Next Article in Journal
Deposition of NiAl/Al3Ni2(CrB2) Coatings from Ni, Al and CrB2 Powders Using Mechanical Synthesis in Planetary Ball Mill
Previous Article in Journal
Thermal Deformation Behavior and Microstructural Evolution of Multicomponent Mg-Li-Zn-Al-Y Alloys under Hot Compression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mapping Uncharted Lead-Free Halide Perovskites and Related Low-Dimensional Structures

by
Anna Dávid
1,*,
Julia Morát
2,
Mengyun Chen
2,
Feng Gao
2,
Mats Fahlman
1 and
Xianjie Liu
1,*
1
Laboratory of Organic Electronics (LOE), Department of Science and Technology, Linköping University, 60174 Norrköping, Sweden
2
Department of Physics, Chemistry and Biology (IFM), Linköping University, 58183 Linköping, Sweden
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(2), 491; https://doi.org/10.3390/ma17020491
Submission received: 4 December 2023 / Revised: 13 January 2024 / Accepted: 16 January 2024 / Published: 19 January 2024

Abstract

:
Research on perovskites has grown exponentially in the past decade due to the potential of methyl ammonium lead iodide in photovoltaics. Although these devices have achieved remarkable and competitive power conversion efficiency, concerns have been raised regarding the toxicity of lead and its impact on scaling up the technology. Eliminating lead while conserving the performance of photovoltaic devices is a great challenge. To achieve this goal, the research has been expanded to thousands of compounds with similar or loosely related crystal structures and compositions. Some materials are “re-discovered”, and some are yet unexplored, but predictions suggest that their potential applications may go beyond photovoltaics, for example, spintronics, photodetection, photocatalysis, and many other areas. This short review aims to present the classification, some current mapping strategies, and advances of lead-free halide double perovskites, their derivatives, lead-free perovskitoid, and low-dimensional related crystals.

1. Introduction: Synthetic Lead-Free Halide Perovskites and Related Low-Dimensional Structures

The invention of crystalline silicon solar cells was a milestone of energy conversion technology development in the 1950s. This achievement led to extensive research and accelerating progress in device fabrication within the photovoltaic field. Now, more than half a century later, several solutions are available [1]. One branch of these devices employs versatile synthetic metal halide perovskites in different architectures, for instance, the Grätzel cell [2,3]. The word perovskite is an umbrella term encompassing both all-inorganic and hybrid organic–inorganic compounds with a perovskite crystal structure. The incorporation of perovskites in dye-sensitized solar (Grätzel) cells catalyzed perovskite research, and remarkable efficiency has been reached with those absorbers. For instance, in 2023, the research group of Prof. De Wolf announced a certified power conversion efficiency (PCE) of 33.2% for monolithic perovskite/silicon tandem solar cell construction. The US National Renewable Energy Laboratory is publishing a constantly updated chart where one can follow the efficiency of devices developed over the timeline (see Figure 1) [4].
In general, perovskite research has found that materials with perovskite crystal structures may also be of interest to several other optoelectronic applications due to their relatively easy processability and unique properties. Other than photovoltaics, examples from the constantly growing list of application areas include photodetectors, spintronic devices, light-emitting diodes, and photocatalysts [5,6]. The most prominent representative compounds from the group of perovskites are lead-based, for instance, methylammonium lead iodide. Lead is abundant, cheap, and easily recyclable from several sources. From a techno-economic aspect, these materials may be competitive in markets such as the Internet of Things and building-integrated photovoltaics due to their high optical absorption coefficient, remarkable defect tolerance, and high electronic dimensionality [7]; moreover, facile film preparation and advanced techniques make in situ and operando studies upon perovskite-based devices feasible to significantly push the development of this field [8]. One of the main issues with these materials is the toxicity of lead, which makes large-scale production challenging from the point of sustainability and safety [9,10,11,12]. The daily intake of lead should not exceed 490 μg for an adult, according to the Food and Agricultural Organization/World Health Organization, but there are no guidelines for children. If lead enters the human body, it is distributed through red blood cells and imposes its toxicity and may cause the symptoms listed in Table 1 [13]. However, in the case of methylammonium lead iodide perovskite, precursor PbI2 only accounts for 3.68% of the global human toxicity potential, while methylammonium iodide has a contribution of 62.31% according to the Life Cycle Assessment by Zhang et al. [14].
Great efforts have been taken to substitute lead in the halide perovskite crystal lattice, but no substitute has shown as remarkable performance in photovoltaics as pristine ones so far. The substitution of lead with other Group IVA elements like tin (or germanium) is an obvious approach; however, perovskites formed with Sn2+ are facing stability issues due to their oxidation from +2 to the toxicologically inactive +4 state [15,16]. Figure 2 shows some examples of lead-free perovskites and the most studied lead-based perovskites from the viewpoint of their bandgaps, which is one crucial factor of absorber materials for developing photovoltaic devices [17]. One must not ignore the fact that the suggested compositions contain elements that are either scarce, reliant on heavy metal production, or toxic [9].
The attempt to find lead-free materials with as good properties as the lead-containing perovskites led to a vast urge to explore the rich perovskite field which covers thousands of compounds demonstrating major potential in, for instance, energy storage devices [18] and optoelectronic [19,20], spintronic [21], and photocatalytic [5] applications. Perovskites are a great branch of materials, and the name denotes a certain crystal structure with an AMX3 stoichiometry, which is composed of MX6 corner-sharing octahedra bordering A-site cations. This review focuses on a subgroup of them, halide perovskite, and related low-dimensional lattices which comprise halide (X) octahedra that surround a positively charged metal ion (M). There are examples of M being an organic monovalent cation [22]. In the interstices of these octahedra, another positive ion (A) can be found. Constituent ion A can be either an inorganic ion, for instance, cesium, or a small organic cation like methyl ammonium ion; however, up to now, no systems with an organic–inorganic mixture for this site [5] have been reported. These building blocks can form several different dimensionalities and symmetries dependent on the constituting or vacant ions’ radii [23], valences, synthetic technique, additives, presence of chiral and racemic ligands, and external conditions [24] (see examples in Figure 3 based on the classification by Akkerman and Mana [25]). Akkerman and Mana presented an overview of the group’s breadth and complexity, classifying the representatives by their crystal structure, and identifying six subgroups [25].
Figure 3 shows not only numerous examples of perovskite structures but also solids that are not built up by the network of corner sharing octahedra due to a lack of stability or because of additives that cause derivation from an originally stable lattice. If the structure involves edge and/or face-sharing octahedra MX6, the literature may name them “perovskitoid” [26]. The research field of low-dimensional structures is closely related to the area of perovskites; one example to illustrate this is their application as scaffolds for perovskite growth for effective surface defect passivation in perovskite solar cells [23,26].
In this short review, we will generally concentrate on exploring lesser-researched lead-free structures and compositions in the perovskite field and related areas. Since tin halide perovskites and vacancy-ordered structures are highly researched as lead-free perovskites, their discussion is outside of the scope of this work, but a comprehensive review was recently published about those branches [27]. Our goal is to help in the rational material design process by sharing recent advancements in fully inorganic lead-free halide double perovskites and lower dimensionality hybrid perovskite-related or derived structures.

2. Extract of Roadmaps on Inorganic Lead-Free Halide Double Perovskites: Versatility, Properties, Development, and Challenges

2.1. Mapping the Stability of Halide Double Perovskites

One way to eliminate divalent lead from the crystal structure is to replace it with a merger of ions with an average charge of +2. This is the case in double perovskites (DPs) (elpasolites), in which this substitution occurs via having a + 1 and a + 3 charged central cation in halide octahedra alternating through the crystal structure, expressed by the A2MM′X6 fixed stoichiometric formula (Figure 4a shows the crystal structure, Figure 4b marks ion M with B) [28]. There are examples of achieving stable rock-salt-ordered DP structures where both M and M′ are divalent, but usually, these components rather form the stoichiometry of a simple perovskite with a disordered configuration, AM(II)M′(II)X3, due to the zero charge difference, as shown in Figure 4b [29]. Having a different order than the rock-salt arrangement by a couple of compositions with A2M(I)M(II)X5 stoichiometry has also been reported [30]. Vacancy-ordered A2M(IV)X6 DPs are not discussed here in detail [31].
There are thousands of combinations of elements that may result in DP stoichiometry; thus, screening them is quite a challenge. Despite the vast number of possible combinations, up until 2023, “only” around 350 halide DPs had been synthesized. Theoretical calculations suggest that around 600 further compounds may be synthesized since many of the combinations from the list of thousands are unstable [33].
In the perovskite field, the crystallographic stability of compounds is usually predicted by three tolerance factors that depend on the oxidation state of ion A (nA) and the ionic radii of A, M, M′, and X ions (rA, rM, and rX, respectively, where rM is the arithmetic mean of the two M-site radii in the DP lattices). These numbers are usually referred to the following names in the literature and can be calculated using the following equations:
-
Goldschmidt tolerance factor:
t = r A + r X 2 r M + r X
-
Octahedral factor:
μ = r M r X
-
Tolerance factor by Bartel et al. [34]:
τ = r X r M n A n A r A r M l n r A r M
To form the perovskite crystal structure, the following criteria must be met: Goldschmidt factor between 0.81 and 1.11, octahedral factor between 0.41 and 0.9, and τ below 4.18 to form a perovskite crystal structure [33,34]. Figure 5 shows the position of several well-known perovskites in the octahedral factor–tolerance factor space, indicating their crystallographic stability.
In rare instances, predictions based solely on geometrical factors can be inaccurate. One example is Cs2CuBiCl6, which is unstable but meets the crystallographic criteria. These cases suggest that it is necessary to analyze the thermodynamic processes involved in the formation and decomposition reactions [36]. These compounds can break down into binary, ternary, or quaternary salts. If the decomposition reactions are poorly understood, there may be contradictions regarding stability. To aid research and development in the field of materials science, an open-access database was created under the name Materials Project, which contains structural information of thousands of compounds [37]. Zhang et al. conducted a study using the Materials Project dataset to map potential DP compositions based on their stability and band gap (see Figure 6) [38]. The study found that as the atomic weight of A increased from Li to Cs, the thermodynamic stability also increased while decreasing X became heavier. As for M, the stability of the DP decreased in the column of alkali metals from Li to Cs, while M′ had less of an impact on stability [38]. The results of these Density Functional Theory (DFT) calculations and data from a plethora of sources were employed in developing machine learning models for predicting thermodynamical stability for lead-free halide perovskites; its general workflow is depicted in Figure 6 [39,40,41].

2.2. Electronic and Optical Characteristics and Application of Halide DPs

Numerous elpasolites have been synthesized and studied, and they are reported to have large bandgaps; their magnitude is usually ranging from 2 to 3.4 eV [5]. The nature of the bandgap for halide DPs may be tunable since both M and M′ contribute to the valence band maximum (VBM) and conduction band minimum (CBM) independently [33]. One way to classify these materials is to consider the valence electrons of ions M and M′. The bandgap is direct if both M and M′ have lone-pair electrons, or if both do not have lone-pairs and for vacancy-ordered DPs (Type I and Type III in Table 2). If only M or M′ have lone-pair electrons such as Cs2AgBiBr6, the bandgap is indirect (Type II in Table 2). In this case, the bonding orbitals of M(nd), M′(ns), and X(np) contribute to the VBM, while the CBM minimum is dominated by the antibonding orbitals M′(np) and X(np) [42]. Table 2 shows these groups with examples from experimental publications and proposed application areas [19].
Differently from the standard AMX3 perovskites, the lattice tilts and octahedral distortion caused by the altering of ion A have no significant effect on the bandgap in A2AgBiBr6 [43]. However, in general, the effect of the substitution of A in DPs is less studied. Since the X ions strongly influence the VBM, it would be one simple method to tune the band structure via replacing or intermixing different halogen ions in the crystal structure, but this strategy fails in a lot of cases due to stability issues, and the most studied elpasolite structures are chloride-based, and very few are bromide-based [33]. It has been demonstrated that a stable DP structure can be built with iodine in the case of Cs2NaBiI6, and this solid shows a bandgap of 1.66 eV [44]. One extensively studied 2D perovskite-related system with iodine octahedra building blocks is Cu2AgBiI6, which is proposed to be a candidate for indoor photovoltaic absorbers [45,46].
In addition to the comprising elements, the band structure could be significantly changed by the control of atomic arrangement via ordering parameters at the M-M′ sublattices, as illustrated in Figure 7 [47]. However, in most cases of DP compositions, the rock-salt order is thermodynamically favored due to the large charge difference between M and M′ and Pauling’s fourth rule. In rock-salt-ordered DPs, the high-valency +3 charged cations inside the octahedral halide cage are the furthest apart from each other, and due to energy or symmetry mismatch between the neighboring M(I) and M(III) frontier orbitals, this leads to (quasi)-0D electronic dimensionality. Therefore, the band edges are relatively localized [30]. Among these perovskites, the unique temperature-dependent light absorption and photoluminescent properties of published double perovskites are currently explained with different mechanisms; one of them is related to the question of atomic arrangement. The observed tendency is that lowering the temperature causes a blue shift of absorption onset and a red shift of emitted light. One of the proposed mechanisms behind this phenomenon is attributed to a local disordered phase, which could establish sub-bandgap emissive states (see Figure 7) [42,48].
One example of a non-rock-salt-ordered structure is Cs2Ag(I)Pd(II)Br5, which shows an indirect bandgap of 1.33 eV. Still, because of altered stoichiometry, strictly speaking, it does not belong to the group of perovskites [30]. On the other hand, if M′ is changed to a vacancy (see Figure 3 for the names of the branches of perovskites), resulting in low electronic dimensionality, some representatives may still exhibit dispersive bands and low bandgaps [5]. For instance, Cs2SnI6 has a relatively low bandgap and was proposed to be applied in thin-film transistors [49]. Its stable mixed-halide derivative, Cs2SnI3Br3, has been introduced in a dye-sensitized solar cell as a hole-transporting material, and the device reached 3.63% PCE [50].
Type III DP (Table 2) crystals have such symmetry conditions that certain optical transitions become symmetry-forbidden (inversion-symmetry-induced parity-forbidden) between CB and VB edges [19,43]. However, several compositions were published as host lattices from this group incorporating homo- and/or heterovalent dopants, which effectively modify the local crystal symmetry and crystal field intensity and result in radiative transition. For example, Figure 8a shows the dual-emission spectrum of Sb3+ and Bi3+ co-doped Cs2NaInCl6 microcrystals by Zhou et al. [51], and Figure 8b shows the photoluminescence quantum yield (PLQY) of blue emission by Sb5+-doped Cs2NaInCl6 single crystals (SC) by Liu et al. [52]. From the aspect of PLQY, some of these systems show remarkable values, for instance, a system of Cs2KInCl6:5%Sb3+ had 93% PLQY [53].
From an application standpoint, DPs have great potential in down-conversion LEDs; Table 3 summarizes numerous device structures featuring powder-based DP compositions for white light emission. In addition to WLEDs, certain DP compositions may be applied as NIR emitters, as was demonstrated by chromium-doped Cs2AgInCl6 crystals [54]. The current challenge of these microcrystalline DPs is to improve operational stability and to achieve a better understanding of charge injection, transport, and recombination [27].
To the best of our knowledge from a research perspective, spectroscopic studies of these systems date back to the 1970s, and until today, several questions have been raised and several energy transfer and quenching mechanisms have been proposed [51,53,56,64,65,66,67,68]. Building on one of these models, a structure–property descriptor was developed for one dopant, Sb3+ for machine learning algorithms, aiming at halide DP design [69]. Metal ion doping/alloying of the DP lattice, however, in addition to bandgap tuning and light-emissive behavior, may lead to other interesting properties, for instance, magnetic response [21]. However, the local environment of dopants/alloying elements might be different from the targeted [70,71]; thus, correct charge states and crystallographic positions of substituted metal ions need to be systematically investigated for halide DPs [72].
In DP nanocrystals (NCs), weak quantum confinement effects can be achieved if the Bohr radius is around 1 nm (e.g., 1.02 nm and 0.82 nm for Cs2AgSbCl6 and Cs2AgInCl6, respectively), giving similar electronic characteristics as for the 10 nm NCs compared to bulk materials [73,74]. Just like DPs in other forms (e.g., polycrystalline films), DP NCs have been universally applied for bright white emissions, solar cells/concentrators, photocatalysis, and X-ray imaging [74,75,76,77,78,79]. In addition, DP NCs share the basic properties of all NCs, like the advantage of being applicable for large-area white light emissions. The performances of recently developed DP NC-based devices in the abovementioned application areas are summarized in Table 4.
To summarize, so far, from the point of application, only 7% of reported lead-free perovskites were double perovskites used for solar cells (the PCE of a champion device with one of the most studied DPs, Cs2AgBiBr6, is 3.31%) [95]. Still, beyond photovoltaics, the potential is significantly larger. A total of 51% of reported lead-free halide DPs were used for developing LEDs, and 45% were used for photocatalysts [5,96]. A couple of representatives have also been proposed for NIR [70] and X-ray [19,97,98] detection. In the latter application area, an excellent review article was published in 2023 comparing the performances of different candidates for ionizing radiation detection [99]. In addition to these fields, certain DP systems were proposed to be employed in memristors [100], and spintronics [21,72], and be exploited for their thermochromic [101,102] behavior in thermometers, temperature sensors, or smart windows.

2.3. Manifesto of DPs: Synthetic Techniques and Technological Challenges

One of the advantageous properties of DPs is good stability under ambient conditions, and they may be synthesized in several forms. Figure 9 shows the three most used methods for the synthesis of these complex compositions: hydrothermal, supersaturation, and the solvent-free solid-state method [96]. Less common ways, but also solvent-free methods, include mechanochemical synthesis [103,104], as well as the melt-based Bridgman technique [105]. The hydrothermal crystal growth of DPs occurs in an aqueous mixture of inorganic salts in a sealed autoclave reactor, for which the system heats up to the reaction temperature and then slowly cools down. In the case of the supersaturation technique, the precursor is prepared at ambient conditions and heated up to a generally lower temperature than in the previous technique. The heat-up is subsequently followed by a cool-down step, which results in microcrystals [96]. The solid-state reaction mixture is sealed in a fused silica ampule under low pressure and heated up to high temperatures without solvents [106]. The Bridgman technique also occurs in a sealed silica tube filled with the precursor, but the tube is moved along the temperature gradient [107]. The mechanosynthesis employs a high-energy ball mill to obtain polycrystalline powders, and except for the intrinsic heat generated by the ball milling process itself, the DPs may be formed at room temperature [103]. Figure 9b shows publication statics on different synthetic routes for several DPs [15].
For the growth of SC DPs, either the hydrothermal reaction, solventless molten-salt medium synthesis, or the Bridgman method is applied. Due to random nucleation and isotropous growth, the hydrothermal method is difficult to scale up for large-scale thin-film production, but via surface engineering, remarkable results have been achieved and published [108]. The method depends on the solubility of inorganic salts in water at different temperatures and vapor pressures, and the formation is complex and relatively slow [109]. However, this facile method can also be applied for the synthesis of complicated composite structures like the DP Cs2AgBiBr6 supported on nitrogen-doped carbon materials for hydrogen evolution [110]. The family of solid-state techniques includes the so-called vapor transport method, which results in high-quality SCs, but it is a less-explored technique [99]. The Bridgman technique was introduced for the fabrication of high-quality perovskite ingots on a large scale [111]. The drawbacks of this technology are high energy demand, mechanical stress, high density of grain boundaries at the surface in close contact with the silica tube, and the breakage of SCs due to subsequent mechanical cutting [99]. For further readings, recently a comprehensive review has been published discussing different SC growth theories and techniques and showing countless examples from lead halide perovskites [107].
Another huge challenge of the development of lead-free halide DP-based devices is that they can be synthesized from the acid solution, but this is not suitable for thin-film device fabrication. In comparison to lead perovskites, they have higher formation temperatures and low solubility, and halide DPs show a quick crystallization rate, which complicates the uniform film formation. This may be circumvented by thermal evaporation, antisolvent engineering techniques, and post-annealing [33,43,112]. Moreover, different from the formation of films or SCs, DP NCs are usually synthesized with the assistance of ligands. Thanks to ligands, which reduce surface energy, doped DP nanocrystals obtain enhanced stability compared to polycrystalline films or SCs [113,114].
There are mainly two synthesis methods for DP nanocrystals: the hot injection (HT) method and the ligand-assisted reprecipitation (LARP) method. Synthesizing Cs2AgBiBr6 and Cs2AgBiCl6 by HT was first reported in 2018 by two different groups, and their approaches are depicted in Figure 10 [114,115]. Sidney E. Creutz et al. injected trimethylsilyl halide (TMSX) into octadecene (ODE) with the precursor metal acetates and ligands (oleic acid (OA) and oleylamine (OAm)) at 140 °C and achieved well-defined cubic Cs2AgBiBr6 and Cs2AgBiCl6 NCs with sizes of around 8–9 nm [115]. Lei Zhou et al. injected Cs-oleate into a high-boiling-point solution of BiBr3 and AgNO3 at 200 °C and achieved Cs2AgBiBr6 NCs with a size of 9.5 nm in cubic shape [114]. According to the report of Bekenstein et al., the two methods yield similar nanocrystals [116]. All in all, HT has been broadly applied for synthesizing DP NCs since 2018.
LARP was first reported by Yang et al. for the synthesis of DP nanocrystals with the following procedure: all the precursor materials were dissolved in a good solvent and injected into an antisolvent to form DP NCs, which turned out to be less regular than that achieved by the HT method [81]. They also found that the addition of OA would enhance the photoluminescence of DP NCs by 100 times through defect passivation, and afterward, OA is used in the LARP synthesis by default [117].
In the past five years, a lot of efforts have been made to dope/alloy metal ions in DP NCs to tune their optical properties [55,77,80,81,82,86,118,119]. Among these, indirect to direct bandgap tuning is achieved by In3+ doping in the case of Cs2AgInxBi1-xCl6 NCs (x = 0.75 and 0.9) [81]. In addition, Han’s group has greatly improved the photoluminescent properties of DP nanocrystals by doping: after doping with Sb2+, green emission with unity PLQY is achieved for Cs2KInCl6 DP NCs, and further Mn2+ doping leads to white light emission with PLQY of 87% due to efficient energy transfer [77].
Besides the abovementioned doping/alloying method, ligand engineering in Cs2NaInCl6/Cs2AgInCl6 synthesis, e.g., by trioctylphosphine (TOP), has also been studied to improve photoluminescence and colloidal stability [83,87,120]. Furthermore, an interesting study on morphology control in DP NCs was also developed by Liu et al. Cs2AgBiX6 (X = Cl, Br, I) two-dimensional nanoplatelets were first reported, which showed better catalytic performance for CO2 photoreduction than their nanocube counterpart [74].
The synthesis of nanomaterials with quaternary elements is complicated and requires careful compositional tuning to ensure product purity. Despite this, a lot of achievements have been reached in DP NCs during the last five years. In the future, the synthesis method is expected to be broadened, e.g., the passivation of halide defects at the surface [121], and morphology engineering, to improve the stability for various applications of DP NCs.

3. Low-Dimensional Lead-Free Halide Perovskite Derivatives

Not featuring the corner-sharing network of octahedra, strictly speaking, the following classes of materials do not belong to perovskites. However, some representatives have similar crystallographic building blocks that otherwise cannot form stable perovskite structures, and some are derived from perovskite structures; thus, one can say that research on them is closely related to the perovskite field. Here, we include recent advances in the field of lead-free hybrid organic–inorganic 2D (fully inorganic 2D structures are out of the scope of this work [23]) and 0D perovskite-related material.

3.1. Hybrid Organic–Inorganic 2D Layered Halide-Perovskite-Related Structures

Hybrid organic–inorganic layered perovskites have been explored since the 1990s and are a promising class of material with potential applications as a surface passivation layer in 3D metal lead-based and lead-free halide photoabsorbers [122,123,124], thermoelectric energy conversion [125], photodetectors [126], magnets [127], memories [128,129], spintronics [130], energy storage, and perhaps many more [131]. Their versatility lies in their rich compositional diversity and structural tunability.
If ion A in the perovskite composition is fully or partially changed to an organic cation that has one or two ionically interacting terminal functional groups (most frequently amines) and is larger than 2.6 Å, the 3D crystal structure may reduce along the (001) plane into slabs of 2D perovskite layers separated by organic layers (there are examples of (110) and (111) oriented structures; they are not discussed here) [23,132,133]. The organic spacers containing hydrophobic alkyl chains ensure protection against moisture and oxygen, thus being more stable than their 3D counterparts, but anisotropic 2D crystals are challenging to synthesize in the form of thin films [131,134,135]. These structures can be described with (S)xAn−1MnX3n+1 stoichiometry, where S is the templating organic cation spacer (if monofunctional, x = 2, if bifunctional, x = 1). If not fully replaced by S, A is the cation incorporated into the octahedral cavities [131,132]. Figure 11a aids in visualizing the dimensional reduction and consequent perovskite lattice distortion through the example of the incorporation of butylammonium (BA) ions in lead-free halide DP Cs2AgBiBr6 [132]. The thickness of the inorganic sheet (number of {MX6} slabs) may be one grouping factor since it determines the phase stability and optoelectronic properties greatly, for instance, the indirect–direct character of bandgaps may be modified by dimensional confinement [131,132].
Figure 11b shows another way of classifying these compounds according to the relative crystallographic phases of the inorganic perovskite layers: in the case of Ruddlesden-Popper (RP) perovskites, the adjacent perovskite layers are displaced by half a crystal unit cell along both in-plane directions (commonly involving monofunctional organic bilayers), while Dion–Jacobson (DJ) phases are defined by no in-plane displacements (mostly having bifunctional organic layers in their structure) [131]. In the case of DJ DP derivatives, due to the alternating cations, the stacking patterns can be formed in two ways, i.e., [0,0] or [1/2,0] [136]. From the aspect of dimensional tuning, there is a third group of structures called the alternating cation interlayer (ACI), where the organic interlayer space consists of different cations in an alternating pattern (guanidinium ion and methylammonium ions are reported to be such cation combinations that may form this specific structure), for which the structure leads to higher crystal symmetry and a narrower optical gap in comparison to the RP perovskite derivatives [131,133,137,138].
Two-dimensional layered structures may allow for ionic compositions in inorganic slabs that otherwise do not form phase pure 3D crystals. One example is the DP combination of Ag+ and Sb3+ in RP (BA)2Ag0.5Sb0.5Br4 [135,139]. Molybdenum can be preserved in oxidation state +3 and combined with monovalent silver in (PPDA)2AgIMoIIICl8 [136] (PPDA = para-phenylenediammonium). Despite the rarity of silver in octahedral iodine coordination, (AE2T)2AgBiI8 [140] (AE2T = 5,5′-diylbis(aminoethyl)-[2,2″-bitiophene]) is synthesizable. A third way of classifying these crystals is to group them according to the constituting inorganic ions, which is shown in Figure 12: the DP derivative (S)4M+M′3+X8 and the (S)2M2+X4-type with Group IVA-type metals (Sn2+ and Ge2+) and with other divalent metal ions (such as Cu2+, Fe2+, Cr2+, etc.) [23].
By separating and connecting inorganic layers, organic spacers are important for determining the system’s functionality. Several different organic spacers have been introduced to these structures, which may even add to the functionality of the corresponding 3D analogs by being inherently electroactive, photoactive, chiral, or mechanochromic [131]. The incorporation of spacers that are insulating leads to natural quantum well behavior, where charges are confined to the inorganic layer [131]. Thus, one grouping point of these low-dimensional structures is the thickness of the inorganic sublattice with which the optoelectronic properties are gradually changing [132]. Using electroactive ligands such as functionalized thiophenes can modulate the electronic structure [131]. The superior optoelectronic characteristics of perovskites may be combined with chirality in these structures (3D perovskites likely remain in the theoretical development stage). If chiral spacers separate the perovskite layers, which leads to phenomena like circular dichroism, nonlinear optics or spin-related effects may be seen [141]. One example is shown in Figure 13, where enantiomeric 2D hybrid copper-based perovskites, (R-MPEA)2CuCl4 and (S-MPEA)2CuCl4, possess chirality confirmed by circular dichroism measurements (MPEA = β-methylphenethylamine) [130].
As proposed in this chapter’s introduction, these 2D structures have significant potential in wide application areas. From the materials design perspective, it is expected to be easier to engineer multifunctional structures based on low-dimensional structures employing A cations with functionalities [23]. Screening these compounds due to their complexity and diversity of their composition is challenging. Li et al. published a DFT-based hierarchical progressive screening method on 3D bulks to compare derived layered structures from the point of view of stabilities and electronic properties [142]. They found that (CH2)8(NH3)2Csn−1SnnBr3n+1 stoichiometry covers a class of material predicted to have high PCE, small carrier effective masses, and good stability and appropriate bandgaps [142]. To aid the discovery of new materials belonging to this group, Chen et al. screened 1,4-butanediamine-templated (BDA) DJ DP for solar application [143]. They published a map of decomposition enthalpies that were calculated considering the decomposition pathways from the desired compound into the corresponding common binary decomposition products. In general, they concluded that all combinations of (BDA)-templated halide structures Na+/K+/Rb+/Cu+/Ag+/Au+ and Bi3+/In3+/Sb3+ all are thermodynamically stable [143].

3.2. Electronically Zero-Dimensional and Perovskite-Inspired Crystal Structures

3.2.1. 0D Anti-Perovskites

Most 0D materials cannot be considered full members of the perovskite classification since the metal halide octahedral corner sharing is absent (see perovskite-inspired structures). However, 0D anti-perovskites do exist. The composition of anti-perovskites can be represented in several ways, one of them is X3BA, where X is a monovalent cation, A is a monovalent anion, and B is a divalent anion, or as [MXx]XA3, where M is a metal, X is a halogen, and A is a cation (see Figure 14a). Another way to represent anti-perovskites is A3BX, but this nomenclature is not so common in the optoelectronic field. The structure shows promising properties for optoelectronics [144,145]. Manganese (Mn)-based organic or inorganic metal halides are typical examples of when a 0D anti-perovskite structure can be favorable. Coupling between the Mn-Mn sites in an ordinary 3D perovskite quenches the emission, but in a 0D anti-perovskite, the distance between the Mn and Mn sites increases, resulting in limited coupling. Also, the confinement effect is enhanced by the structure and will improve the emission. Panda et al. synthesized a 0D anti-perovskite, (Piperidinium)3Cl[MnCl4] (see Figure 14b), with a PLQY of 54.5% [144], and Yan et al. synthesized six different Mn-cesium (Cs) halides, all with a PLQY over 80% [146] (see Figure 14a). The structure consists of octahedral complexes made up of a halide surrounded by organic monovalent cations, while the incorporated divalent anion consists of a tetrahedrally coordinated metal halide. This structure is represented as [MX4]XA3, where A is an alkali metal (I); M is a transition metal(II); and X represents Cl, Br, or I. In an anti-perovskite, XA3 builds up the classical corner sharing octahedra, while [MX4] represents the position in between the octahedra. However, not all [MX4]XA3 materials show an anti-perovskite structure. By comparing the structure of 22 different [MX4]XA3 materials, Yan et al. concluded that the ratio of half the tetrahedral edge length and the sum of the ionic radius of the A and X atoms had to be between 0.544 and 0.644 for 0D anti-perovskites [146].

3.2.2. Perovskite-Inspired Structures

Definition of Different Crystals

0D inorganic–organic hybrid metal halides, as well as fully inorganic metal halides (called 0D metal halides in the rest of the text), consist of the same type of building blocks as 3D metal halide perovskites used in optoelectronic devices such as solar cells and LEDs. However, for the structure to be considered electronically 0D, the octahedra in the perovskite must be separated to limit the metal–metal coupling. Separated octahedra means no corner sharing, and in this way, the perovskite structure is broken and should theoretically not be called perovskites (see Figure 15b). Only increasing the size of the A site molecule/atom in a 3D perovskite until the octahedra is separated results in the composition AMX6, where A is a large monovalent cation, M is a 5+ metal such as Bi5+, Sb5+, or V5+, and X is a halogen [147]. The octahedra can also be separated by 2, 3, or 4 A site monovalent molecules/atoms per unit cell, resulting in the compositions A2MX6 (M = +4) [148], A3MX6 (M = +3) [149], and A4MX6 (M = +2) [150].
So far, we have only discussed metal halides containing octahedral polyhedrons, but also materials containing tetrahedral or square pyramidal [151], or a mixture of tetrahedral and trigonal bipyramidal–polyhedral, can be perovskite-inspired materials. For example, tetrahedral-containing or seesaw-shaped materials with a divalent metal M result in the composition A2MX4 (M = 2+) [152]. Additionally, the A site molecule does not necessarily have to be monovalent, once again changing the A-M-X composition.
As a general conclusion, 0D metal halides consist of anionic metal halide polyhedrons spaced by large cations [153]. The general composition can be described as AiMjXk, a variety of combinations are included, and many of them have been called 0D perovskites, although the structures deviate a lot from the perovskite structure.
Another quite common feature in 0D metal halides is the incorporation of water molecules, sometimes becoming a part of the polyhedron, which then affects the coordination chemistry [154]. By divagating slightly from the 0D structure, we will find structures where two or more metal halide polyhedrons will cluster up, showing corner, edge, or face sharing. This structure can sometimes be called quasi-zero-dimensional (q-0D) structures. The terminology around q-0D is not fully established, and sometimes the quasi term can be used to point out that the type of 0D materials that have been discussed in this section are not 0D in the same sense as quantum dots since they can form macroscopically sized SCs. The 0D structure for this for the materials discussed here is only 0D in electronic matter [155].

Properties

As seen in the sections above, there are many kinds of 0D metal halides, and all of them will not have the same properties. However, they will be more stable than their 3D counterparts since the large organic (or inorganic) cations will act as a protective layer surrounding the metal halide polyhedrons. Electronically low-dimensional materials often have hampered charge transfer and independent luminescent centers that are relatively insensitive to long-distance chemical morphology, resulting in rather stable luminescence. They also have a stronger quantum confinement effect, resulting in closely localized excitons with high exciton binding energy, leading to high PLQY [156,157].
The light emission in 0D metal halides can originate from self-trapped excitons (STE), d-d transitions, d-f transitions, metal–ligand charge transfers (MLCT), defect states, as well as mixtures of these and other emission routes [158,159]. STEs are formed by local lattice distortion induced by the excited state, generating a distortion field trapping the exciton that induced the distortion in the first place (self-trapped). The photoluminescent emission energy can be expressed as EPL = Eg − Eb − Est − Ed, where Eg is the bandgap energy, Eb is the exciton binding energy, Est is the self-trapping energy, and Ed is the lattice deformation energy. Jahn–Teller distortion is the most common cause of self-trapping in perovskite/perovskite-like structures, while the formation of Vk centers is the most common cause in alkali halides. The formation of STEs is increased in low-dimensional crystal structures, making it a common origin for emission in 0D metal halides. Emission by STE gives a broad PL spectrum and large stokes shifts (example: (C4N2H14X)4SnX6) (see Figure 15b,c) [160,161,162,163]. d-d transitions can mainly be seen from transition metals surrounded by ligands, and emission energy originates from the crystal field splitting of the otherwise degenerated d orbitals. d-d transitions with ligand field dependency (changes in orbital occupancy) generate relatively broad emissions due to vibronic broadening, while transitions that are close to being independent of the ligand field (near unchanged electron distribution) reveal a much narrower emission (see Figure 15a) [164,165,166]. The emission becomes stronger when dipole–dipole coupling can be avoided between the metal centers, and therefore the strongest d-d emissions can be found within electronically 0D materials or from doped materials [167]. One of the most typical examples of d-d transitions in 0D metal halides is seen in manganese-based materials, where the octahedral polyhedrons emit weaker than the tetrahedral since the transitions are Laporte-forbidden (d-d and centrosymmetric), as well as oppose the spin selection rule. In contrast, the tetrahedral polyhedrons only oppose the spin selection rule since the shape is not centrosymmetric; nevertheless, both have forbidden transitions, which causes long lifetimes.
Figure 15. (a) (i) Gives an idea of the relative emission width of the different transition mechanisms seen in (iiv). (ii) Metal–ligand charge transfer. (iii) d-d transitions with ligand field dependence and (iv) with ligand field independence. (v) f-f transitions. Reprinted (adapted) with permission from {Chem. Rev. 2023, 123, 12, 7890–7952}. Copyright {2023} American Chemical Society [165]. (b) SnBr64− octahedra surrounded by C4N2H14Br+ ligands, reproduced from Chem. Sci., 2018, 9, 586–593 with permission from the Royal Society of Chemistry [163]. (c) Schematic energy curves of STE formation. GS = ground state; Eg = the bandgap; Eb = exciton binding energy; Est = self-trapping energy; FC = free carrier; FE = free excitons; S = Huang–Rhys factor; Ed = lattice distortion energy. Reproduced from [168]. © 2022 Wiley-VCH GmbH.
Figure 15. (a) (i) Gives an idea of the relative emission width of the different transition mechanisms seen in (iiv). (ii) Metal–ligand charge transfer. (iii) d-d transitions with ligand field dependence and (iv) with ligand field independence. (v) f-f transitions. Reprinted (adapted) with permission from {Chem. Rev. 2023, 123, 12, 7890–7952}. Copyright {2023} American Chemical Society [165]. (b) SnBr64− octahedra surrounded by C4N2H14Br+ ligands, reproduced from Chem. Sci., 2018, 9, 586–593 with permission from the Royal Society of Chemistry [163]. (c) Schematic energy curves of STE formation. GS = ground state; Eg = the bandgap; Eb = exciton binding energy; Est = self-trapping energy; FC = free carrier; FE = free excitons; S = Huang–Rhys factor; Ed = lattice distortion energy. Reproduced from [168]. © 2022 Wiley-VCH GmbH.
Materials 17 00491 g015
The d-f transitions are much faster than the d-d and f-f transitions since they are parity-allowed (example: Cs3CeI6) [169]. Charge transfer relaxation gives broad emission due to geometric distortions, as seen for the metal-to-ligand charge transfers (MLCTs) in Figure 15a(i,ii). MLCTs are not common in 0D metal halides, but there are some materials with ligand-to-metal charge transfer (LMCT)-based emission (example: (PTMA)3Cu3I6) [170] and MLCT-assisted emission (example: [(C6H5)4P]2SbCl5) [171].
As mentioned in the beginning of this section, the emission could also be from defect states. As an example, Ran et al. showed that the emission from (C6H8N)6InBr9 SCs originates from Br vacancies; however, owing to a strong thermal quenching effect, the material showed a low PLQY of only 2.7%. Yet when doping the material with antimony, creating (SbBr6)3− octahedra, the PLQY increased to 71.8%. The increased PLQY originates from STE recombination localized on the (SbBr6)3− octahedra and is enhanced by the charge transfer from the defects [172].

Applications

As for the 3D perovskites, 0D metal halides can be employed as the active layer in LEDs, but they are generally not suitable for solar cells due to high exciton binding energies and electronic confinements. There are several 0D material systems synthesized for LED applications, but one of the major difficulties with 0D-based electrically driven LEDs is low conductivity due to bulky spacing molecules, but with good material engineering and doping, the conductivity can be increased [173]. 0D materials also usually have large Stoke shifts, which is preferable for LEDs since this reduces reabsorption (see Table 5) [174].
Another more common application is to use them as phosphor materials in LEDs, i.e., the phosphor covers the lightbulb or diode surface of a high-energy emissive (UV or blue) diode that can excite the phosphor, which emits a lower energy wavelength. By mixing phosphors with different emission wavelengths, a white light emission can be achieved. 0D materials are often more stable than their 3D counterpart, while showing a high PLQY value, making them suitable as phosphorus materials (see Table 6). Phosphor materials can also be used in applications such as encoding and anti-counterfeiting or re-printable paper. Properties such as quenched emission in the presence of water or some other solvent, as well as dual or triple emission colors that emit stronger/weaker depending on the excitation wavelength, can be utilized for these applications [161,182,183]. X/β-ray scintillators made from 0D metal halides are an emerging field with promising results [184]. There are also reports of fluorescent sensors used for detecting heat, moisture, or specific solvents (see Table 6) [185,186].

Synthesis

One common way to synthesize perovskites as well as perovskite-like materials is by different solution-based methods. The A site molecule/atom or A site halogenated salt is dissolved together with B site halogenated salt or oxide. Sometimes, HI, HBr or HCl is added for excess halogens to enhance the reaction. When the solvent has evaporated, cooled down from a higher temperature (see Section 2.3), or an antisolvent has been diffused into the solvent, the material will precipitate and crystallize [212].
0D metal halides can also be synthesized using mechanochemical techniques such as ball milling or grinding using a mortar; this will create a fine powder of the 0D material without using any solvents or by using only a small amount of solvent to enhance the reaction. The properties of the solution-grown materials and the mechanically grown materials can differ slightly [212].
An autoclave can be used to prepare the synthesis under thermally activated pressure. The chosen molar ratio of the A site and B site molecule/atom is added to the autoclave together with HI, HBr, or HCl depending on the preferred halogen. The B site precursor can be in the form of an oxide or halogenated. This technique can generate high-quality SCs [209,213]. Some materials are synthesized to become both nanocrystals and electronically 0D at the same time. For this type of material, hot injection synthesis is common (see Section 2.3).
For thin-film fabrication, spin coating is the most utilized method in research laboratories. Nonetheless, other solution-based techniques such as spray, dip, slot-die, bar, or blade coating, as well as gravure, screen, or inkjet printing, are also used (see Figure 16) [214]. Additionally, evaporation methods such as physical vapor deposition (PVD) and chemical vapor deposition (CVD) are applicable [169]. Spin coating is an easy and low-cost option and a well-operated coating results in high-quality thin films. The two main problems with spin coating are scalability and material waste; only 2–5% of the solution stays on the substrates. Other solution-based methods are also relatively cheap and more suitable for up-scaling, but so far, the best-performing devices are made by spin coating. The film quality from spray coatings is dependent on the size of the droplets, and it is hard to make the films uniform and with a precise thickness. An advantage of spray coating is the possibility of coating substrates with complex shapes [215,216]. Dip coating gives high crystallinity, but also relatively many pinholes and high surface roughness [217].
If spin coating is the most common coating method on a lab scale, slot-die coatings incorporated with roll-to-roll (R2R) functions are the most common for the industrial manufacturing of perovskite solar cells as well as organic photovoltaics. Slot-die coatings are material-efficient and can easily be made on flexible substrates, plus with the roll-to-roll function, the deposition becomes easily adapted to automatization [218]. Blade coating and bar coating are similar techniques to slot-die coating but with less ink control. The most important factor to receive good-quality films, while using any of the three techniques, is to control the evaporation [219]. Inkjet printing works as a common household printer, and one of the advantages of inkjet printing is high-resolution printing without the usage of masks; another is that the nozzle is never in contact with the substrate [220]. One disadvantage is the formation of the coffee ring effect which originated from increased evaporation on the droplet edge due to curvature and induced flow within the droplet [221]. This effect generates uneven coatings but can be avoided by controlling the evaporation speed concerning the diffusion rate of the solute [222]. Gravure printing is expensive due to equipment, has limited resolution, and is time-consuming, but it has demonstrated itself to be a method to consider by generating PCE values above 19.10%. Most of the abovementioned solution-based techniques have received PCE values higher than 19%: blade coating at 21.09% (2021), slot-die coating at 20.80% (2021), spray coating at 19.40% (2020), and inject printing at 21.60% (2021) [223]. Bar coating and screen printing are lagging behind with PCEs of 17.53% (2021) [224] and 15.89% (2020), respectively [223]. All these techniques are also possible to adopt to R2R [225,226].
In PVD, the material is evaporated in a high-vacuum chamber where it then sublimates on the substrate. The evaporation could be performed from one single source containing the full material or from several sources that can be evaporated sequentially or simultaneously. In CVD, gaseous phases are introduced into the chamber and react on the substrate. The precursors can be evaporated to decomposition to later form the same material again by chemical reactions on the surface, or two or several gaseous precursors react to form a new material. CVD is more complicated than PVD since reactions must form, and many atoms/molecules are needed for a high probability of the reaction. PVD is more expensive to operate than solution-based methods, especially due to the high vacuum needed for good depositions. Advantages are an even film thickness, good film quality, good thickness control, and the avoidance of harmful solvents [227]. Without trapped solvents, one source for generating defects is eliminated. Additionally, there are no solvents to dissolve previous layers.
Depending on the application, as well as the material, different deposition techniques can be preferred. While pinholes are very bad for optoelectronic devices, they do not matter as much for phosphorus materials. And, while the active layers in solar cells preferably would be thin, the layers for scintillators need to be much thicker. Some materials might be sensitive to heat, while others are sensitive to solvents. And sometimes a combination of coating techniques might be the best alternative.

4. Conclusions

Concerns have been raised regarding the toxicity and environmental impact of lead-based perovskite solar cells. Therefore, as a follow-up on explosive perovskite research, efforts have been made to replace lead while preserving the advantageous properties of this structure. The different approaches led to the expansion of the field to structures not strictly meeting the definition of perovskites but still showing some structural and characteristic similarities. In this review, we aim to classify the relatively less-researched examples of synthetic halide perovskites and related structures. In addition, we briefly introduce the mapping strategies, general optoelectronic properties, synthetic techniques, and possible application areas of uncharted classes, i.e., halide double perovskites, hybrid organic–inorganic layered 2D perovskite derivatives, and electronically zero-dimensional structures. Examples of recent advances in the classes presented indicate that more explorative research is needed to find new compositions, and we hope this viewpoint aids in the rational design of these compounds.

Author Contributions

Conceptualization, A.D. and X.L., writing—review and editing, A.D., J.M. and M.C.; project administration, F.G. and M.F.; funding acquisition, F.G. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Knut and Alice Wallenberg Foundation (Dnr. KAW 2019.0082) and by the Swedish Government Strategic Research Area in Materials Science on Functional Materials at Linköping University (Faculty Grant SFO Mat LiU No. 2009 00971).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts 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. O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  3. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. [Google Scholar] [CrossRef] [PubMed]
  4. Best Research-Cell Efficiency Chart. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 8 November 2023).
  5. Jiang, P.; Acharya, D.; Volonakis, G.; Zacharias, M.; Kepenekian, M.; Pedesseau, L.; Katan, C.; Even, J. Pb-Free Halide Perovskites for Solar Cells, Light-Emitting Diodes, and Photocatalysts. APL Mater. 2022, 10, 060902. [Google Scholar] [CrossRef]
  6. Argillander, J.; Alarcón, A.; Bao, C.; Kuang, C.; Lima, G.; Gao, F.; Xavier, G.B. Quantum Random Number Generation Based on a Perovskite Light Emitting Diode. Commun. Phys. 2023, 6, 157. [Google Scholar] [CrossRef]
  7. Xiao, Z.; Song, Z.; Yan, Y. From Lead Halide Perovskites to Lead-Free Metal Halide Perovskites and Perovskite Derivatives. Adv. Mater. 2019, 31, 1803792. [Google Scholar] [CrossRef] [PubMed]
  8. He, B.; Wang, C.; Li, J.; Su, Z.; Xing, G.; Gao, X.; Chen, S. In Situ and Operando Characterization Techniques in Stability Study of Perovskite-Based Devices. Nanomaterials 2023, 13, 1983. [Google Scholar] [CrossRef] [PubMed]
  9. Schileo, G.; Grancini, G. Lead or No Lead? Availability, Toxicity, Sustainability and Environmental Impact of Lead-Free Perovskite Solar Cells. J. Mater. Chem. C 2021, 9, 67–76. [Google Scholar] [CrossRef]
  10. Sofia, S.E.; Wang, H.; Bruno, A.; Cruz-Campa, J.L.; Buonassisi, T.; Peters, I.M. Roadmap for Cost-Effective, Commercially-Viable Perovskite Silicon Tandems for the Current and Future PV Market. Sustain. Energy Fuels 2020, 4, 852–862. [Google Scholar] [CrossRef]
  11. Vidal, R.; Alberola-Borràs, J.; Sánchez-Pantoja, N.; Mora-Seró, I. Comparison of Perovskite Solar Cells with Other Photovoltaics Technologies from the Point of View of Life Cycle Assessment. Adv. Energy Sustain. Res. 2021, 2, 2000088. [Google Scholar] [CrossRef]
  12. 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]
  13. Kumar, A.; Kumar, A.; Cabral-Pinto, M.; Chaturvedi, A.K.; Shabnam, A.A.; Subrahmanyam, G.; Mondal, R.; Gupta, D.K.; Malyan, S.K.; Kumar, S.S.; et al. Lead Toxicity: Health Hazards, Influence on Food Chain, and Sustainable Remediation Approaches. Int. J. Environ. Res. Public. Health 2020, 17, 2179. [Google Scholar] [CrossRef]
  14. Zhang, J.; Gao, X.; Deng, Y.; Li, B.; Yuan, C. Life Cycle Assessment of Titania Perovskite Solar Cell Technology for Sustainable Design and Manufacturing. ChemSusChem 2015, 8, 3882–3891. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, Y.; Cruse, K.; Abdelsamie, M.; Ceder, G.; Sutter-Fella, C.M. Synthetic Approaches for Thin-Film Halide Double Perovskites. Matter 2021, 4, 1801–1831. [Google Scholar] [CrossRef]
  16. Abate, A. Perovskite Solar Cells Go Lead Free. Joule 2017, 1, 659–664. [Google Scholar] [CrossRef]
  17. Ke, W.; Kanatzidis, M.G. Prospects for Low-Toxicity Lead-Free Perovskite Solar Cells. Nat. Commun. 2019, 10, 965. [Google Scholar] [CrossRef] [PubMed]
  18. Ben Bechir, M.; Dhaou, M.H.; Altrifi, S.M. Photoluminescence, Raman and Photosensitive Dielectric Properties of Lead-Free Antimony-Based Cs3Sb2Br9 Single Crystals for Energy Storage Devices. Mater. Res. Bull. 2023, 167, 112381. [Google Scholar] [CrossRef]
  19. Zhao, X.-G.; Yang, D.; Ren, J.-C.; Sun, Y.; Xiao, Z.; Zhang, L. Rational Design of Halide Double Perovskites for Optoelectronic Applications. Joule 2018, 2, 1662–1673. [Google Scholar] [CrossRef]
  20. 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] [PubMed]
  21. Ning, W.; Bao, J.; Puttisong, Y.; Moro, F.; Kobera, L.; Shimono, S.; Wang, L.; Ji, F.; Cuartero, M.; Kawaguchi, S.; et al. Magnetizing Lead-Free Halide Double Perovskites. Sci. Adv. 2020, 6, eabb5381. [Google Scholar] [CrossRef]
  22. Ye, H.-Y.; Tang, Y.-Y.; Li, P.-F.; Liao, W.-Q.; Gao, J.-X.; Hua, X.-N.; Cai, H.; Shi, P.-P.; You, Y.-M.; Xiong, R.-G. Metal-Free Three-Dimensional Perovskite Ferroelectrics. Science 2018, 361, 151–155. [Google Scholar] [CrossRef]
  23. Ning, W.; Gao, F. Structural and Functional Diversity in Lead-Free Halide Perovskite Materials. Adv. Mater. 2019, 31, 1900326. [Google Scholar] [CrossRef] [PubMed]
  24. Morana, M.; Malavasi, L. Pressure Effects on Lead-Free Metal Halide Perovskites: A Route to Design Optimized Materials for Photovoltaics. Sol. RRL 2021, 5, 2100550. [Google Scholar] [CrossRef]
  25. Akkerman, Q.A.; Manna, L. What Defines a Halide Perovskite? ACS Energy Lett. 2020, 5, 604–610. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, J.; Yang, Y.; Dong, H.; Li, J.; Zhu, X.; Xu, J.; Pan, F.; Yuan, F.; Dai, J.; Jiao, B.; et al. Highly Efficient and Stable Perovskite Solar Cells Enabled by Low-Dimensional Perovskitoids. Sci. Adv. 2022, 8, eabk2722. [Google Scholar] [CrossRef] [PubMed]
  27. López-Fernández, I.; Valli, D.; Wang, C.; Samanta, S.; Okamoto, T.; Huang, Y.; Sun, K.; Liu, Y.; Chirvony, V.S.; Patra, A.; et al. Lead-Free Halide Perovskite Materials and Optoelectronic Devices: Progress and Prospective. Adv. Funct. Mater. 2023, 2307896. [Google Scholar] [CrossRef]
  28. Bhaumik, S.; Ray, S.; Batabyal, S.K. Recent Advances of Lead-Free Metal Halide Perovskite Single Crystals and Nanocrystals: Synthesis, Crystal Structure, Optical Properties, and Their Diverse Applications. Mater. Today Chem. 2020, 18, 100363. [Google Scholar] [CrossRef]
  29. Ji, G.; Xiao, Z. Jahn−Teller Distortion-Stabilized Halide Double Perovskites with Unusual Rock-Salt-Type Ordering of Divalent B-Site Cations. Chem. Mater. 2022, 34, 8207–8212. [Google Scholar] [CrossRef]
  30. Ji, G.; Han, C.; Hu, S.; Fu, P.; Chen, X.; Guo, J.; Tang, J.; Xiao, Z. B-Site Columnar-Ordered Halide Double Perovskites: Theoretical Design and Experimental Verification. J. Am. Chem. Soc. 2021, 143, 10275–10281. [Google Scholar] [CrossRef] [PubMed]
  31. Li, X.; Shi, J.; Chen, J.; Tan, Z.; Lei, H. Lead-Free Halide Double Perovskite for High-Performance Photodetectors: Progress and Perspective. Materials 2023, 16, 4490. [Google Scholar] [CrossRef]
  32. Hutter, E.M.; Gélvez-Rueda, M.C.; Bartesaghi, D.; Grozema, F.C.; Savenije, T.J. Band-Like Charge Transport in Cs2AgBiBr6 and Mixed Antimony–Bismuth Cs2AgBi1–xSbxBr6 Halide Double Perovskites. ACS Omega 2018, 3, 11655–11662. [Google Scholar] [CrossRef] [PubMed]
  33. Ji, F.; Wang, F.; Gao, F.; Boschloo, G. Challenges and Progress in Lead-Free Halide Double Perovskite Solar Cells. Sol. RRL 2023, 7, 2201112. [Google Scholar] [CrossRef]
  34. Bartel, C.J.; Sutton, C.; Goldsmith, B.R.; Ouyang, R.; Musgrave, C.B.; Ghiringhelli, L.M.; Scheffler, M. New Tolerance Factor to Predict the Stability of Perovskite Oxides and Halides. Sci. Adv. 2019, 5, eaav0693. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, Y.; Li, X.; Zeng, H. Lead-Free Halide Double Perovskites: Structure, Luminescence, and Applications. Small Struct. 2021, 2, 2000071. [Google Scholar] [CrossRef]
  36. 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]
  37. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A Materials Genome Approach to Accelerating Materials Innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef]
  38. Zhang, T.; Cai, Z.; Chen, S. Chemical Trends in the Thermodynamic Stability and Band Gaps of 980 Halide Double Perovskites: A High-Throughput First-Principles Study. ACS Appl. Mater. Interfaces 2020, 12, 20680–20690. [Google Scholar] [CrossRef] [PubMed]
  39. Liang, G.-Q.; Zhang, J. A Machine Learning Model for Screening Thermodynamic Stable Lead-Free Halide Double Perovskites. Comput. Mater. Sci. 2022, 204, 111172. [Google Scholar] [CrossRef]
  40. Tao, Q.; Xu, P.; Li, M.; Lu, W. Machine Learning for Perovskite Materials Design and Discovery. npj Comput. Mater. 2021, 7, 23. [Google Scholar] [CrossRef]
  41. Lu, S.; Zhou, Q.; Ouyang, Y.; Guo, Y.; Li, Q.; Wang, J. Accelerated Discovery of Stable Lead-Free Hybrid Organic-Inorganic Perovskites via Machine Learning. Nat. Commun. 2018, 9, 3405. [Google Scholar] [CrossRef]
  42. Muscarella, L.A.; Hutter, E.M. Halide Double-Perovskite Semiconductors beyond Photovoltaics. ACS Energy Lett. 2022, 7, 2128–2135. [Google Scholar] [CrossRef] [PubMed]
  43. Tailor, N.K.; Listorti, A.; Colella, S.; Satapathi, S. Lead-Free Halide Double Perovskites: Fundamentals, Challenges, and Photovoltaics Applications. Adv. Mater. Technol. 2023, 8, 2200442. [Google Scholar] [CrossRef]
  44. Zhang, C.; Gao, L.; Teo, S.; Guo, Z.; Xu, Z.; Zhao, S.; Ma, T. Design of a Novel and Highly Stable Lead-Free Cs2NaBiI6 Double Perovskite for Photovoltaic Application. Sustain. Energy Fuels 2018, 2, 2419–2428. [Google Scholar] [CrossRef]
  45. Grandhi, G.K.; Toikkonen, S.; Al-Anesi, B.; Pecunia, V.; Vivo, P. Perovskite-Inspired Cu2AgBiI6 for Mesoscopic Indoor Photovoltaics under Realistic Low-Light Intensity Conditions. Sustain. Energy Fuels 2023, 7, 66–73. [Google Scholar] [CrossRef]
  46. Sansom, H.C.; Longo, G.; Wright, A.D.; Buizza, L.R.V.; Mahesh, S.; Wenger, B.; Zanella, M.; Abdi-Jalebi, M.; Pitcher, M.J.; Dyer, M.S.; et al. Highly Absorbing Lead-Free Semiconductor Cu2AgBiI6 for Photovoltaic Applications from the Quaternary CuI–AgI–BiI3 Phase Space. J. Am. Chem. Soc. 2021, 143, 3983–3992. [Google Scholar] [CrossRef] [PubMed]
  47. Yang, J.; Zhang, P.; Wei, S.-H. Band Structure Engineering of Cs2AgBiBr6 Perovskite through Order–Disordered Transition: A First-Principle Study. J. Phys. Chem. Lett. 2018, 9, 31–35. [Google Scholar] [CrossRef] [PubMed]
  48. Dai, C.; Zhang, T.; Wu, Y.; Chen, S. Halide Double-Perovskite Light-Emitting Centers Embedded in Lattice-Matched and Coherent Crystalline Matrix. Adv. Funct. Mater. 2020, 30, 2000653. [Google Scholar] [CrossRef]
  49. Liu, A.; Zhu, H.; Reo, Y.; Kim, M.-G.; Chu, H.Y.; Lim, J.H.; Kim, H.-J.; Ning, W.; Bai, S.; Noh, Y.-Y. Modulation of Vacancy-Ordered Double Perovskite Cs2SnI6 for Air-Stable Thin-Film Transistors. Cell Rep. Phys. Sci. 2022, 3, 100812. [Google Scholar] [CrossRef]
  50. Kaltzoglou, A.; Antoniadou, M.; Perganti, D.; Siranidi, E.; Raptis, V.; Trohidou, K.; Psycharis, V.; Kontos, A.G.; Falaras, P. Mixed-Halide Cs2SnI3Br3 Perovskite as Low Resistance Hole-Transporting Material in Dye-Sensitized Solar Cells. Electrochim. Acta 2015, 184, 466–474. [Google Scholar] [CrossRef]
  51. Zhou, B.; Liu, Z.; Fang, S.; Zhong, H.; Tian, B.; Wang, Y.; Li, H.; Hu, H.; Shi, Y. Efficient White Photoluminescence from Self-Trapped Excitons in Sb3+/Bi3+-Codoped Cs2NaInCl6 Double Perovskites with Tunable Dual-Emission. ACS Energy Lett. 2021, 6, 3343–3351. [Google Scholar] [CrossRef]
  52. Liu, Y.; Dai, X.; Zeng, X.; Yuan, X.; Wang, Y.; Song, Y.; Chen, H.; Zhang, C.; Wang, Y.; Wan, L.; et al. High-Efficient Blue Emission and Bandgap Engineering from Jahn–Teller Distorted Halide Double Perovskites. Adv. Opt. Mater. 2023, 2301576. [Google Scholar] [CrossRef]
  53. Noculak, A.; Morad, V.; McCall, K.M.; Yakunin, S.; Shynkarenko, Y.; Wörle, M.; Kovalenko, M.V. 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]
  54. Zhao, F.; Song, Z.; Zhao, J.; Liu, Q. Double Perovskite Cs2AgInCl6:Cr3+: Broadband and near-Infrared Luminescent Materials. Inorg. Chem. Front. 2019, 6, 3621–3628. [Google Scholar] [CrossRef]
  55. 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]
  56. Li, H.; Tian, L.; Shi, Z.; Li, Y.; Li, C.; Feng, J.; Zhang, H. Double Perovskite Cs2 NaInCl6 Nanocrystals with Intense Dual-Emission via Self-Trapped Exciton-to-Tb3+ Dopant Energy Transfer. J. Mater. Chem. C 2022, 10, 10609–10615. [Google Scholar] [CrossRef]
  57. Li, S.; Hu, Q.; Luo, J.; Jin, T.; Liu, J.; Li, J.; Tan, Z.; Han, Y.; Zheng, Z.; Zhai, T.; et al. Self-Trapped Exciton to Dopant Energy Transfer in Rare Earth Doped Lead-Free Double Perovskite. Adv. Opt. Mater. 2019, 7, 1901098. [Google Scholar] [CrossRef]
  58. Hu, M.; Luo, J.; Li, S.; Liu, J.; Li, J.; Tan, Z.; Niu, G.; Wang, Z.; Tang, J. Broadband Emission of Double Perovskite Cs2Na0.4Ag0.6In0.995Bi0.005Cl6:Mn2+ for Single-Phosphor White-Light-Emitting Diodes. Opt. Lett. 2019, 44, 4757–4760. [Google Scholar] [CrossRef]
  59. Gray, M.B.; Majher, J.D.; Strom, T.A.; Woodward, P.M. Broadband White Emission in Cs2AgIn1–xBixCl6 Phosphors. Inorg. Chem. 2019, 58, 13403–13410. [Google Scholar] [CrossRef] [PubMed]
  60. Deng, W.; Fan, T.; Wu, X.; Lü, J.; Fan, W.; Deng, T.; He, H. Highly Efficient Double Perovskite (Bi,Gd)-Codoped Cs2Ag0.4Na0.6InCl6 Phosphors for Warm White LEDs. Ceram. Int. 2022, 48, 29991–29996. [Google Scholar] [CrossRef]
  61. Li, S.; Shi, Z.; Zhang, F.; Wang, L.; Ma, Z.; Wu, D.; Yang, D.; Chen, X.; Tian, Y.; Zhang, Y.; et al. Ultrastable Lead-Free Double Perovskite Warm-White Light-Emitting Devices with a Lifetime above 1000 Hours. ACS Appl. Mater. Interfaces 2020, 12, 46330–46339. [Google Scholar] [CrossRef]
  62. Zeng, Z.; Sun, M.; Zhang, S.; Zhang, H.; Shi, X.; Ye, S.; Huang, B.; Du, Y.; Yan, C.-H. Rare-Earth-Based Perovskite Cs2AgScCl6:Bi for Strong Full Visible Spectrum Emission. Adv. Funct. Mater. 2022, 32, 2204780. [Google Scholar] [CrossRef]
  63. Facile Melting-Crystallization Synthesis of Cs2NaxAg1–xInCl6: Bi Double Perovskites for White Light-Emitting Diodes | Inorganic Chemistry. Available online: https://pubs.acs.org/doi/full/10.1021/acs.inorgchem.1c03996 (accessed on 8 January 2024).
  64. Guedel, H.U.; Snellgrove, T.R. Jahn-Teller Effect in the 4T2g State of Chromium(III) in Dicesium Sodium Indium(III) Hexachloride. Inorg. Chem. 1978, 17, 1617–1620. [Google Scholar] [CrossRef]
  65. Reber, C.; Gudel, H.U. Near-infrared luminescence spectroscopy of Al2O3: V3+ and YP3O9: V3+. Chem. Phys. Lett. 1989, 154, 7. [Google Scholar] [CrossRef]
  66. Arfin, H.; Nag, A. Origin of Luminescence in Sb3+- and Bi3+-Doped Cs2SnCl6 Perovskites: Excited State Relaxation and Spin–Orbit Coupling. J. Phys. Chem. Lett. 2021, 12, 10002–10008. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, L.; Xie, X.; Li, D.; Yuan, Y.; Xue, X.; Li, Q.; Xu, J.; Wang, H.; Hu, F.; Zhang, X. Investigation on Lead-Free Mn-Doped Cs2NaInCl6 Double Perovskite Phosphors and Their Optical Properties. Opt. Mater. 2021, 122, 111802. [Google Scholar] [CrossRef]
  68. Li, X.; Shen, X.; Lu, M.; Wu, J.; Zhong, Y.; Wu, Z.; Yu, W.W.; Gao, Y.; Hu, J.; Zhu, J.; et al. Wide-Coverage and Efficient NIR Emission from Single-Component Nanophosphors through Shaping Multiple Metal-Halide Packages. Angew. Chem. Int. Ed. 2023, 62, e202217832. [Google Scholar] [CrossRef] [PubMed]
  69. Boutinaud, P. ERRATUM: Revisiting Duffy’s Model for Sb3+ and Bi3+ in Double Halide Perovskites: Emergence of a Descriptor for Machine Learning. Opt. Mater. X 2021, 12, 100109. [Google Scholar] [CrossRef]
  70. Ji, F.; Huang, Y.; Wang, F.; Kobera, L.; Xie, F.; Klarbring, J.; Abbrent, S.; Brus, J.; Yin, C.; Simak, S.I.; et al. Near-Infrared Light-Responsive Cu-Doped Cs2AgBiBr6. Adv. Funct. Mater. 2020, 30, 2005521. [Google Scholar] [CrossRef]
  71. Liu, G.; Zhang, Z.; Wu, C.; Zhang, Y.; Li, X.; Yu, W.; Yao, G.; Liu, S.; Shi, J.; Liu, K.; et al. Extending Absorption of Cs2AgBiBr6 to Near-Infrared Region (≈1350 nm) with Intermediate Band. Adv. Funct. Mater. 2022, 32, 2109891. [Google Scholar] [CrossRef]
  72. Mopoung, K.; Dávid, A.; Liu, X.; Fahlman, M.; Buyanova, I.A.; Chen, W.M.; Puttisong, Y. Spin Centers in Vanadium-Doped Cs2NaInCl6 Halide Double Perovskites. ACS Mater. Lett. 2024, 6, 566–571. [Google Scholar] [CrossRef]
  73. Dahl, J.C.; Osowiecki, W.T.; Cai, Y.; Swabeck, J.K.; Bekenstein, Y.; Asta, M.; Chan, E.M.; Alivisatos, A.P. Probing the Stability and Band Gaps of Cs2AgInCl6 and Cs2AgSbCl6 Lead-Free Double Perovskite Nanocrystals. Chem. Mater. 2019, 31, 3134–3143. [Google Scholar] [CrossRef]
  74. Liu, Z.; Yang, H.; Wang, J.; Yuan, Y.; Hills-Kimball, K.; Cai, T.; Wang, P.; Tang, A.; Chen, O. Synthesis of Lead-Free Cs2AgBiX6 (X = Cl, Br, I) Double Perovskite Nanoplatelets and Their Application in CO2 Photocatalytic Reduction. Nano Lett. 2021, 21, 1620–1627. [Google Scholar] [CrossRef] [PubMed]
  75. Zdražil, L.; Kalytchuk, S.; Langer, M.; Ahmad, R.; Pospíšil, J.; Zmeškal, O.; Altomare, M.; Osvet, A.; Zbořil, R.; Schmuki, P.; et al. Transparent and Low-Loss Luminescent Solar Concentrators Based on Self-Trapped Exciton Emission in Lead-Free Double Perovskite Nanocrystals. ACS Appl. Energy Mater. 2021, 4, 6445–6453. [Google Scholar] [CrossRef]
  76. Wu, D.; Tao, Y.; Huang, Y.; Huo, B.; Zhao, X.; Yang, J.; Jiang, X.; Huang, Q.; Dong, F.; Tang, X. High Visible-Light Photocatalytic Performance of Stable Lead-Free Cs2AgBiBr6 Double Perovskite Nanocrystals. J. Catal. 2021, 397, 27–35. [Google Scholar] [CrossRef]
  77. Cong, M.; Zhang, Q.; Yang, B.; Chen, J.; Xiao, J.; Zheng, D.; Zheng, T.; Zhang, R.; Qing, G.; Zhang, C.; et al. Bright Triplet Self-Trapped Excitons to Dopant Energy Transfer in Halide Double-Perovskite Nanocrystals. Nano Lett. 2021, 21, 8671–8678. [Google Scholar] [CrossRef]
  78. Ahmad, R.; Nutan, G.V.; Singh, D.; Gupta, G.; Soni, U.; Sapra, S.; Srivastava, R. Colloidal Lead-Free Cs2AgBiBr6 Double Perovskite Nanocrystals: Synthesis, Uniform Thin-Film Fabrication, and Application in Solution-Processed Solar Cells. Nano Res. 2021, 14, 1126–1134. [Google Scholar] [CrossRef]
  79. Zhou, W.; Li, C.; Wu, T.; Liu, R.; Ding, Z.; Zhang, R.; Yu, Y.; Han, P.; Lu, R. Bright Green-Emitting All-Inorganic Terbium Halide Double Perovskite Nanocrystals for Low-Dose X-Ray Imaging. J. Phys. Chem. Lett. 2023, 14, 8577–8583. [Google Scholar] [CrossRef]
  80. Locardi, F.; Cirignano, M.; Baranov, D.; Dang, Z.; Prato, M.; Drago, F.; Ferretti, M.; Pinchetti, V.; Fanciulli, M.; Brovelli, S.; et al. Colloidal Synthesis of Double Perovskite Cs2AgInCl6 and Mn-Doped Cs2AgInCl6 Nanocrystals. J. Am. Chem. Soc. 2018, 140, 12989–12995. [Google Scholar] [CrossRef]
  81. 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]
  82. 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]
  83. Ahmad, R.; Zdražil, L.; Kalytchuk, S.; Naldoni, A.; Mohammadi, E.; Schmuki, P.; Zboril, R.; Kment, Š. Robust Dual Cationic Ligand for Stable and Efficient Warm-White Light Emission in Lead-Free Double Perovskite Nanocrystals. Appl. Mater. Today 2022, 26, 101288. [Google Scholar] [CrossRef]
  84. Li, L.; Shao, H.; Wu, X.; Chen, W.; Zhu, J.; Dong, B.; Xu, L.; Xu, W.; Hu, J.; Zhou, M.; et al. Aluminum-Doped Lead-Free Double Perovskite Cs2AgBiCl6 Nanocrystals with Ultrahigh Stability towards White Light Emitting Diodes. Mater. Res. Bull. 2022, 147, 111645. [Google Scholar] [CrossRef]
  85. Han, P.; Mao, X.; Yang, S.; Zhang, F.; Yang, B.; Wei, D.; Deng, W.; Han, K. Lead-Free Sodium–Indium Double Perovskite Nanocrystals through Doping Silver Cations for Bright Yellow Emission. Angew. Chem. 2019, 131, 17391–17395. [Google Scholar] [CrossRef]
  86. Liu, X.; Xu, X.; Li, B.; Yang, L.; Li, Q.; Jiang, H.; Xu, D. Tunable Dual-Emission in Monodispersed Sb3+/Mn2+ Codoped Cs2NaInCl6 Perovskite Nanocrystals through an Energy Transfer Process. Small 2020, 16, 2002547. [Google Scholar] [CrossRef] [PubMed]
  87. 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]
  88. Wang, Z.; Zhang, R.; Mao, X.; Zheng, D.; Liu, S.; Liu, F.; Han, K.; Yang, B. Boosting the Self-Trapped Exciton Emission in Cs2NaYCl6 Double Perovskite Single Crystals and Nanocrystals. J. Phys. Chem. Lett. 2022, 13, 8613–8619. [Google Scholar] [CrossRef] [PubMed]
  89. Liu, S.; Yang, B.; Chen, J.; Wei, D.; Zheng, D.; Kong, Q.; Deng, W.; Han, K. Efficient Thermally Activated Delayed Fluorescence from All-Inorganic Cesium Zirconium Halide Perovskite Nanocrystals. Angew. Chem. 2020, 132, 22109–22113. [Google Scholar] [CrossRef]
  90. Liu, S.; Yang, B.; Chen, J.; Zheng, D.; Tang, Z.; Deng, W.; Han, K. Colloidal Synthesis and Tunable Multicolor Emission of Vacancy-Ordered Cs2HfCl6 Perovskite Nanocrystals. Laser Photonics Rev. 2022, 16, 2100439. [Google Scholar] [CrossRef]
  91. Luo, C.; Han, P.; Hou, J.; Li, J.; Sun, F.; Li, C.; Liu, J.; Yang, B. Elucidating the Role of Antimony Dopant in Optical Properties of Brightly Luminescent Zero-Dimensional Organic–Inorganic Metal Halides. J. Phys. Chem. C 2023, 127, 10720–10729. [Google Scholar] [CrossRef]
  92. 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]
  93. Jin, S.; Yuan, H.; Pang, T.; Zhang, M.; Li, J.; Zheng, Y.; Wu, T.; Zhang, R.; Wang, Z.; Chen, D. Highly Bright and Stable Lead-Free Double Perovskite White Light-Emitting Diodes. Adv. Mater. 2023, 2308487. [Google Scholar] [CrossRef] [PubMed]
  94. Bai, T.; Yang, B.; Chen, J.; Zheng, D.; Tang, Z.; Wang, X.; Zhao, Y.; Lu, R.; Han, K. Efficient Luminescent Halide Quadruple-Perovskite Nanocrystals via Trap-Engineering for Highly Sensitive Photodetectors. Adv. Mater. 2021, 33, 2007215. [Google Scholar] [CrossRef] [PubMed]
  95. Li, B.; Wu, X.; Zhang, S.; Li, Z.; Gao, D.; Chen, X.; Xiao, S.; Chueh, C.-C.; Jen, A.K.-Y.; Zhu, Z. Efficient and Stable Cs2AgBiBr6 Double Perovskite Solar Cells through In-Situ Surface Modulation. Chem. Eng. J. 2022, 446, 137144. [Google Scholar] [CrossRef]
  96. Marongiu, D.; Lai, S.; Liu, F.; Simbula, A.; Quochi, F.; Saba, M.; Mura, A.; Bongiovanni, G. Halide Double-Perovskites: High Efficient Light Emission and Beyond. APL Energy 2023, 1, 021501. [Google Scholar] [CrossRef]
  97. Hu, Q.; Deng, Z.; Hu, M.; Zhao, A.; Zhang, Y.; Tan, Z.; Niu, G.; Wu, H.; Tang, J. X-Ray Scintillation in Lead-Free Double Perovskite Crystals. Sci. China Chem. 2018, 61, 1581–1586. [Google Scholar] [CrossRef]
  98. Zhou, Y.; Chen, J.; Bakr, O.M.; Mohammed, O.F. Metal Halide Perovskites for X-Ray Imaging Scintillators and Detectors. ACS Energy Lett. 2021, 6, 739–768. [Google Scholar] [CrossRef]
  99. He, X.; Deng, Y.; Ouyang, D.; Zhang, N.; Wang, J.; Murthy, A.A.; Spanopoulos, I.; Islam, S.M.; Tu, Q.; Xing, G.; et al. Recent Development of Halide Perovskite Materials and Devices for Ionizing Radiation Detection. Chem. Rev. 2023, 123, 1207–1261. [Google Scholar] [CrossRef]
  100. Cheng, X.; Qian, W.; Wang, J.; Yu, C.; He, J.; Li, H.; Xu, Q.; Chen, D.; Li, N.; Lu, J. Environmentally Robust Memristor Enabled by Lead-Free Double Perovskite for High-Performance Information Storage. Small 2019, 15, 1905731. [Google Scholar] [CrossRef]
  101. Ji, F.; Klarbring, J.; Zhang, B.; Wang, F.; Wang, L.; Miao, X.; Ning, W.; Zhang, M.; Cai, X.; Bakhit, B.; et al. Remarkable Thermochromism in the Double Perovskite Cs2NaFeCl6. Adv. Opt. Mater. 2023, 2301102. [Google Scholar] [CrossRef]
  102. Ning, W.; Zhao, X.; Klarbring, J.; Bai, S.; Ji, F.; Wang, F.; Simak, S.I.; Tao, Y.; Ren, X.; Zhang, L.; et al. Thermochromic Lead-Free Halide Double Perovskites. Adv. Funct. Mater. 2019, 29, 1807375. [Google Scholar] [CrossRef]
  103. García-Espejo, G.; Rodríguez-Padrón, D.; Luque, R.; Camacho, L.; de Miguel, G. Mechanochemical Synthesis of Three Double Perovskites: Cs2AgBiBr6, (CH3NH3)2TlBiBr6 and Cs2AgSbBr6. Nanoscale 2019, 11, 16650–16657. [Google Scholar] [CrossRef]
  104. Giovilli, G.; Albini, B.; Grisci, V.; Bonomi, S.; Moroni, M.; Mosconi, E.; Kaiser, W.; De Angelis, F.; Galinetto, P.; Malavasi, L. Band Gap Tuning through Cation and Halide Alloying in Mechanochemically Synthesized Cs3(Sb1−xBix)2Br9 and Cs3Sb2(I1−xBrx)9 Solid Solutions. J. Mater. Chem. C 2023, 11, 10282–10291. [Google Scholar] [CrossRef]
  105. Knochenmuss, R.; Reber, C.; Rajasekharan, M.V.; Güdel, H.U. Broadband Near-infrared Luminescence of Cr+3 in the Elpasolite Lattices Cs2NaInCl6, Cs2NaYCl6, and Cs2NaYBr6. J. Chem. Phys. 1986, 85, 4280–4289. [Google Scholar] [CrossRef]
  106. Cao, X.; Kang, L.; Guo, S.; Zhang, M.; Lin, Z.; Gao, J. Cs2NaVCl6: A Pb-Free Halide Double Perovskite with Strong Visible and Near-Infrared Light Absorption. ACS Appl. Mater. Interfaces 2019, 11, 38648–38653. [Google Scholar] [CrossRef] [PubMed]
  107. Yan, J.; Li, H.; Aldamasy, M.H.; Frasca, C.; Abate, A.; Zhao, K.; Hu, Y. Advances in the Synthesis of Halide Perovskite Single Crystals for Optoelectronic Applications. Chem. Mater. 2023, 35, 2683–2712. [Google Scholar] [CrossRef]
  108. Deng, Y.-H.; Yang, Z.-Q.; Ma, R.-M. Growth of Centimeter-Scale Perovskite Single-Crystalline Thin Film via Surface Engineering. Nano Converg. 2020, 7, 25. [Google Scholar] [CrossRef] [PubMed]
  109. Vidyasagar, C.C.; Muñoz Flores, B.M.; Jiménez Pérez, V.M. Recent Advances in Synthesis and Properties of Hybrid Halide Perovskites for Photovoltaics. Nano-Micro Lett. 2018, 10, 68. [Google Scholar] [CrossRef] [PubMed]
  110. Jiang, Y.; Li, K.; Wu, X.; Zhu, M.; Zhang, H.; Zhang, K.; Wang, Y.; Loh, K.P.; Shi, Y.; Xu, Q.-H. In Situ Synthesis of Lead-Free Halide Perovskite Cs2AgBiBr6 Supported on Nitrogen-Doped Carbon for Efficient Hydrogen Evolution in Aqueous HBr Solution. ACS Appl. Mater. Interfaces 2021, 13, 10037–10046. [Google Scholar] [CrossRef] [PubMed]
  111. He, Y.; Matei, L.; Jung, H.J.; McCall, K.M.; Chen, M.; Stoumpos, C.C.; Liu, Z.; Peters, J.A.; Chung, D.Y.; Wessels, B.W.; et al. High Spectral Resolution of Gamma-Rays at Room Temperature by Perovskite CsPbBr3 Single Crystals. Nat. Commun. 2018, 9, 1609. [Google Scholar] [CrossRef] [PubMed]
  112. Gao, W.; Ran, C.; Xi, J.; Jiao, B.; Zhang, W.; Wu, M.; Hou, X.; Wu, Z. High-Quality Cs2AgBiBr6 Double Perovskite Film for Lead-Free Inverted Planar Heterojunction Solar Cells with 2.2% Efficiency. ChemPhysChem 2018, 19, 1696–1700. [Google Scholar] [CrossRef]
  113. Usman, M.; Yan, Q. Recent Advancements in Crystalline Pb-Free Halide Double Perovskites. Crystals 2020, 10, 62. [Google Scholar] [CrossRef]
  114. Zhou, L.; Xu, Y.; Chen, B.; Kuang, D.; Su, C. Synthesis and Photocatalytic Application of Stable Lead-Free Cs2 AgBiBr6 Perovskite Nanocrystals. Small 2018, 14, 1703762. [Google Scholar] [CrossRef] [PubMed]
  115. 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] [PubMed]
  116. 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]
  117. Lv, K.; Qi, S.; Liu, G.; Lou, Y.; Chen, J.; Zhao, Y. Lead-Free Silver-Antimony Halide Double Perovskite Quantum Dots with Superior Blue Photoluminescence. Chem. Commun. 2019, 55, 14741–14744. [Google Scholar] [CrossRef]
  118. Mahor, Y.; Mir, W.J.; Nag, A. Synthesis and Near-Infrared Emission of Yb-Doped Cs2AgInCl6 Double Perovskite Microcrystals and Nanocrystals. J. Phys. Chem. C 2019, 123, 15787–15793. [Google Scholar] [CrossRef]
  119. Lee, W.; Choi, D.; Kim, S. Colloidal Synthesis of Shape-Controlled Cs2NaBiX6 (X = Cl, Br) Double Perovskite Nanocrystals: Discrete Optical Transition by Non-Bonding Characters and Energy Transfer to Mn Dopants. Chem. Mater. 2020, 32, 6864–6874. [Google Scholar] [CrossRef]
  120. Tang, H.; Xu, Y.; Hu, X.; Hu, Q.; Chen, T.; Jiang, W.; Wang, L.; Jiang, W. Facile One-Pot Preparation and Luminescence Mechanism of Stable Bi3+-Doped Cs2AgInCl6 Double Perovskite Nanocrystals. J. Lumin. 2022, 252, 119246. [Google Scholar] [CrossRef]
  121. Yang, J.-N.; Song, Y.; Yao, J.-S.; Wang, K.-H.; Wang, J.-J.; Zhu, B.-S.; Yao, M.-M.; Rahman, S.U.; Lan, Y.-F.; Fan, F.-J.; et al. Potassium Bromide Surface Passivation on CsPbI3-xBrx Nanocrystals for Efficient and Stable Pure Red Perovskite Light-Emitting Diodes. J. Am. Chem. Soc. 2020, 142, 2956–2967. [Google Scholar] [CrossRef]
  122. Ugur, E.; Aydin, E.; De Bastiani, M.; Harrison, G.T.; Yildirim, B.K.; Teale, S.; Chen, B.; Liu, J.; Wang, M.; Seitkhan, A.; et al. Front-Contact Passivation through 2D/3D Perovskite Heterojunctions Enables Efficient Bifacial Perovskite/Silicon Tandem Solar Cells. Matter 2023, 6, 2919–2934. [Google Scholar] [CrossRef]
  123. Garai, R.; Gupta, R.K.; Iyer, P.K. Trap State Passivation for Stabilizing Perovskite Solar Cells via Multifunctional Molecules. Acc. Mater. Res. 2023, 4, 560–565. [Google Scholar] [CrossRef]
  124. Aktas, E.; Rajamanickam, N.; Pascual, J.; Hu, S.; Aldamasy, M.H.; Di Girolamo, D.; Li, W.; Nasti, G.; Martínez-Ferrero, E.; Wakamiya, A.; et al. Challenges and Strategies toward Long-Term Stability of Lead-Free Tin-Based Perovskite Solar Cells. Commun. Mater. 2022, 3, 104. [Google Scholar] [CrossRef]
  125. Hsu, S.-N.; Zhao, W.; Gao, Y.; Akriti; Segovia, M.; Xu, X.; Boudouris, B.W.; Dou, L. Thermoelectric Performance of Lead-Free Two-Dimensional Halide Perovskites Featuring Conjugated Ligands. Nano Lett. 2021, 21, 7839–7844. [Google Scholar] [CrossRef] [PubMed]
  126. Akriti; Shi, E.; Dou, L. A Leap towards High-Performance 2D Perovskite Photodetectors. Trends Chem. 2019, 1, 365–367. [Google Scholar] [CrossRef]
  127. Xue, J.; Wang, Z.; Comstock, A.; Wang, Z.; Sung, H.H.Y.; Williams, I.D.; Sun, D.; Liu, J.; Lu, H. Chemical Control of Magnetic Ordering in Hybrid Fe–Cl Layered Double Perovskites. Chem. Mater. 2022, 34, 2813–2823. [Google Scholar] [CrossRef]
  128. Chao, I.-H.; Yang, Y.-T.; Yu, M.-H.; Chen, C.-H.; Liao, C.-H.; Lin, B.-H.; Ni, I.-C.; Chen, W.-C.; Ho-Baillie, A.W.Y.; Chueh, C.-C. Performance Enhancement of Lead-Free 2D Tin Halide Perovskite Transistors by Surface Passivation and Its Impact on Non-Volatile Photomemory Characteristics. Small 2023, 19, 2207734. [Google Scholar] [CrossRef] [PubMed]
  129. Zhang, F.; Shao, M.; Wang, C.; Wen, W.; Shi, W.; Qin, M.; Huang, H.; Wei, X.; Guo, Y.; Liu, Y. Photoinduced Nonvolatile Memory Transistor Based on Lead-Free Perovskite Incorporating Fused Π-Conjugated Organic Ligands. Adv. Mater. 2024, 36, 2307326. [Google Scholar] [CrossRef] [PubMed]
  130. Sun, B.; Liu, X.-F.; Li, X.-Y.; Zhang, Y.; Shao, X.; Yang, D.; Zhang, H.-L. Two-Dimensional Perovskite Chiral Ferromagnets. Chem. Mater. 2020, 32, 8914–8920. [Google Scholar] [CrossRef]
  131. Milić, J.V. Multifunctional Layered Hybrid Perovskites. J. Mater. Chem. C 2021, 9, 11428–11443. [Google Scholar] [CrossRef]
  132. Connor, B.A.; Leppert, L.; Smith, M.D.; Neaton, J.B.; Karunadasa, H.I. Layered Halide Double Perovskites: Dimensional Reduction of Cs2 AgBiBr6. J. Am. Chem. Soc. 2018, 140, 5235–5240. [Google Scholar] [CrossRef]
  133. Mao, L.; Stoumpos, C.C.; Kanatzidis, M.G. Two-Dimensional Hybrid Halide Perovskites: Principles and Promises. J. Am. Chem. Soc. 2019, 141, 1171–1190. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, A.; Zuo, C.; Niu, X.; Ding, L.; Ding, J.; Hao, F. Recent Promise of Lead-Free Halide Perovskites in Optoelectronic Applications. Chem. Eng. J. 2023, 451, 138926. [Google Scholar] [CrossRef]
  135. McClure, E.T.; McCormick, A.P.; Woodward, P.M. Four Lead-Free Layered Double Perovskites with the n = 1 Ruddlesden–Popper Structure. Inorg. Chem. 2020, 59, 6010–6017. [Google Scholar] [CrossRef] [PubMed]
  136. Vishnoi, P.; Zuo, J.L.; Li, X.; Binwal, D.C.; Wyckoff, K.E.; Mao, L.; Kautzsch, L.; Wu, G.; Wilson, S.D.; Kanatzidis, M.G.; et al. Hybrid Layered Double Perovskite Halides of Transition Metals. J. Am. Chem. Soc. 2022, 144, 6661–6666. [Google Scholar] [CrossRef] [PubMed]
  137. Zhao, J.; Zhang, Z.; Li, G.; Aldamasy, M.H.; Li, M.; Abate, A. Dimensional Tuning in Lead-Free Tin Halide Perovskite for Solar Cells. Adv. Energy Mater. 2023, 13, 2204233. [Google Scholar] [CrossRef]
  138. Sun, P.-P.; Zhang, X.; Yuan, S.; Chi, W.; Cai, G. High-Performance Photoabsorber Based on 2D Perovskites with Alternating Cations in the Interlayer Space: GA(MA)3M3I10. Sol. RRL 2023, 7, 2300441. [Google Scholar] [CrossRef]
  139. Li, Z.; Kavanagh, S.R.; Napari, M.; Palgrave, R.G.; Abdi-Jalebi, M.; Andaji-Garmaroudi, Z.; Davies, D.W.; Laitinen, M.; Julin, J.; Isaacs, M.A.; et al. Bandgap Lowering in Mixed Alloys of Cs2Ag(SbxBi1−x)Br6 Double Perovskite Thin Films. J. Mater. Chem. A 2020, 8, 21780–21788. [Google Scholar] [CrossRef]
  140. Jana, M.K.; Janke, S.M.; Dirkes, D.J.; Dovletgeldi, S.; Liu, C.; Qin, X.; Gundogdu, K.; You, W.; Blum, V.; Mitzi, D.B. Direct-Bandgap 2D Silver–Bismuth Iodide Double Perovskite: The Structure-Directing Influence of an Oligothiophene Spacer Cation. J. Am. Chem. Soc. 2019, 141, 7955–7964. [Google Scholar] [CrossRef]
  141. Coccia, C.; Moroni, M.; Malavasi, L. Chiral Metal Halide Perovskites: Focus on Lead-Free Materials and Structure-Property Correlations. Molecules 2023, 28, 6166. [Google Scholar] [CrossRef]
  142. Li, Q.; Cao, D.; Liu, X.; Zhou, X.; Chen, X.; Shu, H. Hierarchical Computational Screening of Layered Lead-Free Metal Halide Perovskites for Optoelectronic Applications. J. Mater. Chem. A 2021, 9, 6476–6486. [Google Scholar] [CrossRef]
  143. Chen, M.; Shan, Z.; Dong, X.; Liu (Frank), S.; Xu, Z. Discovering Layered Lead-Free Perovskite Solar Absorbers via Cation Transmutation. Nanoscale Horiz. 2023, 8, 483–488. [Google Scholar] [CrossRef]
  144. Panda, D.P.; Swain, D.; Rohj, R.K.; Sarma, D.D.; Sundaresan, A. Elucidating Structure–Property Correlation in Perovskitoid and Antiperovskite Piperidinium Manganese Chloride. Inorg. Chem. 2023, 62, 3202–3211. [Google Scholar] [CrossRef]
  145. Zhao, Y.; Daemen, L.L. Superionic Conductivity in Lithium-Rich Anti-Perovskites. J. Am. Chem. Soc. 2012, 134, 15042–15047. [Google Scholar] [CrossRef] [PubMed]
  146. Yan, S.; Tang, K.; Lin, Y.; Ren, Y.; Tian, W.; Chen, H.; Lin, T.; Qiu, L.; Pan, X.; Wang, W. Light-Emitting Diodes with Manganese Halide Tetrahedron Embedded in Anti-Perovskites. ACS Energy Lett. 2021, 6, 1901–1911. [Google Scholar] [CrossRef]
  147. Ju, M.-G.; Dai, J.; Ma, L.; Zhou, Y.; Zeng, X.C. Zero-Dimensional Organic–Inorganic Perovskite Variant: Transition between Molecular and Solid Crystal. J. Am. Chem. Soc. 2018, 140, 10456–10463. [Google Scholar] [CrossRef] [PubMed]
  148. Karim, M.M.S.; Ganose, A.M.; Pieters, L.; Winnie Leung, W.W.; Wade, J.; Zhang, L.; Scanlon, D.O.; Palgrave, R.G. Anion Distribution, Structural Distortion, and Symmetry-Driven Optical Band Gap Bowing in Mixed Halide Cs2SnX6 Vacancy Ordered Double Perovskites. Chem. Mater. 2019, 31, 9430–9444. [Google Scholar] [CrossRef] [PubMed]
  149. Yan, J.; Zhang, S.; Wei, Q.; Cao, S.; Zhao, J.; Zou, B.; Zeng, R. Stoichiometry-Controlled Phase Engineering of Cesium Bismuth Halides and Reversible Structure Switch. Adv. Opt. Mater. 2022, 10, 2101406. [Google Scholar] [CrossRef]
  150. Dave, K.; Huang, W.-T.; Liu, R.-S. All Inorganic Lead-Free Zero-Dimensional Metal Halide Luminescent Materials and Applications. Crystals 2023, 13, 499. [Google Scholar] [CrossRef]
  151. Peng, H.; Tian, Y.; Yu, Z.; Wang, X.; Ke, B.; Zhao, Y.; Dong, T.; Wang, J.; Zou, B. (C16H28N)2SbCl5: A New Lead-Free Zero-Dimensional Metal-Halide Hybrid with Bright Orange Emission. Sci. China Mater. 2022, 65, 1594–1600. [Google Scholar] [CrossRef]
  152. Zhou, C.; Lin, H.; Shi, H.; Tian, Y.; Pak, C.; Shatruk, M.; Zhou, Y.; Djurovich, P.; Du, M.-H.; Ma, B. A Zero-Dimensional Organic Seesaw-Shaped Tin Bromide with Highly Efficient Strongly Stokes-Shifted Deep-Red Emission. Angew. Chem. Int. Ed. 2018, 57, 1021–1024. [Google Scholar] [CrossRef]
  153. Zhou, C.; Xu, L.-J.; Lee, S.; Lin, H.; Ma, B. Recent Advances in Luminescent Zero-Dimensional Organic Metal Halide Hybrids. Adv. Opt. Mater. 2021, 9, 2001766. [Google Scholar] [CrossRef]
  154. Xiao, H.; Dang, P.; Yun, X.; Li, G.; Wei, Y.; Wei, Y.; Xiao, X.; Zhao, Y.; Molokeev, M.S.; Cheng, Z.; et al. Solvatochromic Photoluminescent Effects in All-Inorganic Manganese(II)-Based Perovskites by Highly Selective Solvent-Induced Crystal-to-Crystal Phase Transformations. Angew. Chem. Int. Ed. 2021, 60, 3699–3707. [Google Scholar] [CrossRef]
  155. Tang, H.; Yuan, K.J.; Zheng, P.F.; Xiao, T.Q.; Zhang, H.W.; Zhao, X.C.; Zhou, W.; Wang, S.Y.; Liu, W.F. Synthesis, Crystal Structure and Optical Properties of the Quasi-0D Lead-Free Organic-Inorganic Hybrid Crystal (C6H14N)3Bi2I9·H2O. J. Solid State Chem. 2023, 323, 124011. [Google Scholar] [CrossRef]
  156. Chu, Z.; Chu, X.; Zhao, Y.; Ye, Q.; Jiang, J.; Zhang, X.; You, J. Emerging Low-Dimensional Crystal Structure of Metal Halide Perovskite Optoelectronic Materials and Devices. Small Struct. 2021, 2, 2000133. [Google Scholar] [CrossRef]
  157. Liu, Y.-H.; Wang, W.-Q.; Zhang, B.-L.; Wang, Y.-J.; Ren, M.-P.; Jing, Z.; Yue, C.-Y. Zero-Dimensional Organic–Inorganic Hybrid Zinc Halide with Broadband Yellow Light Emission. CrystEngComm 2023, 25, 444–449. [Google Scholar] [CrossRef]
  158. Wang, Z.; Zheng, S.; Teng, Q.; Li, C.; Zhuang, B.; Zhang, R.; Huang, F.; Chen, D.; Yuan, F. Opportunity of Lead-Free Metal Halide Perovskites for Electroluminescence. Innov. Mater. 2023, 1, 100015–100088. [Google Scholar] [CrossRef]
  159. Sun, S.; Lu, M.; Gao, X.; Shi, Z.; Bai, X.; Yu, W.W.; Zhang, Y. 0D Perovskites: Unique Properties, Synthesis, and Their Applications. Adv. Sci. 2021, 8, 2102689. [Google Scholar] [CrossRef] [PubMed]
  160. 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]
  161. Zhao, J.-Q.; Shi, H.-S.; Zeng, L.-R.; Ge, H.; Hou, Y.-H.; Wu, X.-M.; Yue, C.-Y.; Lei, X.-W. Highly Emissive Zero-Dimensional Antimony Halide for Anti-Counterfeiting and Confidential Information Encryption-Decryption. Chem. Eng. J. 2022, 431, 134336. [Google Scholar] [CrossRef]
  162. Williams, R.T.; Song, K.S. The Self-Trapped Exciton. J. Phys. Chem. Solids 1990, 51, 679–716. [Google Scholar] [CrossRef]
  163. Zhou, C.; Lin, H.; Tian, Y.; Yuan, Z.; Clark, R.; Chen, B.; van de Burgt, L.J.; Wang, J.C.; Zhou, Y.; Hanson, K.; et al. Luminescent Zero-Dimensional Organic Metal Halide Hybrids with near-Unity Quantum Efficiency. Chem. Sci. 2018, 9, 586–593. [Google Scholar] [CrossRef] [PubMed]
  164. Patra, S.G.; Datta, D. Ligand Field Splitting in Homoleptic Tetrahedral D10 Transition Metal Complexes. Spectrochemical Series. Comput. Theor. Chem. 2018, 1130, 77–82. [Google Scholar] [CrossRef]
  165. Nguyen, H.A.; Dixon, G.; Dou, F.Y.; Gallagher, S.; Gibbs, S.; Ladd, D.M.; Marino, E.; Ondry, J.C.; Shanahan, J.P.; Vasileiadou, E.S.; et al. Design Rules for Obtaining Narrow Luminescence from Semiconductors Made in Solution. Chem. Rev. 2023, 123, 7890–7952. [Google Scholar] [CrossRef] [PubMed]
  166. Kitzmann, W.R.; Moll, J.; Heinze, K. Spin-Flip Luminescence. Photochem. Photobiol. Sci. 2022, 21, 1309–1331. [Google Scholar] [CrossRef] [PubMed]
  167. Pan, H.-M.; Yang, Q.-L.; Xing, X.-X.; Li, J.-P.; Meng, F.-L.; Zhang, X.; Xiao, P.-C.; Yue, C.-Y.; Lei, X.-W. Enhancement of the Photoluminescence Efficiency of Hybrid Manganese Halides through Rational Structural Design. Chem. Commun. 2021, 57, 6907–6910. [Google Scholar] [CrossRef] [PubMed]
  168. Guo, Q.; Zhao, X.; Song, B.; Luo, J.; Tang, J. Light Emission of Self-Trapped Excitons in Inorganic Metal Halides for Optoelectronic Applications. Adv. Mater. 2022, 34, 2201008. [Google Scholar] [CrossRef] [PubMed]
  169. Guo, Q.; Wang, L.; Yang, L.; Duan, J.; Du, H.; Ji, G.; Liu, N.; Zhao, X.; Chen, C.; Xu, L.; et al. Spectra Stable Deep-Blue Light-Emitting Diodes Based on Cryolite-like Cerium(III) Halides with Nanosecond d-f Emission. Sci. Adv. 2022, 8, eabq2148. [Google Scholar] [CrossRef]
  170. Lian, L.; Zhang, T.; Ding, H.; Zhang, P.; Zhang, X.; Zhao, Y.-B.; Gao, J.; Zhang, D.; Zhao, Y.S.; Zhang, J. Highly Luminescent Zero-Dimensional Organic Copper Halide with Low-Loss Optical Waveguides and Highly Polarized Emission. ACS Mater. Lett. 2022, 4, 1446–1452. [Google Scholar] [CrossRef]
  171. Luo, H.; Bu, K.; Yin, Y.; Wang, D.; Shi, C.; Guo, S.; Fu, T.; Liang, J.; Liu, B.; Zhang, D.; et al. Anomalous Charge Transfer from Organic Ligands to Metal Halides in Zero-Dimensional [(C6H5)4P]2SbCl5 Enabled by Pressure-Induced Lone Pair-π Interaction. Angew. Chem. Int. Ed. 2023, 62, e202304494. [Google Scholar] [CrossRef]
  172. Ran, Q.; Zhang, Y.; Yang, J.; He, R.; Zhou, L.; Hu, S. White-Light Defect Emission and Enhanced Photoluminescence Efficiency in a 0D Indium-Based Metal Halide. J. Mater. Chem. C 2022, 10, 1999–2007. [Google Scholar] [CrossRef]
  173. Zhang, Y.; Tirani, F.F.; Pattison, P.; Schenk-Joß, K.; Xiao, Z.; Nazeeruddin, M.K.; Gao, P. Zero-Dimensional Hybrid Iodobismuthate Derivatives: From Structure Study to Photovoltaic Application. Dalton Trans. 2020, 49, 5815–5822. [Google Scholar] [CrossRef] [PubMed]
  174. Yan, S.; Tian, W.; Chen, H.; Tang, K.; Lin, T.; Zhong, G.; Qiu, L.; Pan, X.; Wang, W. Synthesis of 0D Manganese-Based Organic–Inorganic Hybrid Perovskite and Its Application in Lead-Free Red Light-Emitting Diode. Adv. Funct. Mater. 2021, 31, 2100855. [Google Scholar] [CrossRef]
  175. Xu, L.-J.; Sun, C.-Z.; Xiao, H.; Wu, Y.; Chen, Z.-N. Green-Light-Emitting Diodes Based on Tetrabromide Manganese(II) Complex through Solution Process. Adv. Mater. 2017, 29, 1605739. [Google Scholar] [CrossRef] [PubMed]
  176. Qin, Y.; Tao, P.; Gao, L.; She, P.; Liu, S.; Li, X.; Li, F.; Wang, H.; Zhao, Q.; Miao, Y.; et al. Designing Highly Efficient Phosphorescent Neutral Tetrahedral Manganese(II) Complexes for Organic Light-Emitting Diodes. Adv. Opt. Mater. 2019, 7, 1801160. [Google Scholar] [CrossRef]
  177. Li, J.-L.; Sang, Y.-F.; Xu, L.-J.; Lu, H.-Y.; Wang, J.-Y.; Chen, Z.-N. Highly Efficient Light-Emitting Diodes Based on an Organic Antimony(III) Halide Hybrid. Angew. Chem. 2022, 134, e202113450. [Google Scholar] [CrossRef]
  178. Liu, H.; Shonde, T.B.; Gonzalez, F.; Olasupo, O.J.; Lee, S.; Luong, D.; Lin, X.; Vellore Winfred, J.S.R.; Lochner, E.; Fatima, I.; et al. Efficient Red Light Emitting Diodes Based on a Zero-Dimensional Organic Antimony Halide Hybrid. Adv. Mater. 2023, 35, 2209417. [Google Scholar] [CrossRef] [PubMed]
  179. Chen, H.; Zhu, L.; Xue, C.; Liu, P.; Du, X.; Wen, K.; Zhang, H.; Xu, L.; Xiang, C.; Lin, C.; et al. Efficient and Bright Warm-White Electroluminescence from Lead-Free Metal Halides. Nat. Commun. 2021, 12, 1421. [Google Scholar] [CrossRef]
  180. Liu, X.; Yuan, F.; Zhu, C.; Li, J.; Lv, X.; Xing, G.; Wei, Q.; Wang, G.; Dai, J.; Dong, H.; et al. Near-Unity Blue Luminance from Lead-Free Copper Halides for Light-Emitting Diodes. Nano Energy 2022, 91, 106664. [Google Scholar] [CrossRef]
  181. 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]
  182. Liu, S.; Fang, X.; Lu, B.; Yan, D. Wide Range Zero-Thermal-Quenching Ultralong Phosphorescence from Zero-Dimensional Metal Halide Hybrids. Nat. Commun. 2020, 11, 4649. [Google Scholar] [CrossRef]
  183. Cai, W.; Kuang, C.; Liu, T.; Shang, Y.; Zhang, J.; Qin, J.; Gao, F. Multicolor Light Emission in Manganese-Based Metal Halide Composites. Appl. Phys. Rev. 2022, 9, 041409. [Google Scholar] [CrossRef]
  184. Lin, Y.-P.; Hu, S.; Xu, J.; Zhang, Z.; Qi, X.; Lu, X.; Jin, J.; Huang, X.-Y.; Xu, Q.; Deng, Z.; et al. Bright Green Emitter of Mn-Doped C4H12N2ZnX4 (X = Cl, Br) for X-Ray Radiography and WLEDs. Chem. Eng. J. 2023, 468, 143818. [Google Scholar] [CrossRef]
  185. Li, M.; Zhou, J.; Molokeev, M.S.; Jiang, X.; Lin, Z.; Zhao, J.; Xia, Z. Lead-Free Hybrid Metal Halides with a Green-Emissive [MnBr4] Unit as a Selective Turn-On Fluorescent Sensor for Acetone. Inorg. Chem. 2019, 58, 13464–13470. [Google Scholar] [CrossRef]
  186. Shen, W.; Sui, S.; Yuan, W.; Wang, A.; Tao, Y.; Chen, S.; Deng, Z. Precise, Sensitive, and Reversible Thermochromic Luminescent Sensing Facilitated via Bright High-Temperature Luminescent PEAMnBrxI3−x (x = 0/1/2/3). J. Mater. Chem. C 2021, 9, 2729–2737. [Google Scholar] [CrossRef]
  187. 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] [PubMed]
  188. 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]
  189. Peng, H.; Tian, Y.; Wang, X.; Huang, T.; Yu, Z.; Zhao, Y.; Dong, T.; Wang, J.; Zou, B. Pure White Emission with 91.9% Photoluminescence Quantum Yield of [(C3H7)4N]2Cu2I4 out of Polaronic States and Ultra-High Color Rendering Index. ACS Appl. Mater. Interfaces 2022, 14, 12395–12403. [Google Scholar] [CrossRef] [PubMed]
  190. Peng, H.; Wang, X.; Tian, Y.; Zou, B.; Yang, F.; Huang, T.; Peng, C.; Yao, S.; Yu, Z.; Yao, Q.; et al. Highly Efficient Cool-White Photoluminescence of (Gua)3Cu2I5 Single Crystals: Formation and Optical Properties. ACS Appl. Mater. Interfaces 2021, 13, 13443–13451. [Google Scholar] [CrossRef] [PubMed]
  191. 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. Int. Ed. 2022, 61, e202205463. [Google Scholar] [CrossRef] [PubMed]
  192. Morad, V.; Shynkarenko, Y.; Yakunin, S.; Brumberg, A.; Schaller, R.D.; Kovalenko, M.V. Disphenoidal Zero-Dimensional Lead, Tin, and Germanium Halides: Highly Emissive Singlet and Triplet Self-Trapped Excitons and X-Ray Scintillation. J. Am. Chem. Soc. 2019, 141, 9764–9768. [Google Scholar] [CrossRef] [PubMed]
  193. 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]
  194. Wang, Y.; Zhou, T.; Chen, J.; Qin, H.; Wu, J.; Zhang, Q.; Zheng, J.; Li, X.; Sun, Y.-Y.; He, Y.; et al. Zero-Dimensional Organic–Inorganic Hybrid Zinc Halides for Multiple Applications in Anti-Counterfeiting, X-Ray Imaging and White LEDs. Adv. Opt. Mater. 2023, 13, 2301864. [Google Scholar] [CrossRef]
  195. He, S.; Hao, S.; Fan, L.; Liu, K.; Cai, C.; Wolverton, C.; Zhao, J.; Liu, Q. Efficient Solar Spectrum-Like White-Light Emission in Zinc-Based Zero-Dimensional Hybrid Metal Halides. Adv. Opt. Mater. 2023, 11, 2300218. [Google Scholar] [CrossRef]
  196. 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. Int. Ed. 2021, 60, 19653–19659. [Google Scholar] [CrossRef] [PubMed]
  197. Worku, M.; Tian, Y.; Zhou, C.; Lee, S.; Meisner, Q.; Zhou, Y.; Ma, B. Sunlike White-Light-Emitting Diodes Based on Zero-Dimensional Organic Metal Halide Hybrids. ACS Appl. Mater. Interfaces 2018, 10, 30051–30057. [Google Scholar] [CrossRef] [PubMed]
  198. Jiang, X.; Chen, Z.; Tao, X. (1-C5H14N2Br)2MnBr4: A Lead-Free Zero-Dimensional Organic-Metal Halide with Intense Green Photoluminescence. Front. Chem. 2020, 8, 352. [Google Scholar] [CrossRef] [PubMed]
  199. Ma, W.; Liang, D.; Qian, Q.; Mo, Q.; Zhao, S.; Cai, W.; Chen, J.; Zang, Z. Near-Unity Quantum Yield in Zero-Dimensional Lead-Free Manganese-Based Halides for Flexible X-Ray Imaging with High Spatial Resolution. eScience 2023, 3, 100089. [Google Scholar] [CrossRef]
  200. 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, 282–291. [Google Scholar] [CrossRef]
  201. 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]
  202. Chen, D.; Dai, F.; Hao, S.; Zhou, G.; Liu, Q.; Wolverton, C.; Zhao, J.; Xia, Z. Crystal Structure and Luminescence Properties of Lead-Free Metal Halides (C6H5CH2NH3)3MBr6 (M = Bi and Sb). J. Mater. Chem. C 2020, 8, 7322–7329. [Google Scholar] [CrossRef]
  203. 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]
  204. 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] [PubMed]
  205. Wang, H.; Wang, C.; Wang, Y.; Sun, M.; Zhao, G. Efficient White Light Emission of 0D Lead-Free Indium-Based Halide Perovskite and the Intermediate State Promotion Mechanism of the Nonadiabatic Transition to Self-Trapped Exciton via Antimony(III) Cation Doping. J. Phys. Chem. C 2023, 127, 11616–11622. [Google Scholar] [CrossRef]
  206. 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]
  207. Li, Z.; Li, Y.; Liang, P.; Zhou, T.; Wang, L.; Xie, R.-J. Dual-Band Luminescent Lead-Free Antimony Chloride Halides with Near-Unity Photoluminescence Quantum Efficiency. Chem. Mater. 2019, 31, 9363–9371. [Google Scholar] [CrossRef]
  208. He, Q.; Zhou, C.; Xu, L.; Lee, S.; Lin, X.; Neu, J.; Worku, M.; Chaaban, M.; Ma, B. Highly Stable Organic Antimony Halide Crystals for X-Ray Scintillation. ACS Mater. Lett. 2020, 2, 633–638. [Google Scholar] [CrossRef]
  209. Peng, Y.-C.; Zhou, S.-H.; Jin, J.-C.; Zhuang, T.-H.; Gong, L.-K.; Lin, H.-W.; Wang, Z.-P.; Du, K.-Z.; Huang, X.-Y. [PPh3H]2[SbCl5]: A Zero-Dimensional Hybrid Metal Halide with a Supramolecular Framework and Stable Dual-Band Emission. J. Phys. Chem. C 2022, 126, 17381–17389. [Google Scholar] [CrossRef]
  210. Lv, L.; Yang, H.; Cheng, X.; Lin, Y.; Chang, X.; Cheng, T.; Xie, Y.; Han, Y.; Li, J.; Yin, J.; et al. Yellow Phosphor Based on Zero-Dimensional Antimony Halide for White Light-Emitting Diodes. J. Mater. Chem. C 2023, 11, 13335–13342. [Google Scholar] [CrossRef]
  211. Chen, D.; Hao, S.; Zhou, G.; Deng, C.; Liu, Q.; Ma, S.; Wolverton, C.; Zhao, J.; Xia, Z. Lead-Free Broadband Orange-Emitting Zero-Dimensional Hybrid (PMA)3InBr6 with Direct Band Gap. Inorg. Chem. 2019, 58, 15602–15609. [Google Scholar] [CrossRef]
  212. Ben-Akacha, A.; Zhou, C.; Chaaban, M.; Beery, D.; Lee, S.; Worku, M.; Lin, X.; Westphal, R.; Ma, B. Mechanochemical Synthesis of Zero Dimensional Organic-Inorganic Metal Halide Hybrids. ChemPhotoChem 2021, 5, 326–329. [Google Scholar] [CrossRef]
  213. Wei, J.-H.; Liao, J.-F.; Zhou, L.; Luo, J.-B.; Wang, X.-D.; Kuang, D.-B. Indium-Antimony-Halide Single Crystals for High-Efficiency White-Light Emission and Anti-Counterfeiting. Sci. Adv. 2021, 7, eabg3989. [Google Scholar] [CrossRef] [PubMed]
  214. Park, N.-G.; Zhu, K. Scalable Fabrication and Coating Methods for Perovskite Solar Cells and Solar Modules. Nat. Rev. Mater. 2020, 5, 333–350. [Google Scholar] [CrossRef]
  215. Faure, B.; Salazar-Alvarez, G.; Ahniyaz, A.; Villaluenga, I.; Berriozabal, G.; De Miguel, Y.R.; Bergström, L. Dispersion and Surface Functionalization of Oxide Nanoparticles for Transparent Photocatalytic and UV-Protecting Coatings and Sunscreens. Sci. Technol. Adv. Mater. 2013, 14, 023001. [Google Scholar] [CrossRef] [PubMed]
  216. Tzounis, L.; Stergiopoulos, T.; Zachariadis, A.; Gravalidis, C.; Laskarakis, A.; Logothetidis, S. Perovskite Solar Cells from Small Scale Spin Coating Process towards Roll-to-Roll Printing: Optical and Morphological Studies. Mater. Today Proc. 2017, 4 Pt B, 5082–5089. [Google Scholar] [CrossRef]
  217. Li, F.; Bao, C.; Gao, H.; Zhu, W.; Yu, T.; Yang, J.; Fu, G.; Zhou, X.; Zou, Z. A Facile Spray-Assisted Fabrication of Homogenous Flat CH3NH3PbI3 Films for High Performance Mesostructure Perovskite Solar Cells. Mater. Lett. 2015, 157, 38–41. [Google Scholar] [CrossRef]
  218. Richards, D.; Burkitt, D.; Patidar, R.; Beynon, D.; Watson, T. Predicting a Process Window for the Roll-to-Roll Deposition of Solvent-Engineered SnO2 in Perovskite Solar Cells. Mater. Adv. 2022, 3, 8588–8596. [Google Scholar] [CrossRef]
  219. Razza, S.; Castro-Hermosa, S.; Di Carlo, A.; Brown, T.M. Research Update: Large-Area Deposition, Coating, Printing, and Processing Techniques for the Upscaling of Perovskite Solar Cell Technology. APL Mater. 2016, 4, 091508. [Google Scholar] [CrossRef]
  220. Peng, X.; Yuan, J.; Shen, S.; Gao, M.; Chesman, A.S.R.; Yin, H.; Cheng, J.; Zhang, Q.; Angmo, D. Perovskite and Organic Solar Cells Fabricated by Inkjet Printing: Progress and Prospects. Adv. Funct. Mater. 2017, 27, 1703704. [Google Scholar] [CrossRef]
  221. Sliz, R.; Czajkowski, J.; Fabritius, T. Taming the Coffee Ring Effect: Enhanced Thermal Control as a Method for Thin-Film Nanopatterning. Langmuir 2020, 36, 9562–9570. [Google Scholar] [CrossRef]
  222. Sumaiya, S.; Kardel, K.; El-Shahat, A. Organic Solar Cell by Inkjet Printing—An Overview. Technologies 2017, 5, 53. [Google Scholar] [CrossRef]
  223. Parida, B.; Singh, A.; Kalathil Soopy, A.K.; Sangaraju, S.; Sundaray, M.; Mishra, S.; Liu, S.; Najar, A. Recent Developments in Upscalable Printing Techniques for Perovskite Solar Cells. Adv. Sci. 2022, 9, 2200308. [Google Scholar] [CrossRef] [PubMed]
  224. Yoo, J.W.; Jang, J.; Kim, U.; Lee, Y.; Ji, S.-G.; Noh, E.; Hong, S.; Choi, M.; Seok, S.I. Efficient Perovskite Solar Mini-Modules Fabricated via Bar-Coating Using 2-Methoxyethanol-Based Formamidinium Lead Tri-Iodide Precursor Solution. Joule 2021, 5, 2420–2436. [Google Scholar] [CrossRef]
  225. Søndergaard, R.R.; Hösel, M.; Krebs, F.C. Roll-to-Roll Fabrication of Large Area Functional Organic Materials. J. Polym. Sci. Part B Polym. Phys. 2013, 51, 16–34. [Google Scholar] [CrossRef]
  226. Li, H.; Liu, H.; Shi, C.; Qin, J.; Fu, Y.; Song, Y.; Li, Y.; Ling, Z. Roll-To-Roll Fabricating MXene Membranes with Ordered Interlayer Distances for Molecule and Ion Separation. Adv. Mater. Interfaces 2023, 10, 2300301. [Google Scholar] [CrossRef]
  227. Bae, S.-R.; Heo, D.Y.; Kim, S.Y. Recent Progress of Perovskite Devices Fabricated Using Thermal Evaporation Method: Perspective and Outlook. Mater. Today Adv. 2022, 14, 100232. [Google Scholar] [CrossRef]
Figure 1. Cell efficiency chart by the US National Renewable Energy Laboratory in May 2023 [4]. This plot is courtesy of the National Renewable Energy Laboratory, Golden, CO, USA.
Figure 1. Cell efficiency chart by the US National Renewable Energy Laboratory in May 2023 [4]. This plot is courtesy of the National Renewable Energy Laboratory, Golden, CO, USA.
Materials 17 00491 g001
Figure 2. Bandgaps of several lead-free perovskites [17] compared to most studied lead-based ones. Reproduced with permission under the terms of the Creative Commons CC BY license.
Figure 2. Bandgaps of several lead-free perovskites [17] compared to most studied lead-based ones. Reproduced with permission under the terms of the Creative Commons CC BY license.
Materials 17 00491 g002
Figure 3. Classification and structures of perovskite and perovskite-related unit cell structures, where A is an organic/inorganic cation, M is a metal cation, X is a halide anion, Y is chalcogenide, and □ is a vacancy.
Figure 3. Classification and structures of perovskite and perovskite-related unit cell structures, where A is an organic/inorganic cation, M is a metal cation, X is a halide anion, Y is chalcogenide, and □ is a vacancy.
Materials 17 00491 g003
Figure 4. (a) Rock-salt-ordered DP structure [32]. Copyright © 2018 American Chemical Society. This publication is licensed under CC-BY-NC-ND. (b) Phase transitions from the rock-salt-ordered DP to simple perovskite and Jahn–Teller distorted double perovskite [29]. Copyright © 2022 American Chemical Society.
Figure 4. (a) Rock-salt-ordered DP structure [32]. Copyright © 2018 American Chemical Society. This publication is licensed under CC-BY-NC-ND. (b) Phase transitions from the rock-salt-ordered DP to simple perovskite and Jahn–Teller distorted double perovskite [29]. Copyright © 2022 American Chemical Society.
Materials 17 00491 g004
Figure 5. Space of octahedral and tolerance factor values of some halide DPs [35] © 2020 Wiley-VCH GmbH.
Figure 5. Space of octahedral and tolerance factor values of some halide DPs [35] © 2020 Wiley-VCH GmbH.
Materials 17 00491 g005
Figure 6. (a) General workflow machine learning in perovskite and related applications [40]. Reproduced with permission under the terms of the Creative Commons CC BY license. (b) Map of GGA-calculated bandgaps of 980 DB; the black plus signs indicate thermodynamically stable structures [38]. Copyright © 2020, American Chemical Society.
Figure 6. (a) General workflow machine learning in perovskite and related applications [40]. Reproduced with permission under the terms of the Creative Commons CC BY license. (b) Map of GGA-calculated bandgaps of 980 DB; the black plus signs indicate thermodynamically stable structures [38]. Copyright © 2020, American Chemical Society.
Materials 17 00491 g006
Figure 7. (a) Band alignment of the ordered, partially disordered, and fully disordered Cs2AgBiBr6 [47]. Copyright © 2017 American Chemical Society. (b) Formation of local domains leading to multiple emissive domains in Cs2AgBiBr6 [42]. Copyright © 2022 the authors. Published by American Chemical Society. This publication is licensed under CC-BY 4.0.
Figure 7. (a) Band alignment of the ordered, partially disordered, and fully disordered Cs2AgBiBr6 [47]. Copyright © 2017 American Chemical Society. (b) Formation of local domains leading to multiple emissive domains in Cs2AgBiBr6 [42]. Copyright © 2022 the authors. Published by American Chemical Society. This publication is licensed under CC-BY 4.0.
Materials 17 00491 g007
Figure 8. (a) White light PL in Sb3+/Bi3+ co-doped Cs2NaInCl6 microcrystals [51]. Copyright © 2021 American Chemical Society. (b) PLQY of pentavalent-doped Cs2NaInCl6 SCs [52] © 2023 Wiley-VCH GmbH.
Figure 8. (a) White light PL in Sb3+/Bi3+ co-doped Cs2NaInCl6 microcrystals [51]. Copyright © 2021 American Chemical Society. (b) PLQY of pentavalent-doped Cs2NaInCl6 SCs [52] © 2023 Wiley-VCH GmbH.
Materials 17 00491 g008
Figure 9. (a) Different approaches for the synthesis of DPs [63]. Copyright © 2023, AIP Publishing. This publication is licensed under CC-BY. (b) Publication statistics are achieved by text mining for different synthetic routes of numerous halide DPs [15]. Copyright © 2021 Elsevier Inc.
Figure 9. (a) Different approaches for the synthesis of DPs [63]. Copyright © 2023, AIP Publishing. This publication is licensed under CC-BY. (b) Publication statistics are achieved by text mining for different synthetic routes of numerous halide DPs [15]. Copyright © 2021 Elsevier Inc.
Materials 17 00491 g009
Figure 10. Two examples for the synthesis of NCs: (a) schematic illustration of the solution phase synthesis of Cs2AgBiBr6 NCs (hot injection route) [114]. © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Synthesis of Cs2AgBiX6 NCs by Creutz et al. [115]. Copyright © 2018, American Chemical Society. (c) TEM image of Cs2AgBiBr6 NCs prepared with method (a). (d,e) TEM images of Cs2AgBiBr6 NCs prepared with method (b).
Figure 10. Two examples for the synthesis of NCs: (a) schematic illustration of the solution phase synthesis of Cs2AgBiBr6 NCs (hot injection route) [114]. © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Synthesis of Cs2AgBiX6 NCs by Creutz et al. [115]. Copyright © 2018, American Chemical Society. (c) TEM image of Cs2AgBiBr6 NCs prepared with method (a). (d,e) TEM images of Cs2AgBiBr6 NCs prepared with method (b).
Materials 17 00491 g010
Figure 11. (a) Single-crystal structures at 298 K of the (001) layered double perovskites (BA = CH3(CH2)3NH3+). Insets show the Ag+ coordination sphere with selected bond distances in Å, t denotes terminal bromide, and b denotes bridging bromide [132]. Copyright © 2018, American Chemical Society. (b) [0,0] and [1/2,0] stacking patterns in DJ-layered DP derivatives [136]. Copyright © 2022, American Chemical Society.
Figure 11. (a) Single-crystal structures at 298 K of the (001) layered double perovskites (BA = CH3(CH2)3NH3+). Insets show the Ag+ coordination sphere with selected bond distances in Å, t denotes terminal bromide, and b denotes bridging bromide [132]. Copyright © 2018, American Chemical Society. (b) [0,0] and [1/2,0] stacking patterns in DJ-layered DP derivatives [136]. Copyright © 2022, American Chemical Society.
Materials 17 00491 g011
Figure 12. The different types of (001)-layered perovskite structures [23]. © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 12. The different types of (001)-layered perovskite structures [23]. © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Materials 17 00491 g012
Figure 13. SC structures viewed along the c-axis (a) (R-MPEA)2CuCl4; (b) (S-MPEA)2CuCl4 [130]; (c) UV-vis-NIR absorbance spectra; and (d) circular dichroism spectra of abovementioned chiral perovskite derivatives [130]. Copyright © 2020, American Chemical Society.
Figure 13. SC structures viewed along the c-axis (a) (R-MPEA)2CuCl4; (b) (S-MPEA)2CuCl4 [130]; (c) UV-vis-NIR absorbance spectra; and (d) circular dichroism spectra of abovementioned chiral perovskite derivatives [130]. Copyright © 2020, American Chemical Society.
Materials 17 00491 g013
Figure 14. (a) An anti-perovskite structure with the formula [MnX’4]XA3, where [MnX’4] represents B. Reprinted (adapted) with permission from {ACS Energy Lett. 2021, 6, 5, 1901–1911}. Copyright {2023} American Chemical Society [146]. (b) Anti-perovskite with a 0D MnCl4 tetrahedra in the middle representing B. Reprinted (adapted) with permission from {Inorg. Chem. 2023, 62, 7, 3202–3211}. Copyright {2023} American Chemical Society [144].
Figure 14. (a) An anti-perovskite structure with the formula [MnX’4]XA3, where [MnX’4] represents B. Reprinted (adapted) with permission from {ACS Energy Lett. 2021, 6, 5, 1901–1911}. Copyright {2023} American Chemical Society [146]. (b) Anti-perovskite with a 0D MnCl4 tetrahedra in the middle representing B. Reprinted (adapted) with permission from {Inorg. Chem. 2023, 62, 7, 3202–3211}. Copyright {2023} American Chemical Society [144].
Materials 17 00491 g014
Figure 16. Schematic figures of most solution-based deposition methods used for perovskite. Note that the setup for the different techniques can vary. (a) Spray coating; (b) bar coating; (c) blade coating; (d) slot-die coating; (e) spin coating; (f) gravure printing; (g) rotary screen printing; (h) inkjet printing.
Figure 16. Schematic figures of most solution-based deposition methods used for perovskite. Note that the setup for the different techniques can vary. (a) Spray coating; (b) bar coating; (c) blade coating; (d) slot-die coating; (e) spin coating; (f) gravure printing; (g) rotary screen printing; (h) inkjet printing.
Materials 17 00491 g016
Table 1. Generalized clinical symptoms of Pb poisoning in humans [13]. © 2020 by the authors. Licensee MDPI, Basel, Switzerland.
Table 1. Generalized clinical symptoms of Pb poisoning in humans [13]. © 2020 by the authors. Licensee MDPI, Basel, Switzerland.
Body Organ/SystemClinical Symptoms of Pb Poisoning
EyesBlindness of parts of visual field
Hallucinations
EarsHearing loss
MouthUnusual taste
Slurred speech
Blue line along the gum
KidneyStructural damage and failure
Changes in the excretory function
LiverJaundice
Lead-induced oxidative stress
Decreased liver function
Microvesicular and macrovesicular steatosis
Hemosiderosis and cholestasis
SkinPallor and/or lividity
Central nervous systemInsomnia
Loss of appetite
Decreased libido
Irritability
Cognitive deficits
Memory loss
Headache
Personality changes
Delirium
Coma
Encephalopathy
Reproductive organsSperm dysfunctions
Pregnancy complications
Preterm birth
Abdomen/StomachPain
Nausea
Diarrhea
Constipation
BloodAnaemia
GeneralMalaise
Fatigue
Weight loss
Neuro-muscularTremor
Pain
Delayed reaction times
Loss of coordination
Convulsions
Foot or ankle drop
Seizers
Weakness
BonesMineralizing bones and teeth
Decreased bone density
Table 2. Double perovskites grouped into three classes according to the valence electrons M and M′ ions and their potential application areas [19] © 2018 Elsevier Inc.
Table 2. Double perovskites grouped into three classes according to the valence electrons M and M′ ions and their potential application areas [19] © 2018 Elsevier Inc.
A2M(I)M(III)X6Optoelectronic PropertiesSynthesized Compounds(Potential) Application
Type I:
s2 + s2
direct bandgap
suitable bandgap values
strong light absorption
high electronic dimensionality
expected defect tolerance
(MA)2TlBiBr6(solar cell)
(light-emission device)
Type II:
s0 + s2
indirect bandgap
large bandgap values
reduced electronic dimensionality
long carrier lifetime
not good carrier transport
Cs2AgBiCl6
Cs2AgBiBr6
(MA)2AgBiBr6
(MA)2AgbiI6
Cs2AgSbCl6
(MA)2AgSbCl6
(MA)2KBiCl6
Cs2NaBiCl6
solar cell
X-ray detector
photocatalysis
(X-ray imaging)
Type III:
s0 + s0
direct bandgap
dipole-forbidden transition
large bandgap values
reduced electronic dimensionality
Cs2AgInCl6
(MA)2KGdCl6
(MA)2KYCl6
Cs2NaGaF6
photodetector
laser
light-emission device
Vacancy-ordereddirect bandgap
strong light absorption
existence of deep mid-gap defects
not good carrier transport
Cs2SnI6
Cs2PdBr6
Cs2Ti(Br/I)6
Cs2TeI6
solar cell
light-emission device
(X-ray imaging)
Table 3. Performance summary of powder based WLEDs [27] © 2023 the authors. Advanced functional materials published by Wiley-VCH GmbH.
Table 3. Performance summary of powder based WLEDs [27] © 2023 the authors. Advanced functional materials published by Wiley-VCH GmbH.
Device StructurePLQY [%]CIE (x, y)CCT [K]CRIRef.
UV LED/Cs2AgIn0.7Bi0.3Cl6 NCs/PMMA4(0.36, 0.35)444391[55]
UV LED/Cs2NaInCl6:2.5%Sb, 45%Tb, 3%Mn NCs74(0.41, 0.39)337189.2[56]
UV LED/Cs2(Na, Ag)InCl6:7.09%Ho3+57.09(0.39, 0.46)N/A75.4[57]
UV LED/Cs2Na0.4Ag0.6In0.995Bi0.005Cl6:Mn2+31.8(0.3784, 0.4216)4323.482.6[58]
UV LED/Cs2AgIn1−xBixCl639(0.417, 0.391)311985[59]
UV LED/Cs2Ag0.4Na0.6InCl6:1%Bi, 1%/BaMgAl10O17:Eu2+98.6(0.4, 0.38)443095.7[60]
UV LED/Cs2Ag0.4Na0.6InCl6:Bi, Gd/BaMgAl10O17:Eu2+87.57(0.3464, 0.3224)481893.9[60]
UV LED/Cs2Ag0.7Na0.3InCl6:Bi87.2(0.38, 0.44)434787.8[61]
UV LED/Cs2AgScCl6:0.05Bi60(0.366, 0.367)410096[62]
UV LED/Cs2Na0.4Ag0.6InCl6:Bi73.3(0.461, 0.443)2930–695784.8–97.1[63]
Table 4. Recent advances in DP NC-based optoelectronic devices and photocatalysts. (1) EL: electroluminescence (2) EQE: external quantum efficiency; (3) OQE: optical quantum efficiency; (4) PCE: power conversion efficiency.
Table 4. Recent advances in DP NC-based optoelectronic devices and photocatalysts. (1) EL: electroluminescence (2) EQE: external quantum efficiency; (3) OQE: optical quantum efficiency; (4) PCE: power conversion efficiency.
Luminescence
CompositionPL Peak [nm]PLQY (%)Ref.
Cs2AgInCl6:Mn62016.4[80]
Cs2AgBiCl6:In57036.6[81]
Cs2AgInCl6:Bi58011.4[82]
Cs2AgInCl6:Na/Bi60040[83]
Cs2AgBiCl6:Al63017.2[84]
Cs2NaInCl6:Ag53531.1[85]
Cs2NaInCl6:Sb/Mn455/62224[86]
Cs2NaInCl6:Sb44850[87]
Cs2KInCl6:Sb51595[77]
Cs2KInCl6:Sb/Mn510/63087[77]
Cs2NaYCl6:Sb46151.8[88]
Cs2NaTbCl6:SbMultiple peaks (green emission)24[79]
Cs2ZrCl644660.37[89]
Cs2HfCl662840.71[90]
MA4InCl7:Sb62084[91]
Light-emitting diodes
EL (1) peak [nm]Efficinecy
Cs2AgInCl6:Bi55758 cd/m2 (luminance); 0.01% (EQE (2))[92]
Cs2AgInCl6:Na/Bi/Tb6102793 cd/m2 (luminance);
0.76% (EQE (2))
[93]
Photocatalysis
CompositionReactionEfficiencyRef.
Cs2AgBiX6 (X = Cl, Br, I)CO2
photoreduction
0.035% (EQE (2))[74]
Cs2AgBiBr6NO removal97% (removal rate)[76]
Luminescent solar concentrators
Cs2AgInCl6:Na/Bi 21.2% (internal OQE (3))[75]
Solar cells
Cs2AgBiBr6 0.46% (PCE (4))[78]
Scintillator
Cs2NaTbCl6:Sb 140 nGyair/s
(detection limit)
[79]
Photodetector
Cs4Cd0.75Mn0.25Bi2Cl12 0.98 × 104 A W−1 (responsivity); 3 × 106% (EQE (2))[94]
Table 5. Electronically driven LEDs based on electronically 0D materials.
Table 5. Electronically driven LEDs based on electronically 0D materials.
MaterialMetalDevice StructureEQE (%)Emission Wavelength [nm]Luminescence [cd m−2]PLQY (%)CommentRef.
(ABI)4MnBr6MnITO/PEDOT:PSS/poly TPD/(ABI)4MnBr6/TPBi/LiF/Al9.8629470080PEO 1 wt% additive[174]
[PPh4]2[MnBr4]MnITO/PEDOT:PSS/TCTA:26DCZPPY (1:2):(Ph4P)2[MnBr4]/BmPyPb LiF/Al9.65182339 Active layer mixed with hole transport[175]
DBFDPO-MnBr2MnITO/MoO3/TAPC/TCTA (50 wt%) DBFDPO-MnBr2/TmPyPB/LiF/Al10.5552 81.4Active layer mixed with hole transport[176]
(MePPh3)2SbCl5SbITO/PEDOT:PSS/Poly-TP/TAPC:2,6-DCZPPY:(MePPh3)2SbCl5 (6:3:1)/TPBi/LiF/A3.1593350099.4Active layer mixed with hole and electron transport[177]
TPPcarzSbBr4SbITO/PEDOT:PSS/PVK/TPPcarzSbBr4/ZnO/LiF/Al5.12653 nm5957 cd m−293.8% [178]
Cs3Cu2I5:CsCu2I3CuITO/PEDOT:PSS, /Cs3Cu2I5:CsCu2I3/TmPyPB, /LiF/Al3.1565157030Mixture of 0D and 1D[179]
TEA2Cu2Br4CuITO/PEDOT:PSS/TEA2Cu2Br4/TPBi/LiF/Al0.114638594.73 [Cu2Br4] units [180]
Cs3CeBr6Ce(ITO)/ZnO/Al2O3/Cs3CeBr6/TCTA/TAPC/HAT-CN/Al0.42391/421 91 [181]
Table 6. A selection of electronic 0D materials for various applications. Abbreviations: SSL = solid-state lightning, * LTC = lowering temperature crystallization.
Table 6. A selection of electronic 0D materials for various applications. Abbreviations: SSL = solid-state lightning, * LTC = lowering temperature crystallization.
FormulaPLQY
(%)
FormValencyEm λ [nm]Isolated PolyhedralFWHM/
Stokes Shift
[nm]
Method/
Application/
Emission
Ref.
Cs3Cu2I558filmI440 [Cu2I3]81/155Thermal evaporation/LED/STE[187]
Cs3Cu2I598.71SCI443–456 [Cu2I5]3−99/135Annealed together/ /STE[188]
[(C3H7)4N]2Cu2I491.9SCI483, 637[CuI2] Solvent evaporation/ /STE[189]
(Gua)3Cu2I596SCI481[Cu2I5]3−125/156Heated solution/WLED/STE[190]
Cs4SnBr6 with SnF262.8powderII540 [SnBr6]4− Ball milling/WLED/STE[191]
(C4N2H14Br)4SnBr695SCII570[SnBr6]4−105/215Antisolvent diffusion/phosphor/STE[163]
(C4N2H14I)4SnI675SCII620[SnBr6]4−118/210Antisolvent diffusion/phosphor/[163]
(Bmpip)2SnBr475SC 666[SnBr4]2−69/326LTC/X-ray scintillator/[192]
Cs3ZnBr57.89NCII468 [ZnBr4]2−76/193Hot injection/SSL, display X-ray scintillator/STE[193]
(ABI)2ZnCl424.28SCII395[ZnCl4]2−92Solvent evaporation/Anti-Counterfeiting, X-ray scintillator, WLED/[194]
(PMA)2ZnCl437.2crystalII413, 440[ZnX4]2− LTC */WLED[195]
Cs3MnBr548 (1.29 with 2H2O)NCsII520 [MnBr4]2−43Hot injection/anti-counterfeiting/d-d transition, and possible STE[196]
(C9NH20)2MnBr481.08CrystalII528[MnBr4]2−64LTC/sensor/d–d transition[185]
(Ph4P)2MnBr498CrystalII512[MnBr4]2−48/52Diffusion, ball milling/WLED/d–d transition[197]
(1-mPQBr)2MnBr4 60.70 SCII 520 [MnBr4]2− 43 Solvent evaporation/ /d–d transition[198]
(TBA)2MnBr4 93.76 SCII 512 [MnBr4]2− 38.7/ Solvent evaporation/X-ray scintillator/d–d transition[199]
Cs3BiCl6 NCsIII391 [BiCl6]3−60/59Hot injection/ /STE[200]
Cs3Bi2Br9with oleic acid4.5NCsIII460 [BiBr6]3−45antisolvent injection[201]
(PMA)3BiBr6<1%CrystalIII405/510[BiBr6]3−153/160LTC/ /STEs[202]
Cs3InBr622.3NCs, lower PLQY as SCIII450 [InBr6]3−/75Hot injection/ /singlet and triplet STE/lighting, displays[203]
Cs2InBr5(H2O) 33SCIII695 [InBr5(H2O)]2−/340LTC/Water sensor/STE[204]
(DETA)InBr61.40
(24.12 Sb3+doping)
SCIII400, 500-700[InBr6]3−134/200LTC/ /STE[205]
K3SbCl6 Mn2+-doped 22.3NCsIII440 [SbCl6]3−102/120Hot injection/WLED/STE and defect[206]
(C9NH20)2SbCl598SCII590[SbCl5]2−119/210Antisolvent diffusion/phosphor/[163]
TEBA2SbCl598SC or powderIII590[SbCl5]2−140/250antisolvent diffusion or injection/WLED/STE[207]
(PPN)2SbCl598.1SCIII635[SbCl5]2−142/225Antisolvent injection/X-ray scintillator[208]
(PPh3H]2SbCl574.50SCIII653[SbCl5]2−/283Autoclave/ /singlet, triplet emission[209]
(NII)2SbCl588.9SCIII610[SbCl5]2−118/248Antisolvent diffusion/SSL/STE[210]
(Bmpip)2GeBr41SCII670[GeBr4]2−/330Antisolvent injection/SSL, X-ray scintillator/singlet, triplet exciton emission, lone pairs[192]
(PMA)3InBr635CrystalIII610[InBr6]3− LTC/WLED/STEs[211]
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

Dávid, A.; Morát, J.; Chen, M.; Gao, F.; Fahlman, M.; Liu, X. Mapping Uncharted Lead-Free Halide Perovskites and Related Low-Dimensional Structures. Materials 2024, 17, 491. https://doi.org/10.3390/ma17020491

AMA Style

Dávid A, Morát J, Chen M, Gao F, Fahlman M, Liu X. Mapping Uncharted Lead-Free Halide Perovskites and Related Low-Dimensional Structures. Materials. 2024; 17(2):491. https://doi.org/10.3390/ma17020491

Chicago/Turabian Style

Dávid, Anna, Julia Morát, Mengyun Chen, Feng Gao, Mats Fahlman, and Xianjie Liu. 2024. "Mapping Uncharted Lead-Free Halide Perovskites and Related Low-Dimensional Structures" Materials 17, no. 2: 491. https://doi.org/10.3390/ma17020491

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