Next Article in Journal
Quantum Behaviour of Mg and Mg-Al-Zn Microstructure
Next Article in Special Issue
Ba0.9A0.1MnO3 (A = Ce, La, Mg) Perovskite-Type Mixed Oxides: Effect of Partial Substitution of Ba on the Catalytic Performance for the Oxidation of CO in Simulated Automobile Exhaust Conditions
Previous Article in Journal
Research Progress in Capping Diamond Growth on GaN HEMT: A Review
Previous Article in Special Issue
Evaluation of the Minority-Carrier Lifetime of IMM3J Solar Cells under Proton Irradiation Based on Electroluminescence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

All Inorganic Lead-Free Zero-Dimensional Metal Halide Luminescent Materials and Applications

Department of Chemistry, National Taiwan University, Taipei 106, Taiwan
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(3), 499; https://doi.org/10.3390/cryst13030499
Submission received: 15 February 2023 / Revised: 4 March 2023 / Accepted: 10 March 2023 / Published: 14 March 2023
(This article belongs to the Special Issue Perovskites – New and Old Materials)

Abstract

:
Recently, zero-dimensional luminescent material has attracted researchers because of its optical properties, which is a possible candidate to replace lead halide perovskite. This review focused on the recent development of tetrahedrally and octahedrally coordinated inorganic halide semiconductor luminescent materials. We discuss the synthesis methods and crystal structures of these materials in this review. The materials are categorized based on the valence of central metal cations (monovalent, divalent, and trivalent). Finally, we have summarized the applications of these luminescent materials, such as light-emitting diodes, ultrafast switching memories, photodetectors, and scintillators. This review article provides an overview of recent progress on zero-dimensional materials and their applications for further development in the future.

1. Introduction

Lead halide (CsPbX3) (X = Cl, Br, and I) perovskite draws the attention of researchers because of its attractive properties for optoelectronics applications such as solar cells, light-emitting diodes, photodetectors, and scintillators [1,2,3,4,5]. Although these materials are used for various applications, the presence of lead (Pb) is a source of concern for researchers because of its toxicity. Several countries restrict the use of heavy metals for electronic devices, including lead. These concerns encourage researchers to move towards lead-free technology for these optoelectronics. Tin ion (Sn2+) is an ideal element to substitute for Pb2+ because it is in the same group in the periodic table. However, Sn2+ is unstable in the atmosphere and oxidizes into Sn4+, which makes an unstable perovskite structure [6]. It was initially observed that double perovskites are replaceable but have broadband white light emission with UV light excitation [7,8]. Several research groups have focused on developing zero-dimensional inorganic materials (AhBkXl), which are ideal to replace lead halide perovskite [9]. In AhBkXl-type compounds, A is the monovalent cations (K+, Cs+, and Rb+); B can be in monovalent, divalent, or trivalent cations; X can be halides (Cl, Br, and I); and (1 < h, k, l < 9). The zero-dimensional structure has isolated tetrahedral [BX4] (Figure 1a) or octahedral [BX6] (Figure 1b) coordination units [9,10]. The dimension relates to the arrangement of the inorganic unit at the molecular level rather than the particle size [9]. These structures attract researchers because of the isolated polyhedral inorganic units and extraordinary optical properties [11].
Only a few review articles have been published on zero-dimensional perovskite. Wang et al. [10] discuss the luminescence mechanism in zero-dimensional lead halide Cs4PbBr6. Zhou et al. [12] summarized the work on organic–inorganic zero-dimensional material and their application for optoelectronics. Li et al. [9] discussed the organic–inorganic hybrid as well as all inorganic zero-dimensional materials with different optoelectronics applications. Recently several research articles have been published on all inorganic zero-dimensional materials with synthesizing techniques and applications, which is the focus of this review.
In this review article, we have summarized recent development in tetrahedrally and octahedrally coordinated with all inorganic zero-dimensional material and their luminescent properties. We have deliberated the synthesis method, crystal structure, and luminescence properties. Finally, we have discussed the application of these materials for optoelectronics.

Luminescence Mechanism of Zero-Dimensional Metal Halide

Figure 2a illustrates the mechanism of light absorption and relaxation to emission lights Figure 2b–f. The photoluminescence mechanisms are categorized into intrinsic and extrinsic properties for solids. Intrinsic photoluminescence has band-to-band emission and exciton emission. In the band-to-band emission, the electron from the conduction band recombines with the valence band hole, shown in Figure 2b [13,14]. In exciton luminescence, photogenerated electron-holes attract each other, and the resulting recombination emits photons, as shown in Figure 2c [13]. On the other hand, for extrinsic luminescence, the impurities, doping ions, and defects make levels where electrons from the conduction band transfer and then relax to the ground level via radiative recombination, as shown in Figure 2d,e [14]. Additionally, in intrinsic and extrinsic photoluminescence, several alkali halides, metal halides, organic–inorganic halide crystals, and low-dimensional materials display a self-trap exciton (STE) emission [15]. The STE emission material has a large Stokes shift and broadband emission [7]. In this material, upon light illumination, the strong electron–phonon coupling between the excited electron and crystal lattice results in strong carrier–photon interactions, as shown in Figure 2f. In emission mechanisms, photogenerated electrons are trapped at different STE states to lower the energy, and holes also increase the energy [16]. Thus, the emitted energy lowers the excited energy in STE-derived emission [17]. These types of luminescence have been observed in halide-based soft lattices [18]. Most zero-dimensional materials also have STE emissions, which we discuss in the next section.

2. Zero-Dimensional Metal Halide Luminescent Materials

We have discussed in the introduction the general formula of compound ABX. This section discusses the coordination environment of B with monovalent, divalent, and trivalent coordination.

2.1. Mono-Valent Coordinated Zero-Dimensional Materials

2.1.1. Cu-Based

Copper-based CsCu2I3, Cs3Cu2I5, and Cs3Cu2Br5−xIx are studied because of the high quantum yield for lighting and scintillation [19,20,21,22]. Liu et al. [19] used thermal evaporation to deposit Cs3Cu2I5 for deep blue light emission at 440 nm. This method is suitable for making uniform and highly crystalline thin films. Cs3Cu2I5 has an orthorhombic phase, with Cu+ having two tetragonal and trigonal sites. The thermal evaporation was carried out using heating CsI and CuI above the melting point and deposited on a quartz/ITO substrate. The uniform thin film has a photoluminescence quantum yield of up to 58% [19]. In the past, several researchers reported using solution process synthesis nanocrystals first, followed by the deposition of these nanocrystals on the substrate. Wang et al. [20] synthesized Cs3Cu2I5 nanocrystals by hot injection for blue light-emitting application. The crystal structure of Cs3Cu2I5 has an orthorhombic phase, the same as thermal evaporated fabricated thin films. The photoluminescence quantum yield of this nanocrystal reached up to 87%, which is higher than the thin film and further applied for the blue LED [20]. Along with this, a solid-state reaction was used to synthesize Cs3Cu2I5 bulk crystal with a quantum efficiency of up to 80% for white light emission on a 310 nm UV chip with green and red-emissive phosphors [23]. Roccanova et al. [21] used a solid solution to synthesize Cs3Cu2Br5−xIx and achieved a PLQY of up to 98.7% for Cs3Cu2I5. So far, all these studies have found that the bulk Cs3Cu2I5 has high photoluminescence quantum yield compared to nanocrystals. Cs3Cu2I5 is well synthesized for single crystals, nanocrystals, and microcrystals and used for blue light-emitting diodes and χ-ray imaging applications, which we discuss in the application sections.

2.1.2. Ag-Based

Ag-based compounds are reported with tunable emissions such as Cs2AgCl3, Cs2AgBr3, CsAgCl2, CsAgBr2, and Rb2AgBr3 [24,25]. Zhang et al. [24] used the ball milling method to synthesize bimetallic compound with tunable emission Cs-Ag-X(x = Cl and Br). Cs2AgX3 has an orthorhombic crystal system with a Pnma space group. These materials are excited with high energy photons and give emissions for Cs2AgCl3, Cs2AgBr3, CsAgCl2, and CsAgBr2 at 397, 524, 620, and 820 nm, respectively. Resultantly these materials are unsuitable for WLED application and need further research to improve the issue. Defect-free single crystals with this system can be used for X-ray scintillators with enhancing quantum yield. Toward the application of scintillation, Zhang et al. [25] synthesized a single crystal of Rb2AgBr3. Rb2AgBr3 has broadband emissions centered at 468 nm. This system achieved a scintillation yield of up to 25,600 photons MeV−1 for χ-ray imaging. These results suggest the possible tunable emission with Ag-based compounds is attributed to a change in crystal structure and halides. We have added comparative photo-physical properties of 0D metal halide with monovalent units in Table 1.

2.2. Di-Valent Coordinated Zero-Dimensional Materials

2.2.1. Sn-Based

Tin ion (Sn2+) is an ideal element to replace the lead ion (Pb2+) in a three-dimensional APbX3 perovskite system because they belong to the same group in the periodic table. However, Sn2+ is unstable and oxidizes into Sn4+ in moisture and water. Although several research groups have tried to make Sn-based three-dimensional perovskites, they are still not sufficiently stable for industrial applications. Recently several groups reported zero-dimensional Sn-based Cs4SnX6 (x = Cl, Br, and I) systems that can stabilize structures with Sn2+ for lighting applications. Both nanocrystal and bulk crystals are synthesized for Cs4SnX6, which gives emission via the STE mechanism [26]. Cs4SnBr6 has a trigonal crystal system with the space group of R-3c, where [SnBr6]4– is in isolated octahedral coordination [27]. Nanocrystals of Cs4SnX6 are synthesized by the hot-injection method in oleic acid and oleylamine in octadecene solvents. In the hot-injection method, Cs2CO3 and SnBr2 precursors are dissolved in the nonpolar solvent 1-octadecene with ligands oleic acid and oleylamine at 120–150 °C in a vacuum separately. After that, Cs precursor is injected into Sn precursor in a nitrogen atmosphere at >160 °C. Further, the reaction is quenched by an ice or water bath to complete the formation of nanocrystals. Tan et al. [28] synthesized Cs4SnBr6 with partial and complete substitution with halide by the hot-injection method. The synthesized Cs4SnBr6 shows emission at 534 nm with an excitation wavelength of 365 nm. Partial substitution with iodide shifts emissions to 546 nm and complete substitution to 578 nm in Cs4SnBrxI6-x, as shown in Figure 3a. Moreover, with substitutions, the PLQY for Cs4SnBr6, Cs4SnBrxI6–x, and Cs4SnI6 are 21%, 6.6%, and 0.7%, respectively. These three systems have a large Stokes shift with high FWHM, confirming the STE-derived emission. Xu et al. [29] modified the hot-injection method to improve luminescence properties, as shown in Figure 3b for Cs4SnBr6, Cs4SnBr3I3, and Cs4SnI6. The synthesized Cs4SnBr6 compound has an emission at 537 nm with a PLQY of 52%. It is the highest PLQY reported for the nanocrystals to date. However, the Cs4SnBr3I3 nanocrystals have emission at 601 nm and Cs4SnI6 at 578 nm. The possible reason for the unusual redshift of Cs4SnBr3I3 is mixed halogen, which increases the lattice deformation energy [29]. Synthesized Cs4SnI6 nanocrystals have a PLQY of 27%. In this article, the thermodynamic stability of zero-dimensional and three-dimensional Sn-based materials are also studied comparatively. Density functional theory simulation is used to calculate enthalpies of formation (∆fHm). Zero-dimensional Cs4SnBr6 and Cs4SnI6 have negative values of −1721, and −1496 kJ mol−1, respectively, and three-dimensional CsSnBr3 and CsSnI3 have −642 and −522 kJ mol−1, respectively. Therefore, zero-dimensional luminescent materials have more negative value than three-dimensional materials, making them more stable.
Several studies have been carried out to understand the bulk of zero-dimensional Cs4SnX6. The Kovalenko group used a solid-state synthesis method for Cs4SnBr6 and mixed it with cations (Rb+ and K+) and anions (I) [27]. In this method, CsBr and SnBr2 were mixed in a stoichiometric ratio and heated in a vacuum at 360 °C in the pellet. With tuning from the Br to I, the excitation and emission band redshift until the 50% substitution, as presented in Figure 3c. The emission band shifted from 540 to 620 nm for Cs4SnBr6 of Cs4SnBr3I3, respectively, although no emission was observed at room temperature with more than 50% of substitutions. These results are similar to the reported results for nanocrystals in Figure 3b. The substitution of monovalent cations (Rb+ and K+) at the Cs+ site blue shift the emission peak, as shown Figure 3d. Moreover, 25% substitution of Rb+ and K+ shift emission bands to 519 and 510 nm, respectively. They have studied the tunable emission in Cs4SnBr6 with different amounts of Rb+, K+, and I in UV light, as shown in Figure 3e. They obtained the highest quantum efficiency of 15% for Cs4SnBr6. Recently Zhang et al. [26] synthesized Cs4SnBr6 powder by the ball mill method, and further oxidation was reduced by partial substitution by F, as shown in Figure 3f. The PLQY was enhanced from 2.8% to 62.8% after a partial substitution of F, which is the highest PLQY achieved for the bulk compound. F substitution at the Br site breaks the lattice symmetry by introducing lattice distortion in the octahedral, a possible reason for enhancing emission. Moreover, this research finds the reduction in charge transfer between oxygen and Sn2+ by calculating the Bader atomic charge [26]. The Bader charge gives information about the charge transfer between O2 and Sn2+ [26]. Cs4SnBr6 has a Bader charge of 0.421e, and after F substitution, it changes to 0.299e, which accomplishes that charge transfer inhibition and resultantly suppresses oxidation.
Both nanocrystal and bulk crystals exhibit the possible emission mechanism of luminescence Cs4SnBr6 via STE in Figure 3g, and the emission energy can describe by the given equation for the mechanism.
E p l = E g E b E S T E E d
Epl is the photoluminescence energy, Eg is the bandgap, Eb binding energy, ESTE is self-trapping exciton energy, and Ed is lattice deformation energy.
All these studies regarding the Sn-based confirm that Cs4SnBr6 has green-emissive luminescent properties and can be a possible candidate for lighting applications.

2.2.2. Zn-Based

Zn-derived zero-dimensional luminescence materials are synthesized with different dopants. In Cs2ZnCl4, [ZnCl4]2− presents in the tetrahedral coordination with cation Cs+ in the void. Although Cs2ZnCl4 does not emit light, doping induces the emission via self-trapped exciton emission. Wang et al. [30] used the acid precipitation method to synthesize Sn2+-doped into Cs2ZnCl4. Sn2+ has an outer cell electronic configuration of ns2, which resultantly have the possible transition from 1P1 to 1S0 and 3P1 to 1S0. They observed the emission centered at 648 nm because of the transition from 3P1 to 1S0. The theoretical calculation found isolated tetrahedral [ZnCl4]2− possesses almost regular arrangement in Cs2ZnCl4. Sn-doped Cs2ZnCl4 has disphenoidal [SnCl4]2− units to capture the exciton for STE. Moreover, these systems have a binding energy of 15 meV, which is lower and insufficient for lighting applications. Although introducing organic cations increases the binding energy in a Zn-based system, we do not cover articles related to organic–inorganic hybrid materials in this review.
Zero-dimensional Cs3ZnX5 compounds are also developed in areas for lighting applications. Cs3ZnX5 (x = Cl and Br) has a tetragonal crystal system with a space group of I4/mcm where [ZnX4]2− is in tetrahedral coordination [31]. The hot-injection method is used to synthesize the nanocrystals Cs3ZnCl5 and Cs3ZnBr5. Cs3ZnCl5 emits 468 nm at the excitation wavelength of 275 nm, which shifts to 484 nm in Cs3ZnBr5. Recently Liu et al. [32] used doping of Sb3+ in Cs3ZnCl5, where Sb partially occupies the Zn positions. Doping of Sb3+ gives an emission from red to orange with increased dopant concentrations, as shown in Figure 4a. The compound has two excitation peaks because of the transition of 1S0 3P1 and 1S0 3P1 at 275 nm and 320 nm in Figure 4b. The emission after Sb3+ causes a broadband emission with large stock because of the STE at the emission center at 640 nm in Figure 4c. It is well-studied that the emission occurred via triplet self-trapped excitons in Sb3+-doped systems. The PLQY of the Sb3+-doped system achieved up to 34.18% at the excitation wavelength of 320 nm. However, this system can still be excitable by UV light and stability of this compound needs to be studied in detail. We have added comparative photo-physical properties of 0D metal halide with divalent units in Table 2.

2.2.3. Mn-Based

Manganese-derived luminescence material already attracts the researcher to green- and red-emissive materials Cs3MnBr5 and CsMnBr3. CsMnBr3 are in one-dimensional electronic configuration with octahedral coordination and Cs3MnBr5 in zero-dimensional electronic configurations with tetrahedral coordination. Several groups used different types of nanocrystals, single crystals, and bulk crystals for Cs3MnBr5 and CsMnBr3. Nanocrystals are synthesized by hot injection, where cesium acetate and manganese acetate dissolve in a nonpolar solvent with organic ligands. Organic bromide ligand precursors were further added for complete reaction at a high temperature in the inert atmosphere to complete the reaction. It is noted that all nanocrystals are stored in a nonpolar solvent with additional organic ligands to stabilize them. Prof. Manna’s group synthesized CsMnBr3, and Cs3MnBr5 nanocrystals doped with rare earth elements and found that CsMnBr3 is the ideal compound to use as a sensitizer for energy transfer in rare earth elements [33]. They tuned the temperature and stoichiometric ratio of Cs2CO3, manganese(II) acetate, and benzoyl bromide for the synthesizing of CsMnBr3 and Cs3MnBr5 separately. The quantum efficiency for NIR-1 is very low, and it needs to be enhanced. Recently Almutlaq et al. synthesized CsMnBr3 nanocrystals and achieved a PLQY of up to 56% [34]. Kong et al. used phase engineering to change the crystal structure with different emissions [35]. They used the hot-injection method with different ratios of precursors to synthesize Cs3MnBr5 and CsMnBr3 emissions at 520 and 660 nm, respectively, as shown in Figure 5a. However, they found that the CsMnBr3 can transform into Cs3MnBr5 after a reaction with isopropanol, and both the nanocrystals of CsMnBr3 and Cs3MnBr5 can absorb water and transform into blue-emissive Cs2MnBr4·2H2O at 440 nm. Although the PLQY of Cs2MnBr4·2H2O is low, it can further transform into the mixture of Cs3MnBr5 and CsMnBr3 after dehydration, as shown in Figure 5b,c. These recent studies found that the Mn-based material can be a possible candidate to replace lead halide perovskites for future optoelectronics applications. However, several studies needed to be carried out for industrial applications, such as the low PLQY of nanocrystals and the stability of these compounds.
Table 2. Properties of 0D metal halide with divalent units.
Table 2. Properties of 0D metal halide with divalent units.
FormulaMethodIsolated
Species
Em.
Wave. (nm)
FWHM
(nm)
PLQY
(%)
Ref.
Cs4SnBr6 NCHot injection[SnBr6]4−537 52[29]
Cs4SnI6 NCHot injection[SnI6]4−578 27[29]
Cs4SnBr6-SnF2Ball Milling[SnBr6]4−540 62.8[26]
Cs3ZnCl5Hot injection[ZnCl4]2−48479 [31]
Cs3ZnBr5Hot injection[ZnBr4]2−468767.98[31]
Cs3MnBr5 NCHot injection[MnBr4]2− 5204348[35]
Cs2MnBr4·2H2OHot injection/Humid environment[MnBr4(H2O)2]2−440 1.29[35]

2.3. Tri-Valent Coordinated Zero-Dimensional Materials

2.3.1. Bi-Based

Different kinds of Bi-derived inorganic luminescence materials are reported, such as zero-dimensional (Cs3BiX6) and two-dimensional (Cs3Bi2X9) materials [36,37]. In Cs3BiX6, [BiX6]3− are in isolated octahedral coordination, and in Cs3Bi2X9, the [BiX6]3− octahedral coordination is attached to each other’s to form layers. Cs3BiCl6 has a monoclinic phase with the C2/c space group [38]. In this work, they also find the decomposition of Cs3BiX6 to Cs3Bi2X9 at heating to 220 °C. Cs3BiCl6 has broadband with emission centered at 391 nm. Cs3BiCl6 can further transform into an anion exchange reaction to Cs3BiBr6; by iodide exchange, it transforms into Cs3Bi2I9. Along with it, zero-dimensional Cs3BiX6 can transform into Cs3Bi2X9 after the insertion of BiX3. So far, Bi-based zero-dimensional material is not able to make tunable emissions. We have added comparative photo-physical properties of 0D metal halide with trivalent units in Table 3.

2.3.2. In-Based

Indium (In)-based zero-dimensional materials, such as Cs3InBr6, Cs2InCl5(H2O), Rb2InCl5(H2O), and Cs2InBr5(H2O), and doping with different dopants are reported in [39,40,41,42,43]. Cs3InBr6 nanocrystals were synthesized by the hot-injection method, and the material exhibits broadband luminescence at an emission of 450 nm in Figure 6a [42]. The emission was due to distortions in the [InBr6]3− octahedral. Recently Cs2InCl5 (H2O), Rb2InCl5(H2O), and Cs2InBr5 (H2O) compounds have also attracted interest, which makes them tunable emission with dopants. Rb2InCl5(H2O) has [InCl5(H2O)]2− an isolated octahedral coordinate (In) with the orthorhombic crystal system of the Pnma space group [39]. The material has low photoluminescence quantum yield due to parity forbidden transition and parity broken after substitution of Sb3+ at the In3+ site. Resultantly, the PLQY of Rb2InCl5(H2O) enhances from 1 to 90% after Sb3+ doping at the emission center at 600 nm, which makes a yellow emission. Furthermore, Rb3InCl6:Sb shows a green emission at 497 nm with a PLQY of 95%, as shown in Figure 6b. These results attribute the blue shift in the STE state of Rb3InCl6:Sb compared to Rb2InCl5·H2O: Sb. The luminescence shifts with the change in STE states. Zhou et al. [40] synthesized a Cs3InBr5(H2O) single crystal with a quantum yield of 33%, as shown in Figure 6c. Figure 6d shows red light emission centered at 695 nm with UV excitation.
Table 3. Properties of 0D metal halide with trivalent units.
Table 3. Properties of 0D metal halide with trivalent units.
FormulaMethodIsolated
Species
Em.
Wave. (nm)
FWHM
(nm)
PLQY
(%)
Ref.
Cs3BiCl6 NCHot injection[BiCl6]3−39160 [38]
Cs3Bi2Br9Hot injection[BiBr6]3−46840 4.5%[37]
Cs3InBr6 NCHot injection[InBr6]3−450∼10022.3%[42]
Cs2InBr5(H2O)Temperature-lowering crystallization[InBr5(H2O)]2−695 33%[40]
K3SbCl6 NCHot injection[SbCl6]3−440102 22.3%[44]
Cs3YbCl6Heating-up[YbCl6]3−4297659.8%[45]
Cs3CeBr6 SCSolid state[CeBr6]3−391,
421
92.5%[16]

2.3.3. Sb-Based

Antimony (Sb)-based zero-dimensional materials such as K3SbCl6 and Rb7Sb3Cl16 doping with different dopants are reported [44,46,47]. K3SbCl6 nanocrystals were synthesized by hot-injection method with Mn2+ dopant [44]. Pristine K3SbCl6 nanocrystal has a monoclinic crystal system where isolated [SbCl6]3− octahedral are surrounded by K+ ions. Pristine K3SbCl6 has blue emission centered at 440 nm with a quantum yield of 22.3%. Mn2+ substitute Sb3+ in octahedral coordination gives emission at 600 nm by 4T16A1 transition by energy transfer (Figure 6e). Figure 6f represents the possible energy transfer from STE to Mn2+. The white LED was fabricated further on a UV chip with K3SbCl6: Mn2+ nanocrystals with single materials. Also, zero-dimensional Rb7Sb3Cl16 nanocrystals are synthesized, exhibiting yellow broadband emission at 567 nm [46]. Pinchetti et al. used Rb7Sb3Cl16 nanocrystals doped with Mn2+ showing energy transfer and magnetic properties [47]. These results are attributed to the STE-derived energy transfer to Mn2+. The material with low doping shows paramagnetic nature, and an increased amount of dopant changes it to antiferromagnetic nature.

2.3.4. Rare-Earth-Based

Rare earth trivalent cations can occupy the B site with the octahedral units. So far, only articles report the synthesis of Cs3LnCl6 (Ln = Y, Ce, Gd, Er, Tm, and Yb) nanocrystals by the heating-up method [45]. In this method, both precursors are mixed with oleic acid and oleylamine ligand with the nonpolar solvent. The reaction was first heated at 130 °C to remove water content from nonpolar solvent and then heated at 270 °C to complete the formation of nanocrystals. All compound shows the highest quantum efficiency for blue light. Figure 6g shows the emission of Cs3YbCl6 centered at 429 nm at the excitation of 355 nm with the quantum efficiency of 59.8%. These results suggest zero-dimensional compounds are materials that can be made with rare earth elements. However, further study is needed on their characteristic emissions in the infrared region. So far, only CsMnBr3 is the ideal sensitizer material to transfer energy to rare-earth elements [33].
Wang et al. also report the synthesis of Cs3CeBr5 with a quantum efficiency of up to 90% for powder and thin film [16]. Cs3CeBr6 has orthorhombic crystal systems with the Pbcm space group. The solid-state method is used to synthesize powder of Cs3CeBr6 with isolated [CeBr6]3− in octahedral coordination. Thin films of Cs3CeBr6 were fabricated by thermal evaporation of CsBr and CeBr3. The Cs3CeBr6 system has dual emissions at 391 and 421 nm due to the transition between Ce 5d Figure 6h, and the external quantum efficiency of Cs3CeBr6 was recorded at a maximum of up to 0.46%.

3. Application of Zero-Dimensional Materials

The zero-dimensional materials are used for several optoelectronics applications, such as a light-emitting diode, ultrafast switching memory, photodetector, and scintillators.

3.1. Light-Emitting Diode

The white light-emitting diode is fabricated on a UV chip using zero-dimensional materials, as shown in Figure 7a,b. High luminescence green-emissive Cs4SnBr6 are mixed with blue- and red-emissive phosphors BaMgAl10O17:Eu2+ and CaAlSiN3:Eu2+ to generate white light, as shown in Figure 7b, respectively. The fabricated devices have a color-rendering index of 96% with CIE coordinates of (0.34, 0.33) at 20 mA. The driving current increased to 200 mA, and only the color rendering index was reduced to 86%, making the device stable. The device was continuously operated at 1000 h, the color rendering index remained unchanged, and the relative luminous efficiency was reduced by up to 79%. Along with it, the CIE coordinate only shift from (0.34, 0.33) to (0.33, 0.32) after 1000 h. These results show that the Cs4SnBr6-prepared device has excellent operation stability. Xu et al. [29] also reported a device using Cs4SnBr6 nanocrystals with red- and green-emissive phosphors and achieved a color rendering index of 92%.
Lead-free copper-based zero-dimensional material is a possible candidate for blue light-emitting diode application because of the high quantum yield at the emission center at ∼450 nm [20]. Cs3Cu2I5 nanocrystals are used for deep blue LED and have achieved external quantum efficiency of up to 1.12%. As shown in Figure 7c,d, the device was fabricated with an active layer of Cs3Cu2I5 nanocrystal with 1,3,5-tri(phenyl-2-benzimi-dazolyl)-benzene (TP) as the electron injecting layer and NiO as the hole injecting layer for the device. The fabricated device has a coordinate of (0.16, 0.07) for blue LED with NTSC standards. The device was operated for 170 h, and after 1 h of relaxation, electroluminescence almost recovered to the initial level, which indicates the stability of the devices. These recent studies suggest the use of zero-dimensional material for WLED and blue LED.

3.2. Ultrafast Switching Memory

Lead halide perovskite is developed as the ideal candidate for resistive switching memory because of multilevel data storage and low operational voltage [51,52]. Halide perovskites have applied resistive switching because of the presence of halide defects or migration of ions [48]. Toxicity and stability issues mean that lead is not environment friendly for further application. Cs3Sb2I5-based resistive switching memory was fabricated, as shown in Figure 7e [48]. The electrical signals are recorded by applying a voltage at Au when ITO is grounded. Figure 7f,g showed the I-V characteristics of the device when the applied voltage was set at 6 V with a width of 20 ns and a reset voltage of −6 V with a width of 20 ns. To find the fast switching, bias voltage was used from 0 to 0.4 V to find the resistance state before and after the pulse. With the applied positive voltage pulse, the state changes from a high-resistance state to a low-resistance state, and further negative voltage pulse shifts the state and vice versa. All these results confirm that the Cs3Sb2I5-based device can be operated with a fast switching speed of 20 ns. So far, only Cs3Sb2I5 has been applied for fast-switching memory applications.

3.3. Photodetectors and Scintillators

After using CsPbX3 and lead-free double perovskite for photodetector applications, zero-dimensional Cs3BiBr6 lead-free luminescent materials were used to detect UV–Vis light [49]. Cs3BiBr6 was deposited on the ITO substrate, and the photocurrent was measured at the bias voltage of 6V by 400 nm monochromatic illumination, as shown in Figure 7h. The fabricated device has 0.3 nA dark current and increases with an increasing power density, as shown in Figure 7i. The fabricated device has responsivity and detectivity of ∼25 mA W−1 and ∼0.8 × 109 Jones, respectively. Highly crystalline zero-dimensional materials could be ideal for detection application in UV light with a high PLQY.
Scintillators convert high energy χ-ray, γ-ray, and α particles into low energy UV– visible light. Recently doped and un-doped Cs3Cu2I5 were used for χ-ray and γ-ray detections [22,50,53]. Perovskite has high photoluminescence quantum yield and high effective atomic number, making it a more suitable candidate for scintillator application. The Bridgman method is used to synthesize the single crystal, as shown in Figure 7j [50]. Cs3Cu2I5 has a high quantum yield of 68% at an emission center of ∼450 nm, and further doping with Tl and In-enhanced it to 79.2% and 88.4%, respectively. Single crystal of Cs3Cu2I5:Tl has a scintillation yield of 1,50,000 photons/MeV with a detection limit of 66.3 nGyairs−1. Additionally, Cs3Cu2I5:Tl single crystal has a scintillation yield of 87,000 photons/MeV under 137Cs γ-ray detection, as shown in Figure 7k [50]. The In-doped Cs3Cu2I5 obtained a scintillation yield of 53,000 photons/MeV for χ-ray [22]. These results show that Cs3Cu2I5 highly crystalline materials can be used for high-energy photon detections.

4. Conclusions and Outlook

Lead halide perovskite’s toxicity and stability issues concern finding an alternative with similar properties. Tin could be alternative to replace Pb, although it is not stable and oxidizes in contact with the environment. Several research groups worked on Lead-free double perovskite, which has broadband emission and is excitable by UV light. So far, octahedral and tetrahedral coordinated luminescence materials have been reported with different applications. The challenges and perspectives regarding zero-dimensional discussed as follows:
  • Copper-derived Cs3Cu2I5 has emission at ∼445 nm, and the quantum yield increased to ∼88%. This system also exhibits external quantum efficiency of ∼1.12% and further improvement will make it a promising candidate for deep blue LEDs. However, further improvement in the future is needed to improve the quantum efficiency. Along with LED, these compounds are widely applied for χ-ray scintillation, where Cs3Cu2I5: Tl has a higher scintillation yield than CsI: Tl.
  • Bimetallic Cs-Ag-X systems emission can be tuned with the change in the crystallographic dimensions and halides. However, these systems achieved high quantum yields between 17 and 68% but need to be excited by the UV light. This concern restricts the use of the material for LED. Along with these bi-metallic systems, Bi-based Cs3Bi2X9 nanocrystals are reported but suffer from a very low quantum yield of up to 4.5% only.
  • The Quantum efficiency of Cs4SnBr6 enhanced achieved up to ∼62%, making it the ideal material for LED applications with high stability. However, the material can only be excitable by UV chip for WLED, so the stability of the Cs4SnX6 needed to be detail study.
  • Cs3MnBr5 and CsMnBr3 have green and red light emissions with tetrahedral and octahedral coordination, respectively. These materials have emissions around ∼520 nm and ∼620 nm, which are ideal for lighting applications. However, further research is needed to enhance quantum efficiency and stability for industrial applications. The hot-injection method is complicated for the nanocrystal, and further research is needed to improve the process.
  • So far, these zero-dimensional materials are applied for several applications discussed in the application sections. However, these materials could be applied for solar cells, photo-catalysis, and sensing applications.
Finally, several zero-dimensional materials are studied but still need deep fundamental research to develop materials for optoelectronics industries in the future. Ag, Bi, and Mn-based compounds have tunable emissions but must improve the emission properties for real applications.

Author Contributions

Conceptualization, K.D., W.-T.H. and R.-S.L.; writing—review and editing, K.D., W.-T.H. and R.-S.L.; visualization, K.D. and W.-T.H.; supervision, R.-S.L.; project administration, R.-S.L.; funding acquisition, R.-S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Science and Technology Council in Taiwan (contract NSTC 109-2113-M-002-020-MY3).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with Wide Color Gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Swarnkar, A.; Marshall, A.R.; Sanehira, E.M.; Chernomordik, B.D.; Moore, D.T.; Christians, J.A.; Chakrabarti, T.; Luther, J.M. Quantum Dot–Induced Phase Stabilization of α-CsPbI3 Perovskite for High-Efficiency Photovoltaics. Science 2016, 354, 92–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Dave, K.; Bao, Z.; Nakahara, S.; Ohara, K.; Masada, S.; Tahara, H.; Kanemitsu, Y.; Liu, R.S. Improvement in Quantum Yield by Suppression of Trions in Room Temperature Synthesized CsPbBr3 Perovskite Quantum Dots for Backlight Displays. Nanoscale 2020, 12, 3820–3826. [Google Scholar] [CrossRef] [PubMed]
  4. Dou, L.; Yang, Y.; You, J.; Hong, Z.; Chang, W.H.; Li, G.; Yang, Y. Solution-processed Hybrid Perovskite Photodetectors with High detectivity. Nat. Commun. 2014, 5, 5404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Dave, K.; Fang, M.H.; Bao, Z.; Fu, H.T.; Liu, R.S. Recent Developments in Lead-Free Double Perovskites: Structure, Doping, and Applications. Chem. Asian J. 2020, 15, 242–252. [Google Scholar] [CrossRef]
  6. Gupta, S.; Bendikov, T.; Hodes, G.; Cahen, D. CsSnBr3, A Lead-Free Halide Perovskite for Long Term Solar Cell Application: Insights on SnF2 Addition. ACS Energy Lett. 2016, 1, 1028–1033. [Google Scholar] [CrossRef] [Green Version]
  7. Luo, J.; Wang, X.; Li, S.; Liu, J.; Guo, Y.; Niu, G.; Yao, L.; Fu, Y.; Gao, L.; Dong, Q.; et al. Efficient and Stable Emission of Warm–White Light from Lead–Free Halide Double Perovskites. Nature 2018, 563, 541–545. [Google Scholar] [CrossRef]
  8. Dave, K.; Huang, W.T.; Leśniewski, T.; Lazarowska, A.; Jankowski, D.; Mahlik, S.; Liu, R.S. Photoluminescence Enhancement Study in a Bi-doped Cs2AgInCl6 Double Perovskite by Pressure and Temperature-dependent Self-Trapped Exciton Emission. Dalton Trans. 2022, 51, 2026–2032. [Google Scholar] [CrossRef]
  9. Li, M.; Xia, Z. Recent Progress of Zero-Dimensional Luminescent Metal Halides. Chem. Soc. Rev. 2021, 50, 2626–2662. [Google Scholar] [CrossRef]
  10. Wang, L.; Liu, H.; Zhang, Y.; Mohammed, O.F. Photoluminescence Origin of Zero-Dimensional Cs4PbBr6 Perovskite. ACS Energy Lett. 2020, 5, 87–99. [Google Scholar] [CrossRef]
  11. 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]
  12. Zhou, C.; Xu, L.J.; Lee, S.; Lin, H.; Ma, B. Recent Advances in Luminescent Zero-Dimensional Organic Metal Halide Hybrids. Adv. Optical Mater. 2021, 9, 2001766. [Google Scholar] [CrossRef]
  13. Schubert, E.F. Light-Emitting Diodes; Schubert, E.F., Ed.; Rensselaer Polytechnic Institute: New York, NY, USA, 2018. [Google Scholar]
  14. Zhou, L.; Liao, J.F.; Kuang, D.B. An Overview for Zero-Dimensional Broadband Emissive Metal-Halide Single Crystals. Adv. Optical Mater. 2021, 9, 2100544. [Google Scholar] [CrossRef]
  15. Song, K.W.; Williams, R.T. Self-Trapped Excitons; Springer Science & Business Media: Berlin, Germany, 2013; Volume 105. [Google Scholar]
  16. Wang, L.; Guo, Q.; Duan, J.; Xie, W.; Ji, G.; Li, S.; Chen, C.; Li, J.; Yang, L.; Tan, Z.; et al. Exploration of Nontoxic Cs3CeBr6 for Violet Light-Emitting Diodes. ACS Energy Lett. 2021, 6, 4245–4254. [Google Scholar] [CrossRef]
  17. 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]
  18. Fowler, W.B.; Marrone, M.J.; Kabler, M.N. Theory of Self-Trapped Exciton Luminescence in Halide Crystals. Phys. Rev. B 1973, 8, 5909–5919. [Google Scholar] [CrossRef]
  19. Liu, X.; Yu, Y.; Yuan, F.; Zhao, C.; Dong, H.; Jiao, B.; Wu, Z. Vacuum Dual-Source Thermal-Deposited Lead-Free Cs3Cu2I5 Films with High Photoluminescence Quantum Yield for Deep-Blue Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2020, 12, 52967–52975. [Google Scholar] [CrossRef]
  20. 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]
  21. Roccanova, R.; Yangui, A.; Nhalil, H.; Shi, H.; Du, M.H.; Saparov, B. Near-Unity Photoluminescence Quantum Yield in Blue-Emitting Cs3Cu2Br5–xIx (0 ≤ x ≤ 5). ACS Appl. Electron. Mater. 2019, 1, 269–274. [Google Scholar] [CrossRef]
  22. Wang, Q.; Zhou, Q.; Nikl, M.; Xiao, J.; Kucerkova, R.; Beitlerova, A.; Babin, V.; Prusa, P.; Linhart, V.; Wang, J.; et al. Highly Resolved X-Ray Imaging Enabled by In(I) Doped Perovskite-Like Cs3Cu2I5 Single Crystal Scintillator. Adv. Opt. Mater. 2022, 10, 2200304. [Google Scholar] [CrossRef]
  23. Huang, X.; Sun, Q.; Devakumar, B. Facile Low-temperature Solid-state Synthesis of Efficient Blue-emitting Cs3Cu2I5 Powder Phosphors for Solid-state Lighting. Mater. Today Chem. 2020, 17, 100288. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Zhao, R.; Teng, S.; Huang, K.; Zhang, L.; Wang, D.; Yang, W.; Xie, R.; Pradhan, N. Color Tunable Self-Trapped Emissions from Lead-Free All Inorganic IA-IB Bimetallic Halides Cs-Ag-X (X = Cl, Br, I). Small 2020, 16, 2004272. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, M.; Wang, X.; Yang, B.; Zhu, J.; Niu, G.; Wu, H.; Yin, L.; Du, X.; Niu, M.; Ge, Y.; et al. Metal Halide Scintillators with Fast and Self-Absorption-Free Defect-Bound Excitonic Radioluminescence for Dynamic X-Ray Imaging. Adv. Funct. Mater. 2021, 31, 2007921. [Google Scholar] [CrossRef]
  26. Zhang, Q.; Liu, S.; He, M.; Zheng, W.; Wan, Q.; Liu, M.; Liao, X.; Zhan, W.; Yuan, C.; Liu, J.; et al. Stable Lead-Free Tin Halide Perovskite with Operational Stability >1200 h by Suppressing Tin(II) Oxidation. Angew. Chem. 2022, 61, e202205463. [Google Scholar]
  27. Benin, B.M.; Dirin, D.N.; Morad, V.; Wörle, M.; Yakunin, S.; Rainò, G.; Nazarenko, O.; Fischer, M.; Infante, I.; Kovalenko, M.V. Highly Emissive Self-Trapped Excitons in Fully Inorganic Zero-Dimensional Tin Halides. Angew. Chem. Int. Ed. 2018, 57, 11329–11333. [Google Scholar] [CrossRef] [Green Version]
  28. Tan, L.; Wang, W.; Li, Q.; Luo, Z.; Zou, C.; Tang, M.; Zhang, L.; He, J.; Quan, Z. Colloidal Syntheses of Zero-Dimensional Cs4SnX6 (X = Br, I) Nanocrystals with High Emission Efficiencies. Chem. Commun. 2020, 56, 387–390. [Google Scholar] [CrossRef]
  29. Xu, K.; Wei, Q.; Wang, H.; Yao, B.; Zhou, W.; Gao, R.; Chen, H.; Li, H.; Wang, J.; Ning, Z. The 3D-Structure-Mediated Growth of Zero-Dimensional Cs4SnX6 Nanocrystals. Nanoscale 2022, 14, 2248–2255. [Google Scholar] [CrossRef]
  30. Wang, X.; Shen, Q.; Chen, Y.; Ali, N.; Ren, Z.; Bi, G.; Wu, H. Self-Trapped Exciton Emission in an Sn(ii)-Doped all-Inorganic Zero-Dimensional Zinc Halide Perovskite Variant. Nanoscale 2021, 13, 15285–15291. [Google Scholar] [CrossRef]
  31. Zhao, H.; Wei, Y.; Chen, Y.; Liu, C.; Jiang, F.; Liu, Y.; Hong, M. Blue-Emitting 0D Cs3ZnX5 (X = Cl, Br) Perovskite Nanocrystals Based on Self-Trapped Excitons. J. Lumin. 2022, 249, 119048. [Google Scholar] [CrossRef]
  32. Liu, X.; Zhang, W.; Xu, R.; Tu, J.; Fang, G.; Pan, Y. Bright Tunable Luminescence of Sb3+ Doping in Zero-Dimensional Lead-Free Halide Cs3ZnCl5 Perovskite Crystals. Dalton Trans. 2022, 51, 10029–10035. [Google Scholar] [CrossRef]
  33. Bahmani Jalali, H.; Pianetti, A.; Zito, J.; Imran, M.; Campolucci, M.; Ivanov, Y.P.; Locardi, F.; Infante, I.; Divitini, G.; Brovelli, S.; et al. Cesium Manganese Bromide Nanocrystal Sensitizers for Broadband Vis-to-NIR Downshifting. ACS Energy Lett. 2022, 7, 1850–1858. [Google Scholar] [CrossRef]
  34. Almutlaq, J.; Mir, W.J.; Gutiérrez Arzaluz, L.; Yin, J.; Vasylevskyi, S.; Maity, P.; Liu, J.; Naphade, R.; Mohammed, O.F.; Bakr, O.M. CsMnBr3: Lead-Free Nanocrystals with High Photoluminescence Quantum Yield and Picosecond Radiative Lifetime. ACS Mater. Lett. 2021, 3, 290–297. [Google Scholar] [CrossRef]
  35. Kong, Q.; Yang, B.; Chen, J.; Zhang, R.; Liu, S.; Zheng, D.; Zhang, H.; Liu, Q.; Wang, Y.; Han, K. Phase Engineering of Cesium Manganese Bromides Nanocrystals with Color-Tunable Emission. Angew. Chem. 2021, 60, 19653–19659. [Google Scholar] [CrossRef]
  36. Yang, H.; Cai, T.; Dube, L.; Chen, O. Synthesis of Double Perovskite and Quadruple Perovskite Nanocrystals Through Post-Synthetic Transformation Reactions. Chem. Sci. 2022, 13, 4874–4883. [Google Scholar] [CrossRef]
  37. 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]
  38. Yang, H.; Cai, T.; Liu, E.; Hills-Kimball, K.; Gao, J.; Chen, O. Synthesis and Transformation of Zero-Dimensional Cs3BiX6 (X = Cl, Br) Perovskite-Analogue Nanocrystals. Nano Res. 2020, 13, 282291. [Google Scholar] [CrossRef] [Green Version]
  39. Han, P.; Luo, C.; Yang, S.; Yang, Y.; Deng, W.; Han, K. All-Inorganic Lead-Free 0D Perovskites by a Doping Strategy to Achieve a PLQY Boost from <2% to 90%. Angew. Chem. Int. Ed. 2020, 59, 12709–12713. [Google Scholar]
  40. Zhou, L.; Liao, J.F.; Huang, Z.G.; Wei, J.H.; Wang, X.D.; Li, W.G.; Chen, H.Y.; Kuang, D.B.; Su, C.Y. A Highly Red-Emissive Lead-Free Indium-Based Perovskite Single Crystal for Sensitive Water Detection. Angew. Chem. Int. Ed. 2019, 58, 5277–5281. [Google Scholar] [CrossRef]
  41. Chang, T.; Wang, H.; Gao, Y.; Cao, S.; Zhao, J.; Zou, B.; Zeng, R. Component Engineering to Tailor the Structure and Optical Properties of Sb-Doped Indium-Based Halides. Inorg. Chem. 2022, 61, 1486–1494. [Google Scholar] [CrossRef]
  42. Zhang, F.; Yang, D.; Shi, Z.; Qin, C.; Cui, M.; Ma, Z.; Wang, L.; Wang, M.; Ji, X.; Chen, X.; et al. Stable Zero-Dimensional Cesium Indium Bromide Hollow Nanocrystals Emitting Blue Light from Self-Trapped Excitons. Nano Today 2021, 38, 101153. [Google Scholar] [CrossRef]
  43. Wei, J.H.; Luo, J.B.; Liao, J.F.; Ou, W.T.; Kuang, D.B. Te4+-Doped Cs2InCl5·H2O Single Crystals for Remote Optical Thermometry. Sci. China Mater. 2022, 65, 764–772. [Google Scholar] [CrossRef]
  44. 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]
  45. Pinchetti, V.; Moro, F.; Zhang, B.; Fanciulli, M.; De Trizio, L.; Meinardi, F.; Manna, L.; Brovelli, S. Magnetic Transitions and Energy Transfer Processes in Sb-Based Zero-Dimensional Metal Halide Nanocrystals Doped with Manganese. ACS Energy Lett. 2022, 7, 1566–1573. [Google Scholar] [CrossRef]
  46. Shao, T.; Fang, Y.; He, C.; Zhang, L.; Wang, K. Pressure Tuning of Optical Properties and Structures in All-Inorganic Halide Perovskite Rb7Sb3Cl16. Inorg. Chem. 2022, 61, 5184–5189. [Google Scholar] [CrossRef]
  47. Lee, M.; Lee, D.H.D.; Hong, S.V.; Woo, H.Y.; Chae, J.Y.; Lee, D.W.; Han, M.J.; Paik, T. Highly Luminescent and Multifunctional Zero-Dimensional Cesium Lanthanide Chloride (Cs3LnCl6) Colloidal Nanocrystals. Adv. Opt. Mater. 2022, 10, 2102727. [Google Scholar] [CrossRef]
  48. Park, Y.; Kim, S.H.; Lee, D.; Lee, J.S. Designing Zero-Dimensional Dimer-Type All-Inorganic Perovskites for Ultra-Fast Switching Memory. Nat. Commun. 2021, 12, 3527. [Google Scholar] [CrossRef]
  49. Tang, Y.; Liang, M.; Chang, B.; Sun, H.; Zheng, K.; Pullerits, T.; Chi, Q. Lead-Free Double Halide Perovskite Cs3BiBr6 with Well-Defined Crystal Structure and High Thermal Stability for Optoelectronics. J. Mater. Chem. C 2019, 7, 3369–3374. [Google Scholar] [CrossRef] [Green Version]
  50. Cheng, S.; Nikl, M.; Beitlerova, A.; Kucerkova, R.; Du, X.; Niu, G.; Jia, Y.; Tang, J.; Ren, G.; Wu, Y. Ultrabright and Highly Efficient All-Inorganic Zero-Dimensional Perovskite Scintillators. Adv. Opt. Mater. 2021, 9, 2100460. [Google Scholar] [CrossRef]
  51. Seo, J.Y.; Choi, J.; Kim, H.S.; Kim, J.; Yang, J.M.; Cuhadar, C.; Han, J.S.; Kim, S.J.; Lee, D.; Jang, H.W.; et al. Wafer-Scale Reliable Switching Memory Based on 2-Dimensional Layered Organic–Inorganic Halide Perovskite. Nanoscale 2017, 9, 15278–15285. [Google Scholar] [CrossRef] [Green Version]
  52. Kim, H.; Choi, M.J.; Suh, J.M.; Han, J.S.; Kim, S.G.; Le, Q.V.; Kim, S.Y.; Jang, H.W. Quasi-2D Halide Perovskites for Resistive Switching Devices with ON/OFF Ratios Above 109. NPG Asia Mater. 2020, 12, 21. [Google Scholar] [CrossRef]
  53. Cheng, S.; Beitlerova, A.; Kucerkova, R.; Nikl, M.; Ren, G.; Wu, Y. Zero-Dimensional Cs3Cu2I5 Perovskite Single Crystal as Sensitive X-Ray and γ-Ray Scintillator. Phys. Status Solidi RRL 2020, 14, 2000374. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of electronically zero-dimensional with the isolated (a) tetrahedral and (b) octahedral.
Figure 1. Crystal structure of electronically zero-dimensional with the isolated (a) tetrahedral and (b) octahedral.
Crystals 13 00499 g001
Figure 2. The possible emission mechanism for (a) light absorption, (b) band-to-band emission, (c) exciton emission, (d) emission from defects, (e) emission from ions, (f) self-trap exciton (STE) emission.
Figure 2. The possible emission mechanism for (a) light absorption, (b) band-to-band emission, (c) exciton emission, (d) emission from defects, (e) emission from ions, (f) self-trap exciton (STE) emission.
Crystals 13 00499 g002
Figure 3. (a) Absorbance, photoluminescence excitation and emission spectra of Cs4SnBr6 Cs4SnBr6I6–x and Cs4SnI6. Reproduced with permission from [28], Copyright 2020, The Royal Society of Chemistry. (b) Normal and UV light images of Cs4SnBr6, Cs4SnBr3I3, and Cs4SnI6. Reproduced with permission from [29], Copyright 2022, The Royal Society of Chemistry. (c) Photoluminescence excitation and emission spectra of Cs4Sn(Br,I)6. (d) Photoluminescence excitation and emission spectra of Cs4-–xAxSn(Br,I)6, where A = K+, Rb+. (e) UV light image of Cs4–xAxSn(Br, I)6 powders. Reproduced with permission from [27], Copyright 2022, Wiley-VCH. (f) Normal and UV light image of Cs4SnBr6 and Cs4SnBr6–SnF2. Reproduced with permission from [26], Copyright 2022, Wiley-VCH. (g) Schematic illustration of STE emission. Reproduced with permission from [29], Copyright 2022, The Royal Society of Chemistry.
Figure 3. (a) Absorbance, photoluminescence excitation and emission spectra of Cs4SnBr6 Cs4SnBr6I6–x and Cs4SnI6. Reproduced with permission from [28], Copyright 2020, The Royal Society of Chemistry. (b) Normal and UV light images of Cs4SnBr6, Cs4SnBr3I3, and Cs4SnI6. Reproduced with permission from [29], Copyright 2022, The Royal Society of Chemistry. (c) Photoluminescence excitation and emission spectra of Cs4Sn(Br,I)6. (d) Photoluminescence excitation and emission spectra of Cs4-–xAxSn(Br,I)6, where A = K+, Rb+. (e) UV light image of Cs4–xAxSn(Br, I)6 powders. Reproduced with permission from [27], Copyright 2022, Wiley-VCH. (f) Normal and UV light image of Cs4SnBr6 and Cs4SnBr6–SnF2. Reproduced with permission from [26], Copyright 2022, Wiley-VCH. (g) Schematic illustration of STE emission. Reproduced with permission from [29], Copyright 2022, The Royal Society of Chemistry.
Crystals 13 00499 g003
Figure 4. (a) Sunlight and UV light image of different amount of Sb3+ in Cs3ZnCl5. (b) Photoluminescence excitation and (c) emission spectra for Cs3ZnCl5:Sb3+ at 620 nm and 320 nm, respectively. Reproduced with permission from [32], Copyright 2022, The Royal Society of Chemistry.
Figure 4. (a) Sunlight and UV light image of different amount of Sb3+ in Cs3ZnCl5. (b) Photoluminescence excitation and (c) emission spectra for Cs3ZnCl5:Sb3+ at 620 nm and 320 nm, respectively. Reproduced with permission from [32], Copyright 2022, The Royal Society of Chemistry.
Crystals 13 00499 g004
Figure 5. (a) Synthesis process for CsMnBr3 and Cs3MnBr5 and their crystal structures, respectively. (b) Transformation of CsMnBr3 to Cs2MnBr4·2H2O in moisture (99%) and further transformation by dehydration to CsMnBr3 and Cs3MnBr5. (c) Transformation of Cs3MnBr5 to Cs2MnBr4·2H2O in moisture (99%) and further transformation by dehydration to CsMnBr3 and Cs3MnBr5. Reproduced with permission from [35], Copyright 2021, Wiley-VCH.
Figure 5. (a) Synthesis process for CsMnBr3 and Cs3MnBr5 and their crystal structures, respectively. (b) Transformation of CsMnBr3 to Cs2MnBr4·2H2O in moisture (99%) and further transformation by dehydration to CsMnBr3 and Cs3MnBr5. (c) Transformation of Cs3MnBr5 to Cs2MnBr4·2H2O in moisture (99%) and further transformation by dehydration to CsMnBr3 and Cs3MnBr5. Reproduced with permission from [35], Copyright 2021, Wiley-VCH.
Crystals 13 00499 g005
Figure 6. (a) UV–vis, photoluminescence excitation, and emission spectra of Cs3InBr6 nanocrystals in solution. Reproduced with permission from [42], Copyright 2021, Elsevier. (b) Photoluminescence emission spectra of Rb3InCl6:Sb and Rb2InCl5(H2O):Sb. Reproduced with permission from [39], Copyright 2020, Wiley–VCH (c) Ambient and UV light image of Cs2InBr5.H2O (d) Photoluminescence excitation (λem = 695 nm) and emission(λex = 355 nm) spectra of Cs2InBr5.H2O. Reproduced with permission from [40], Copyright 2019, Wiley–VCH. (e) Photoluminescence emission spectra of K3SbCl6 with Mn2+ doping (0.12%, 1.5%, 3.7%, 4.2%, and 5.8%). (f) Mechanism of energy transfer of self-trapped exciton emission to Mn2+. Reproduced with permission from [44], Copyright 2021, Elsevier. (g) Photoluminescence excitation (λem = 429 nm) and emission (λex = 355 nm) spectra of Cs3YbBr6 nanocrystal. Reproduced with permission from [45], Copyright 2022, Wiley–VCH. (h) Photoluminescence excitation and emission spectra of Cs3CeBr6 nanocrystals at different excitation and emission wavelength. Reproduced with permission from [16], Copyright 2021, the American Chemical Society.
Figure 6. (a) UV–vis, photoluminescence excitation, and emission spectra of Cs3InBr6 nanocrystals in solution. Reproduced with permission from [42], Copyright 2021, Elsevier. (b) Photoluminescence emission spectra of Rb3InCl6:Sb and Rb2InCl5(H2O):Sb. Reproduced with permission from [39], Copyright 2020, Wiley–VCH (c) Ambient and UV light image of Cs2InBr5.H2O (d) Photoluminescence excitation (λem = 695 nm) and emission(λex = 355 nm) spectra of Cs2InBr5.H2O. Reproduced with permission from [40], Copyright 2019, Wiley–VCH. (e) Photoluminescence emission spectra of K3SbCl6 with Mn2+ doping (0.12%, 1.5%, 3.7%, 4.2%, and 5.8%). (f) Mechanism of energy transfer of self-trapped exciton emission to Mn2+. Reproduced with permission from [44], Copyright 2021, Elsevier. (g) Photoluminescence excitation (λem = 429 nm) and emission (λex = 355 nm) spectra of Cs3YbBr6 nanocrystal. Reproduced with permission from [45], Copyright 2022, Wiley–VCH. (h) Photoluminescence excitation and emission spectra of Cs3CeBr6 nanocrystals at different excitation and emission wavelength. Reproduced with permission from [16], Copyright 2021, the American Chemical Society.
Crystals 13 00499 g006
Figure 7. (a) Schematic diagram of LED devices based on Cs4SnBr6-SnF2. (b) Emission spectra of the fabricated white light diode (WLED) by blue-emissive BaMgAl10O17:Eu2+, green-emissive Cs4SnBr6-SnF2 sand red-emissive CaAlSiN3:Eu2+ phosphors. Reproduced with permission from [26], Copyright 2022, Wiley-VCH. (c) Schematic diagram of blue light-emitting diode (LED) made by Cs3Cu2I5 nanocrystals. (d) Electroluminescence spectra of Cs3Cu2I5 nanocrystals made blue LED device for fresh and at different time intervals. Reproduced with permission from [20], Copyright 2020, the American Chemical Society. (e) Schematic diagram of resistive switching memory made by Cs3Sb2I9. I-V curve of resistive switching memory made (f) set pulse (g) reset pulse. Reproduced with permission from [48], Copyright 2021, the Nature Publishing Group. (h) Schematic diagram of Photodetector device made by Cs3BiBr6. (i) I-V characteristic of photodetector device at different power densities. Reproduced with permission from [49], Copyright 2019, Royal Society of Chemistry) (j) Tl-doped Cs3Cu2I5 crystal grown by Bridgman method. (k) Comparison with different scintillators. Reproduced with permission from [50], Copyright 2021, Wiley-VCH.
Figure 7. (a) Schematic diagram of LED devices based on Cs4SnBr6-SnF2. (b) Emission spectra of the fabricated white light diode (WLED) by blue-emissive BaMgAl10O17:Eu2+, green-emissive Cs4SnBr6-SnF2 sand red-emissive CaAlSiN3:Eu2+ phosphors. Reproduced with permission from [26], Copyright 2022, Wiley-VCH. (c) Schematic diagram of blue light-emitting diode (LED) made by Cs3Cu2I5 nanocrystals. (d) Electroluminescence spectra of Cs3Cu2I5 nanocrystals made blue LED device for fresh and at different time intervals. Reproduced with permission from [20], Copyright 2020, the American Chemical Society. (e) Schematic diagram of resistive switching memory made by Cs3Sb2I9. I-V curve of resistive switching memory made (f) set pulse (g) reset pulse. Reproduced with permission from [48], Copyright 2021, the Nature Publishing Group. (h) Schematic diagram of Photodetector device made by Cs3BiBr6. (i) I-V characteristic of photodetector device at different power densities. Reproduced with permission from [49], Copyright 2019, Royal Society of Chemistry) (j) Tl-doped Cs3Cu2I5 crystal grown by Bridgman method. (k) Comparison with different scintillators. Reproduced with permission from [50], Copyright 2021, Wiley-VCH.
Crystals 13 00499 g007
Table 1. Properties of 0D metal halide with monovalent units.
Table 1. Properties of 0D metal halide with monovalent units.
FormulaMethodIsolated
Species
Em.
Wave. (nm)
FWHM
(nm)
PLQY
(%)
Ref.
Cs3Cu2I5 Film Thermal evaporation [Cu2I3] 440 70 58[19]
Cs3Cu2I5 NC Hot injection [Cu2I5]3− 445 6387[20]
Cs2AgCl3Ball Milling[AgCl4]3−397..17[24]
Cs2AgBr3Ball Milling[AgBr4]3−524..21[24]
Rb2AgBr3 SCCooling crystallization[AgBr4]3−488..27[25]
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

Dave, K.; Huang, W.-T.; Liu, R.-S. All Inorganic Lead-Free Zero-Dimensional Metal Halide Luminescent Materials and Applications. Crystals 2023, 13, 499. https://doi.org/10.3390/cryst13030499

AMA Style

Dave K, Huang W-T, Liu R-S. All Inorganic Lead-Free Zero-Dimensional Metal Halide Luminescent Materials and Applications. Crystals. 2023; 13(3):499. https://doi.org/10.3390/cryst13030499

Chicago/Turabian Style

Dave, Kashyap, Wen-Tse Huang, and Ru-Shi Liu. 2023. "All Inorganic Lead-Free Zero-Dimensional Metal Halide Luminescent Materials and Applications" Crystals 13, no. 3: 499. https://doi.org/10.3390/cryst13030499

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