (c,d) reprinted with permission from Ref. [186]. Copyright 2020, American Chemical Society. (e) Energy band of the Cs3Cu2I5 NC-based LED; (f) EQE-voltage curve of Cs3Cu2I5 NCs PeLED; (e,f) reprinted with permission from Ref. [187]. Copyright 2020, American Chemical Society. (g) Energy band diagram of Cs2AgIn0.9Bi0.1Cl6 NC-based PeLED; (h) current efficiency-voltage curve (black line) and EQE-Voltage curve of Cs2AgIn0.9Bi0.1Cl6 NC-based PeLED; (g,h) reprinted with permission from Ref. [188]. Copyright 2021, Wiley-VCH.
Materials 16 06317 g008
Figure 9. Illustration of techniques for enhancing the material properties of PeNCs.
Figure 9. Illustration of techniques for enhancing the material properties of PeNCs.
Materials 16 06317 g009
Table 1. Optical and photophysical properties of PeNCs.
Table 1. Optical and photophysical properties of PeNCs.
PerovskiteAbsorption Peak (nm)PL Peak (nm)Stokes Shift (nm)PLQY (%)PL Lifetime (ns)Reference
CsPbCl3-408-657.63[24]
CsPbBr3-512-9212.52[24]
CsPbI3-691-5820.99[24]
FAPbBr3-54528-18.7[25]
CsPb(Br/I)3 NC film-~650-1526.2[32]
CsSnBr3 nanocages655685302.16.52[59]
PFOA-CsSnBr3 nanocages-683-1.80.25[59]
OA/OAm-CsSnCl3~300~440~140-7.33[61]
Gelatin-CsSnCl3303442139-8.84[61]
Cs3Bi2Br9439460214.5~2.7[70]
Cl-MA3Bi2Br93884223454.12.17[95]
Cs3Sb2Br937541035464.285[97]
K3SbCl6~32844012022.3535.2[99]
K3SbCl6:Mn2+318~32060028037.23.2 × 106[99]
CsCu2I3321561240110.1 × 103[107]
Cs3Cu2I530044015872.22.8 × 103[122]
Cs3Cu2I5:InI329544014596.61491[102]
Cs2CuCl4~3003888851.8-[106]
Cs2CuBr43603933337.5-[106]
Cs2AgBiBr6500625125-7.5[128]
Cs2AgBiCl6~360395~356.7-[132]
Cs2SnCl63174381214.3711[147]
Cs4CdBi2Cl12331~602~271<0.1-[152]
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

Kim, D.-Y.; Jung, J.-G.; Lee, Y.-J.; Park, M.-H. Lead-Free Halide Perovskite Nanocrystals for Light-Emitting Diodes. Materials 2023, 16, 6317. https://doi.org/10.3390/ma16186317

AMA Style

Kim D-Y, Jung J-G, Lee Y-J, Park M-H. Lead-Free Halide Perovskite Nanocrystals for Light-Emitting Diodes. Materials. 2023; 16(18):6317. https://doi.org/10.3390/ma16186317

Chicago/Turabian Style

Kim, Do-Young, Jae-Geun Jung, Ye-Ji Lee, and Min-Ho Park. 2023. "Lead-Free Halide Perovskite Nanocrystals for Light-Emitting Diodes" Materials 16, no. 18: 6317. https://doi.org/10.3390/ma16186317

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