Next Article in Journal
Pyro-Phototronic Effect Enhanced MXene/ZnO Heterojunction Nanogenerator for Light Energy Harvesting
Previous Article in Journal
Direct Current Triboelectric Nanogenerators, a Perspective from Material Selections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Influence of Binary Ligands in Designing Cesium Lead Halide (CsPbX3, X = Cl, Br, I) Perovskite Nanocrystals-Oleic Acid and Oleylamine

by
Ananthakumar Soosaimanickam
*,
Alejandro Saura
,
Noemi Farinós
and
Rafael Abargues
*
Institut de Ciencia dels Materials (ICMUV), Universitat de València, c/Catedràtic José Beltrán 2, 46980 Paterna, Spain
*
Authors to whom correspondence should be addressed.
Nanoenergy Adv. 2023, 3(4), 376-400; https://doi.org/10.3390/nanoenergyadv3040019
Submission received: 21 June 2023 / Revised: 2 November 2023 / Accepted: 3 November 2023 / Published: 15 November 2023

Abstract

:
The surface chemistry of cesium lead halide perovskite nanocrystals has been elaborately studied in recent years and has proved the critical role of carboxylic acids and amines in the formation and stability of the nanocrystals. Specifically, a slight change in the concentration and ratio of the frequently used oleic acid and oleylamine critically influences the resultant phase and physical properties. Thus, understanding the delicate surface of cesium lead halide perovskite nanocrystals mainly relies on chemical bonding and the dynamic ligand environment of these two organic species. In this aspect, this review summarizes experimental findings about the critical role of oleic acid and oleylamine on the nucleation, growth, stability, phase, and morphology of cesium lead halide perovskite nanocrystals and their effect under different circumstances.

Graphical Abstract

1. Introduction

Organic-inorganic hybrid perovskite (OIHP) and pure inorganic cesium lead halide perovskite nanocrystals (Cs LHP NCs) possess outstanding optical properties for several kinds of applications [1,2,3,4]. In particular, a narrow photoluminescence (PL) spectrum width, tunable optical spectra, and easy synthesis approaches result from Cs LHP NCs (CsPbX3 NCs, X = Cl, Br, I), positioning them as prominent candidates for optoelectronic applications. Moreover, a tunable bandgap from 1.77 to 3.1 eV is obtained by carefully tuning the halide composition [5]. Figure 1 shows a schematic representation of the surface termination of the CsPbBr3 NC lattice and a representation of colloidal CsPbX3 NCs and their PL spectra. Countless articles deal with synthesizing and fabricating Cs LHP thin films using solution-processing methods [6,7,8]. The high photoluminescence quantum yield (PLQY) of Cs LHP NCs is related to several factors, including the nature of precursors, halide concentration in solution, synthesis methods, organic ligands used in the reaction, and the composition of the finally formed NCs [9,10,11]. The synthesis of perovskite NCs through solution-processed methods has been accelerated to explore the possibilities of preparing them with different functionalities and morphologies. The ionic Pb-X bond, stronger ionic interaction of Cs+ ions with a PbX6 framework, and lower lattice energy result in Cs LHP NCs, which are very sensitive to ligand interaction [12,13,14]. Because of their ionic nature, the synthesis of LHP NCs in solution is fast, but the kinetics of the formation mechanism are less studied [15,16,17]. Moreover, the surface properties of Cs LHP NCs differ from the traditional II–VI and IV–VI chalcogenide semiconductor quantum dots (QDs). For example, unlike CdSe or PbS QDs, the surface of the CsPbX3 NCs is given as [CsPbX3] [PbX2] {AX}, where A = oleylammonium and X = halides and/or oleate [18]. Irrespective of the defect tolerance, the freshly prepared Cs LHP NCs have surface defects due to the deficiency of Pb2+ and halide ions. These defects must be passivated with cationic and anionic ligands like aliphatic amines (i.e., oleylamine (OAm)) and carboxylic acids (i.e., oleic acid (OA)). These organic molecules bind to the surface of Cs LHP NCs, forming a protective layer. In general, ligands play an active role in controlling the size, shape, and polydispersity of NCs. Additionally, they provide stability and rule their surface properties and reactivity. Moreover, the selection of solvents and ligands could influence the reaction rate and growth of the Cs LHP NCs [5,19,20]. It is well known that OA and OAm are suitable for preparing Cs LHP NCs using hot-injection and ligand-assisted reprecipitation (LARP) methods. In the presence of amines and carboxylic acids, the optical properties and phase formation of perovskite NCs are affected. Importantly, the role of carboxylic acids and amines in the structural framework of Cs LHP NCs is being characterized by different spectroscopy techniques by several research groups. Among others, binary ligands system such as oleic acid (OA) and oleylamine (OAm) play a significant role in stabilizing Cs LHP NCs and enhancing their optical properties. The ionized state of these ligands offers multiple roles in providing surface passivation and the dissolution of precursors. Even in NC films, it is important to understand the critical role of OA and OAm to elucidate optoelectronic devices’ charge transport and efficiency. Although these two ligands are extensively investigated in preparing traditional semiconductor NCs, understanding their chemistry is essential because of the unique surface properties of Cs LHP NCs. In particular, the binding of OA and OAm to nanocrystals (NCs) is highly susceptible to the solvent used during the purification process. Compared with their interaction with traditional semiconductor NCs, they are more labile on the Cs LHP NCs surface. This is also helpful in the ligand-exchange approach [21] to improve the optical properties and the efficiency of devices like LEDs, solar cells and photodetectors [20,22,23,24,25]. There are different strategies used to introduce and eliminate the OA/OAm ligand pair that significantly influence the structural and optical properties of Cs LHP NCs. For instance, an excess of OAm could turn the CsPbBr3 NCs into the lead-depleted non-luminescent perovskite phase, Cs4PbBr6 [26,27]. Also, the excess of ligands could induce different morphologies in Cs LHP NCs.
OA and OAm also direct the self-assembly of NCs to form well-ordered thin films [28,29]. The Bronsted acid–base equilibria between OA and OAm greatly influence the synthesis and assembly of Cs LHP NCs. Also, the presence of these ligands on the NC surface plays a key role in charge transport properties. The inclusion or removal of OA and OAm is useful in tuning the structural and optical properties of the Cs LHP NCs. A few research groups have spectroscopically confirmed the specific attachment and structural transformation of Cs LHP NCs under the influence of OA and OAm [30,31]. On one hand, these ligands show excellent surface capping properties due to their large size. On the other hand, large ligands strongly hinder the charge transport properties in NC films. Therefore, a common approach used in chalcogenide QDs to enhance the performance of optoelectronic devices is to carry out a ligand exchange with shorter ligands to improve charge carrier mobility [22,23,32,33]. To understand more about the crucial role of OA and OAm, this review summarizes important contributions rendered by these binary ligands to the growth, physical, and chemical properties of colloidally synthesized Cs LHP NCs.

2. Significance of Oleic Acid (OA) and Oleylamine (OAm) in the Formation and Stabilization of Cs LHP NCs

Capping ligands are essential for synthesizing, growing, stabilizing and fabricating densely packed, structured, super-lattice NC films. Figure 2 shows CsPb(Cl, Br)3 NC synthesis using hot injection and the ligand-assisted room-temperature precipitation method (LARP). According to Protesescu et al., who first demonstrated the hot-injection synthesis of Cs LHP NCs, OA and OAm are essential for dissolving the precursors and stabilizing the resulting NCs [34]. In the LARP approach, adding OA and OAm enhances the solubility of the precursors up to 50 times in dimethyl formamide (DMF) [5]. Also, to increase the reaction temperature, the concentration of OA/OAm is essential for dissolving PbX2 [35]. Since ligands play a critical role in the stabilization and properties of any NC, it is essential to know their fundamental role in remodifying the surface of Cs LHP NCs. The reactivity of Cs LHP NCs is much larger than that of II–VI or IV–VI chalcogenide semiconductor NCs. Therefore, the attachment of OA and OAm is more crucial in Cs LHP NCs for improving stability. This is mainly attributed to the extreme ionic character of the LHP NCs. The ionicity (difference in Pauling electronegativity) of Cs-X (~1.9–2.4) and Pb-X (~0.8–1.3) is much larger than that of chalcogenide semiconductors like (Pb/Cd)-(S/Se/Te) (~0.2–0.9) [36]. Higher ionicity in bonds reduces the activation energy required for ion migration within the structure, thereby increasing reactivity values. Here, ligands must be used to minimize reactivity by avoiding ion migration, in which an LHP NC releases or captures species from the environment.
In the hot-injection method, the high-temperature synthesis of Cs LHP NCs is typically carried out in the presence of OAm and OA and a non-coordinating solvent, 1-octadecene (ODE) [37]. Here, a preheated solution of Cs-oleate is injected into the PbX2/OAm/OA mixture, and the extremely fast crystallization leads to the formation of Cs LHP NCs. In the LARP method, high-boiling-point solvents (i.e., DMF and dimethyl sulfoxide (DMSO)) are used as solvents of the precursors at room temperature. Here, irrespective of the type of solvents, OAm and OA are commonly used to stabilize NCs, but they have different roles. Generally, OAm enhances the dissolution of precursors and controls the nucleation and growth of NCs in solution, whereas OA avoids the agglomeration of the synthesized NCs [38]. A density functional theory (DFT)–based calculation indicates that the chain length of OA and OAm is estimated as 24.34 and 23.85 A° [39]. This longer chain length of OA/OAm pair greatly influences the structural and morphological properties of Cs LHP NCs.
According to HSAB theory, Pb2+ is an intermediate Lewis acid, expected to interact with hard bases like oleate (OA) and OAm. Therefore, the interaction between OA, OAm and Pb2+ results in (1) the direct binding of OAm with Pb2+ through donating lone-pair electrons from nitrogen, (2) the direct binding of OA with Pb2+ to form a Pb-oleate complex, and (3) the formation of Pb-oleate coordination due to the substitution of halides [40]. However, an acid–base reaction between OA and OAm occurs during the synthesis. Here, OAm is protonated to form an oleylammonium cation (OAm+) and OA deprotonates to form an oleate anion (OA), as shown in Equation (1). A dynamic acid–base equilibrium between OAm+ and OA occurs, forming a hydrogen-bond acid–base pair. The equilibrium between OAm and OA can be represented by Equation (1) [27].
R-N H 2 + R-COOH R-N H 3 + + R-CO O
Thus, the resulting OAm+/OA ligand pair forms stable colloidal Cs LHP NCs through this dynamic ligand binding effect [38]. In detail, the OAm+ binds with surface halides by hydrogen bonding, whereas OA binds with the uncoordinated surface Pb2+ or Cs+ and forms Pb-O or Cs-O bonds [41]. Later, it was observed that the OAm+/OA ligand pair plays a critical role in controlling the structural and optical properties of Cs LHP NCs. Moreover, the metal–ligand interaction in Cs LHP NCs is not strong enough when OAm and OA ligands are used separately. The experimental results from hot-injection and LARP approaches indicate that only using OA does not lead to the synthesis of Cs LHP NCs. In contrast, employing OAm alone leads to the synthesis of Cs LHP NCs, but these NCs exhibit reduced stability compared to the OA/OAm-capped NCs [42,43,44].

2.1. Structural Modification of Cs LHP NCs Using OA/OAm Ligand Pair

The OAm/OA ligand pair plays an active role in the synthesis and passivation of Cs LHP NCs. The OAm/OA ligand pair can be located between NCs to stabilize the colloids sterically [30,45,46,47]. However, OAm and OA show a different type of interaction with LHP NCs, which can dramatically affect the LHP stability. OAm+ cations can partially replace the Cs+ from the structure of Cs LHP NCs or cap the Cs+ vacancies to control the phase, morphology, and dimensions of the Cs LHP NCs. Therefore, OAm+ can reconstruct PbX42− octahedra by stabilizing the {110} facet [48], which strongly influences the formation of the perovskite phase [49]. It should be noted that the unprotonated OAm can donate two electrons to the metal center, which also serves for passivation. Moreover, alkylammonium ions can also interact with halide on the surface of the LHP NCs to immobilize them [50]. The concentration of protonated OAm+ can be mainly enhanced at lower pH by increasing the concentration of OA [27].
While OAm+ substitutes Cs+, the carboxylate group in OA can bind with Pb2+ by partially removing the halides [40], leading to a loss of stability. If the deprotonation of OA is hindered, for example, through surface-binding ligands such as ionic liquids, surface passivation is enhanced [51]. Also, OAm+ can degrade the NCs by removing lead-oleate from the surface through the formation of an OA/OAm ligand pair [52,53]. Due to the surface interaction of LHP NCs with the OA/OAm ligand pair being relatively weak, their stability in solution is compromised. Consequently, they exhibit outstanding optical properties over short periods.
Despite the substitution of Cs+ ions with incoming OAm+ ions, the surface termination of Cs LHP NCs is predominantly associated with the presence of metal halides: either PbX2 or CsX termination [54]. Through experimental and theoretical model analysis, Ravi et al. proposed that CsBr terminates the surface of the CsPbBr3 NCs. Concurrently, OAm+ ions also participate in hydrogen bonding interactions with the surface bromide ions [55], as confirmed by X-ray scattering and NMR [56,57]. In general, varying the concentration of OA and OAm drastically alters the optical properties, morphology, and crystalline phase of the Cs LHP NCs. In addition to these, the solubility of PbX2 is also affected concerning the concentration of ligands. For example, increasing the concentration of OA/OAm in the reaction enhances the precipitation temperature of PbBr2 up to 290 °C [27].
Compared with the mere addition of OA or OAm, the OA/OAm ligand pair improves the PL properties. Here, the dynamic equilibrium between OA and OAm generates a shell around the NCs, which enhances the PL intensity. Interestingly, when a pre-synthesized halide salt of OAm (i.e., OAmBr) is directly used as a reactant, the resulting CsPbBr3 NCs show a near-unity PLQY [9]. Similarly, different kinds of long-chain ammonium halides help to promote the optical properties of Cs LHP NCs [58]. The post-addition of OAmBr also alters the morphology and resists the phase transformation of CsPbBr3 NCs [59]. It should be noted that the vapor pressure values of OAm and OA are ~10 and ~103 Pa at 120 °C. However, this does not affect the removal or evaporation of these ligands in films [38].
As stated before, spectroscopic observations are indeed helpful in exploring the existence of OA and OAm with Cs LHP NCs. The bonding nature of OA and OAm can be explored through core-loss electron energy loss spectroscopy. Using this method, Kirakosyan et al. found the σ* peak at 292 eV for the OAm-capped NCs and the π* feature of the C K-edge of OAm [49]. NMR also helps to identify the interaction of OA/OAm with the Cs LHP NCs. Because of the coordination between ligands, the corresponding resonance appears broader and shifted compared to their spectrum in the free state [54]. Also, the condensation reaction between carboxylic acid and amine results in amide formation, which is detected from the appearance of a peak at δ~3.10 ppm [60]. The dynamic equilibrium between OA and OAm in Cs LHP NCs is schematically given in Figure 3.
The interaction of OA/OAm with Cs LHP NCs’ surfaces can be detected and confirmed through 1H-NMR analysis. The OA/OAm-capped LHP NCs usually show a peak at ~1.66 ppm owing to the existence of β-CH2 groups. The free OAm on the surface reveals an upfield resonance at δ = 5.55 ppm, whereas the bonded OAm shows a downfield resonance at δ = 5.68 ppm [30,49]. The protonation and deprotonation between this ligand pair could be observed from the α1-CH2 1H signal at 12.16 ppm. Furthermore, the protonation of OAm by OA and halo acids can be confirmed through the downward shift of the α–CH2 resonance in the NMR spectra of the OAm [61].
Figure 4 provides a schematic representation illustrating the modification of OAmBr on the surface of CsPbBr3 NCs. Additionally, it includes images of the OA/OAm mixture, the chemical structure of oleylammonium oleate, and their respective 1H-NMR spectra. Although the OA/OAm ligand pair serves as a capping agent, the combination of OA and OAm is also associated with surface Pb2+ etching when OAm and OA are protonated and deprotonated, respectively. As a result, the acid–base equilibrium in the solution is broken and becomes immobilized within the ligand shell. [49]. Also, OA could coordinate with Pb2+, but the reaction kinetics is slower than that of OAm with Pb2+ [40].

2.2. Morphological Changes of Cs LHP NCs in the Presence of OA/OAm

The structural influence of the OA/OAm ligand pair on Cs LHP NCs also leads to the induction of morphological changes. Apart from reaction conditions such as temperature and precursors, the concentration of the OA/OAm ligand pair can produce different morphologies [62]. It is well known that OA and OAm preferentially bind on the Cs-Br and Pb-Br terminated NC surface. The concentration of OA/OAm should be optimal to dissolve the PbBr2, but an excess of concentration could dissolve partially Cs LHP NCs [35]. Therefore, it is essential to meticulously control the concentration and ratio of the OA/OAm ligand pair to achieve a narrow distribution of a specific morphology. For example, upon reaching the optimal concentration of the OA/OAm pair ligand for solubilizing PbBr2, a narrow size distribution of CsPbBr3 nanocubes is observed [27]. The competition between Cs+ and OAm+ dramatically accelerates the growth of Cs LHP NCs with different morphologies. Using halo acids such as HCl, HBr, and HI results in the formation of nanowires (NWs) and nanosheets (NSs) [61,63]. Here, the protonation of OAm by halo acids sterically influences the growth conditions, which direct different morphologies. The advantage of using halo acids is that they can readily supply the halide ions, which react with cationic species. It should be noted that the formation of halo acids also occurs when PbX2 reacts with OA, which can also influence the NC morphology. Additionally, the protonation of OAm by halo acid (HBr) slows down the growth of NCs in the vertical direction, which leads to the control of the thickness of the CsPbBr3 nanoplatelets (NPLs) [64]. This kind of thickness control in CsPbBr3 NPLs has also been achieved by supplying halide (i.e., Br) to the OAm+ ions in the presence of benzoyl bromide [65]. Furthermore, with respect to the concentration of halo acids, the phase transformation of CsPbBr3 to CsPb2Br5 is demonstrated [66].
As stated earlier, it is possible to achieve different morphologies of Cs LHP NCs by simply varying the molar ratio of the OA/OAm ligand pair. Li et al. have observed the formation of CsPbBr3 NSs with a low amount of OA, whereas the formation of NWs is favored with a high amount of OA [67]. According to the authors, the high presence of OAm disfavors the growth of NCs in the (110) plane, which directs the formation of NPL assembly. In contrast, Almeida et al. have observed the formation of NPLs when the concentration of OA increases [27]. With high OA concentration, the acid–base equilibrium shifts towards OAm protonation, which directs NPL formation [36]. Ji et al. have synthesized CsPbBr3 NCs (NSs and NWs) with different morphologies by simply varying the ratio of OA/OAm [68]. Here, uniform NW distribution is achieved using an OA/OAm ratio of 2. Interestingly, Li et al. have synthesized ultrasmall OA/OAm-capped CsPbBr3 QDs (~4.5 nm) with high exciton binding energy (Eb = 268.7 meV) using a high OA concentration without 1-ODE [69]. However, by varying the concentration of OA/OAm in the presence of 1-ODE, Liang et al. achieved QDs (2.4 nm) for the ratio 0.6 mL/0.3 mL, whereas NPLs and NSs morphologies were achieved for the ratios 0.5 mL/0.5 mL and 0.2 mL/0.8 mL [70]. Although these investigations deal with the influence of ligands, variation in the reaction temperature could also significantly influence the morphology of Cs LHP NCs. Zhang et al. found that under optimized experimental conditions, the formation of CsPbBr3 NWs takes place at 150 °C, whereas over 180 °C is required to synthesize CsPbI3 NWs [71]. In this case, with respect to the reaction time, the morphological conversion of CsPbBr3 NCs from nanocubes, NWs, NSs and larger size crystals is observed. The formation mechanism of the NWs of Cs LHP NCs consists of different postulates such as dipole-driven attachment, template-directed growth, surfactant-directed growth, seed-mediated growth, etc. Additionally, the modification in morphology can be achieved irrespective of the synthesis method followed. For example, Seth et al. have varied the concentration of OA/OAm to synthesize CsPbX3 NCs in toluene using the LARP method [43]. In this study, it is observed that a higher concentration of OAm (70 μL) leads to larger nanocubes, whereas a lower concentration (20 μL) leads to the formation of NR morphology. Interestingly, under prolonged reaction time, these NRs are converted to NWs, which clearly emphasizes the influence of solvents and ligands on morphology transformation. Most of the synthesis protocols of the NWs of Cs LHP NCs are accompanied by OA/OAm ligand pair in the presence of various solvents and ligands. This includes LARP [72], hot injection [71], ultrasonication [73], microfluidic reactor [74], etc. Moreover, this kind of NW can also be achieved through a post-synthetic treatment approach. For example, Yang et al. have achieved ultrathin CsPbBr3 NWs (2.5 nm width) by post-synthetically treating Cs4PbBr6 NCs with PbBr2 in short-chain ligands (hexanoic acid, octylamine) [75]. Here, it is proposed that the PbBr2-ligand intermediate, which serves as a lamellar template in the reaction, direct this NWs formation. Similarly, Fanizza et al. have achieved different morphologies of CsPbBr3 NCs (nanocubes, NPLs, NWs) through the post-synthetic addition of OA and OAm [76]. Generally, the post-synthetic surface modification of Cs LHP NCs is influenced by several parameters, such as ligands, the polarity of the solvent, temperature, light, pressure, etc. [77]. The formation of different morphologies of CsPbBr3 NCs under the post-addition of OA and OAm is schematically provided in Figure 5. Other than this, replacing OA/OAm with short- or long-chain ligands also critically influences the morphology of the Cs LHP NCs [78]. For instance, Pan et al. have modified the chain length of the amine and carboxylic acids in place of OA/OAm in the reaction [31]. In this case, at 170 °C, it was found that decreasing the chain length of carboxylic acids leads to the formation of nanocubes with different sizes, while in the case of amines, NPLs with different thicknesses are achieved. A similar kind of approach using the solvothermal synthesis method (at 100 °C) results in the formation of NWs with different aspect ratios [79]. Along with morphology variation, the thickness of the NWs could be controlled by varying the chain length of the acid/amine in the reaction, which results in NWs with a high PLQY (up to 77%) with significant shifts in the absorption and emission spectra [80]. These results clearly envisage the critical role of the acid/amine pair concerning the morphological variation of Cs LHP NCs. When OA is replaced by a fatty acid—for example, olive oil, the nanocubic morphology of Cs LHP NCs can still be achieved [81]. This is because, at high temperatures, the OA from olive oil produces Cs-oleate and Pb-oleate in the reaction medium [82].
The inclusion of excess OA also favors the improvement of PL properties, because the addition of OA reduces the desorption of oleate ions, enhancing the PL characteristics [83]. Despite their equal contribution, it was found that high polar solvents such as acetone selectively remove OAm+ from the NC surface, leaving OA on it [31]. Thus, weakly bounded OAm+ ions direct the structure and morphology of Cs LHP NCs. At the same time, OA remains a dense layer on the surface and stimulates the two-dimensional (2D) growth pattern to form NPLs with different thicknesses [17,49]. Here, the OA excess is assumed to exfoliate the structural units, forming different monolayers. Likewise, OA’s rich condition in the reaction favors the formation of thicker NPLs, which strongly influence the PL spectra [65]. In addition to this, OA helps to solubilize the cesium source. The solubilization rate depends on the value of heat of formation energy. For example, it is observed that OA and OAm could solubilize cesium acetate (CsCH3COO (CsOAc)). In contrast, CsBr and CsCl are only partially solubilized due to the higher heat of the formation energy [4]. In another study, it is evidently observed that replacing Cs2CO3 with CsOAc is very helpful in achieving Cs LHP NCs with different morphologies in the presence of OA/OAm [31]. The morphological analysis of the formation of CsPbX3 NWs under the influence of different amounts of HX (X = Cl, Br, I) in the presence of OA/OAm and a comparison graph of the influence of the ratio of OA/OAm on the PL spectra of the CsPbBr3 NPLs are given in Figure 6.
Another intriguing aspect of the OA/OAm ligand pair is its capacity for the self-assembly of LHP NCs. The use of OAm and OA direct self-assembled super-lattice structures in NC films due to the solvent, ligand, and light-induced interaction. Self-assembly has been carried out in the nanocubes [44,84], NPLs [85,86,87], NWs [88,89], and nanorods (NRs) [90] of Cs LHP NCs. Generally, the self-assembly of Cs LHP NCs in solution takes place when the dispersed NCs are kept for a long time. Here, together with OA/OAm, the dielectric constant and polarity of the dispersing solvent significantly influence the rate of the self-assembly process. In the case of NC films, with respect to the solvent and ligands, the self-assembly is influenced. Moreover, surface treatment using smaller molecules could also motivate self-assembly, owing to the close contact of NCs. For example, when CsPbBr3 nanocubes are treated by thiocyanate (SCN) molecules, the oleyl molecules are partially replaced by SCN, which brings the nanocubes very close to each other, favoring the self-assembly [91]. The interparticle spacing between the OA/OAm-capped NCs in such a supramolecular assembly is approximately 11.4 ± 0.1 nm in the colloidal state [92]. Liu et al. used a solar simulator with the power of 1.7 suns (170 mW/cm2) to remove OA/OAm ligands from the as-synthesized CsPbBr3 NCs (PL lifetime~2 ns) [88]. In this case, the removal of OA/OAm accelerated the formation of NWs (PL lifetime~125 ns) through self-assembly. Similarly, by varying the intensity of light (532 nm and 670 nm), Pramanik et al. achieved nanobelts and nanoplatelet morphologies from OA/OAm-capped CsPbI3 QDs [87]. This kind of UV-light irradiation on the 1-alkynyl acid-capped CsPbBr3 QDs could induce the formation of self-assembled nanostructures through a catalysis reaction in the presence of a reaction intermediate, CsBr [93]. It is possible to achieve self-assembly in Cs LHP NCs by the sole use of OAm as a ligand [94]. Nevertheless, OA could support the achievement of a long-range ordered NC assembly [95]. It is demonstrated that the high concentration of OAm could lead from NSs to NWs through a dipole–dipole interaction in the presence of OA [68]. Bi et al. have indicated that during the self-assembly process, OA is gradually desorbed and under the influence of OAm, the growth of CsPbI3 NWs takes place [72]. Thus, the structural participation of OAm+ ions also influences the morphology of Cs LHP NCs. During synthesis, the self-assembled one-dimensional CsPbBr3 NWs can be achieved by forming lamellar structures [70,96]. Moreover, the OAm-rich synthesis reaction at 180 °C could direct the formation of self-assembled CsPbBr3 NWs [97]. In this case, the self-assembly of NWs takes place by connecting [PbBr6]4− octahedrons under a longer reaction period. Furthermore, it is demonstrated that together with these factors, surface halide vacancies of Cs LHP NCs could promote self-assembly [72,98]. The self-assembly of Cs LHP NCs into NWs could be distinguished through a shift in the PL spectrum depending on the aspect ratio. Interestingly, it was found that while aging CsPbBr3 NCs under ambient conditions, the top surface of the resultant NC super-lattice is converted to bulk-like particles through a coalescence process [99]. Along with morphology, aging Cs LHP NCs could also induce a phase transformation. By aging the pre-synthesized cubic CsPb(BrxI1−x)3 NCs in an equal volume of toluene and chloroform, it is possible to achieve orthorhombic CsPb(BrxI1−x)3 NWs [100]. In this case, the spontaneous coalescence of NCs leads to NWs through an oriented attachment process. Other than these results, as discussed previously, the post-synthetic transformation of Cs LHP NCs to NWs is also demonstrated. Factors such as dipole–dipole, hydrophobic and van der Waals interactions of the NC capping agents could lead to super-lattice nanostructures in Cs LHP NCs [28,68]. As mentioned, self-assembly is also directed by solvents with different polarity. If the interparticle spacing between the NCs is less than twice the OA/OAm ligand length (~4 nm), the solvent could induce self-assembly [101]. Mehetor et al. observed a polarity-dependent self-digestive conversion of CsPbBr3 quantum wires to quantum rods [90]. In this case, the authors observed a rapid conversion of this morphology change in chloroform (CHCl3) but a very slow conversion in hexane. Similarly, Soetan et al. observed a one-dimensional super-lattice nanostructure transformation of CsPbBr3 NCs when redispersed in toluene (non-polar), whereas two-dimensional short-range nanostructures were observed in polar solvents [101]. A mixed-solvent system for ex: toluene/chloroform was also found to be useful in expediting the self-assembly of CsPb(BrxI1−x)3 nanocubes to NWs under aging [99]. Here, along with the partial removal of ligands, the strong dipolar interaction of nanocubes leads to the formation of NWs. Similar to these investigations, there are different studies elucidating the role of solvents and ligands in the self-assembly of Cs LHP NCs [102,103,104,105,106]. Because of polarity, alcohols play an important role in the formation of super-lattice nanostructures of Cs LHP NCs. For instance, synthesizing CsPbBr3 NCs in the presence of methanol/cesium cholate directs the formation of NRs through oriented attachment [47]. Like methanol, the addition of ethanol into pre-synthesized CsPbX3 NPLs was found to direct the formation of NWs [17]. It was also found that CsPbBr3 NPLs could spontaneously transform into nanobelts and nanotiles under prolonged storage [107]. This kind of spontaneous self-assembly is generally directed by NC surface ligands through interparticle interactions [108]. In most of these investigations, the desorption of OA/OAm under different conditions leads to super-lattice nanostructures of Cs LHP NCs. The schematic representations of the formation of the one-dimensional super-lattice assembly of OA/OAm-capped CsPbBr3 NCs, the two-dimensional OA-induced long-range assembly of CsPbBr3 NPLs, and the structural and morphological analysis of CsPbBr3 NCs and NWs are given in Figure 7.

2.3. Phase Transformation of Cs LHP NCs in the Presence of OA/OAm

A variation in the concentration of OA/OAm will additionally impact the resulting phase of the Cs LHP NCs. Even a slight variation in their concentration alters the structural arrangement of atoms and directs the phase transformation. Two kinds of perovskite phases possibly form under the influence of OA/OAm, namely Cs4PbX6 and CsPb2Br5. Usually, a high OAm concentration with CsPbBr3 promotes the formation of the lead-depleted rhombohedral Cs4PbB6 phase [109,110,111]. This is actually a reversible transformation that takes place between corner shared PbX6 octahedra to isolated PbX6 octahedra. It is observed that the formation of the Cs4PbX6 occurs through forming halo-plumbate species in solution under a high Pb2+-ligand ratio [61]. While Cs4PbBr6 itself is non-luminescent, the presence of a small amount of CsPbBr3 introduces fascinating optical characteristics. This has been experimentally demonstrated by different research groups [112,113,114]. Jing et al. have adopted an amine-free synthesis approach to synthesize CsPbBr3 NCs [115]. In this case, the authors have observed that adding OAm with the pre-prepared CsPbBr3 NCs leads to Cs4PbBr6 phase transformation. This phase transformation by OAm could be accelerated in the presence of potential organic compounds such as thiol [109]. Here, the formation of Cs4PbBr6 NCs is confirmed through the emergence of the peak at ~315 nm in the UV-visible spectra. This phase conversion can be achieved by increasing the concentration of OA/OAm [27]. Furthermore, it is possible to achieve the same when the NCs are redispersed in hexane. The mechanism behind this phase transformation is still unclear. Accordingly, in most of the observations, it is stated that the dissolution-recrystallization accompanied by the influence of excess OAm is the reason for converting the luminescent to the non-luminescent phase [109,116]. Other possibilities, such as ligand environment and soft ligand templating, are also proposed as reasons for this phase transformation. Li et al. observed that when the synthesis reaction is carried out in an OAm-rich atmosphere, temperature and reaction time significantly influence the formation of perovskite NCs [97]. In this case, the authors achieved Cs4PbBr6 NCs at 160 °C and CsPbBr3 NCs at 180 °C. It was found that in contrast with CsPbBr3 NCs, where the substitution of Cs+ takes place by OAm+ cations, Cs4PbBr6 is comprised of both OAm and OA in a bonded state as an OAm+-OA complex [49]. The different roles of OAm and the phase transformation of CsPbBr3 NCs under the influence of OA/OAm with their structural, optical, and morphological characterization are combinedly given in Figure 8.
Structural models demonstrate the reversible transformation between luminescent CsPbBr3 and non-luminescent Cs4PbBr6 NCs. In general, Cs4PbBr6 NCs are obtained after adding an extra number of OA/OAm ligands once the CsPbBr3 NCs are synthesized. Reversibly, the Cs4PbBr6 NCs are transformed back to highly luminescent CsPbBr3 NCs after adding PbBr2. The CsPbBr3 NCs structure is characterized by corner-sharing [PbX6]4− octahedra with Cs+ ions filling the voids created by four neighboring [PbX6]4− octahedra. In the Cs4PbBr6 NCs structure, adjacent [PbBr6]4− octahedra do not share any corners and they are entirely decoupled in all directions. Thus, excess PbBr2 can be integrated into the crystal structure of Cs4PbBr6 NCs to form compact CsPbBr3 NCs [117].

3. Potential Ligands and Solvents Replacing OA and OAm: A Step towards Enhancing Optical Properties and Stability of Cs LHP NCs

Although OA and OAm are helpful for the passivation of Cs LHP NCs, their role as ligands is limited. When highly ionic Cs LHP NCs are prepared with OA/OAm ligands, their attachment with the surface is relatively weak, and they could easily detach even during the first purification cycle. This purification step mainly depends on the polarity of the solvents. For instance, solvents such as methyl acetate and ethyl acetate are much preferred compared with acetone, ethanol, and isopropanol. Purification by suitable solvents greatly assists in controlling ligand density in Cs LHP NCs. For example, purification using an ethyl acetate/hexane solvent mixture was found to precisely regulate OA/OAm ligand density in Cs LHP NCs with excellent PL properties [118,119,120]. When the prepared crude solid Cs LHP NCs undergo multiple purification cycles, the detachment of ligands substantially impacts the colloidal stability as well as the lifetime and PLQY of Cs LHP NCs. The weak binding nature is also attributed to the low binding constant value (6020 M−1) of OAm+ over short-chain amine ions [121]. Moreover, the removal of OA/OAm is also very sensitive to thermal annealing and UV exposure [38].
Recently, this problematic issue has been solved using strong binding ligands. Compared with OA/OAm, these ligands resist the antisolvents and do not detach from the NC surface during purification. Thus, the colloidal stability and the optical properties are well preserved even after purification, and highly luminescent NCs are achieved. Several kinds of non-traditional ligands and potential additives have been explored recently, and their effect on the NC surface is documented [122,123]. For example, phosphonic ligands are identified as potential candidates that resist solvent washing and produce smaller NCs [124,125]. Here, the P-O bond is relatively stronger, with the Pb2+ leading to a network with octylphosphonate molecules through hydrogen bonds. Additionally, unlike OA/OAm reversible dynamic binding ligands, the irreversible phosphonate binding results in an increase in PL intensity [57,126]. In addition to binding ability, it is demonstrated that oleylphosphonic acid (pKa = 2) could effectively replace DDA+ ions from the surface of CsPbBr3 NCs without affecting the optical properties [127]. Although these ligands provide stability, OAm is still used to dissolve the Pb precursor, PbBr2. However, this use of OAm can be efficiently replaced by TOPO, as demonstrated experimentally [126]. Likewise, a complete amine-free synthesis of CsPbBr3 NCs is reported using TOP as a solvent with HBr or 1-bromopropane as a bromide source [128,129].
Interestingly, when trioctylphosphine (TOP) is used as a solvent to synthesize Cs LHP NCs, TOP is oxidized to TOPO and prevents the oxidation of Pb2+ [130]. Similarly, it is observed that using a quaternary ammonium halide salt, didodecyldimethylammonium bromide (DDAB), instead of OAm significantly improves the PLQY of Cs LHP NCs, up to near-unity [4,127,131]. For this, DDAB in a non-polar solvent, specifically in toluene, is used to carry out the surface modification. Here, DDAB specifically heals the CsX surface vacancies, resulting in an impressive PLQY. Thus, DDAB enables a reduction in the number of ligands present on the surface of Cs LHP NCs without impacting their optical properties. This treatment is particularly useful for improving the charge transport in optoelectronic devices like LEDs [132,133]. Because of the shorter chain length over the OA/OAm ligand pair, DDAB specifically replaces OAm and could efficiently bind on the surface of the Cs LHP NCs [134,135]. The efficiency of DDAB treatment could be enhanced by the inclusion of a co-additive like ZnBr2 [134]. However, improving the PLQY using DDAB should be carried out with careful observations, since this may lead to a phase transformation [136,137]. Similar to DDAB, different kinds of organic and inorganic compounds are used for the post-synthetic modification of Cs LHP NCs in solution or in solid-state in order to regulate the density of the OA/OAm binary ligands. Ligand exchange in perovskite NCs is a complex process that requires careful consideration of the desired modifications and their impact on stability and performance. Researchers continue to investigate and refine ligand-exchange techniques to unlock the full potential of perovskite NCs for various applications in optoelectronics and beyond.
During the synthesis reaction, the formation of OAmBr frequently leads to a decrease in surface Br ions on CsPbBr3, leading to a deterioration in optical properties. To prevent this, substituting OAm with HBr is a viable approach. Akhil et al. have adopted this approach and demonstrated the completely amine-free synthesis of CsPbBr3 NCs through the hot-injection method [129,138]. Here, it is observed that an increase in Br concentration results in the formation of CsPbBr3 NCs. The authors have observed that excessive halide concentration in the reaction helps to achieve CsPbBr3 NCs with high stability. Furthermore, the NC films demonstrated enhanced stability over a three-month period, suggesting the potential of halo acids as a promising alternative to the OA/OAm ligand pair for achieving highly stable and luminescent Cs LHP NCs.
Recently, zwitterionic ligands have been delivering promising results in the stabilization and maintenance of the optical properties of Cs LHP NCs. Zwitterionic ligands are molecules featuring both positively charged groups (e.g., ammonium) and negatively charged groups (e.g., carboxylates, sulfonates, and phosphonates). These bidentate ligands demonstrate a more robust interaction with the surface of NCs and exhibit a lower detachment rate when compared to the OA/OAm ligand pair. Consequently, this allows for the long-term stability of optical properties [139]. In particular, zwitterionic ligands possess strong chelating ability and can stand several purification cycles with different solvents—for example, ethyl acetate (EtOAc) [140,141]. The existence of acidic and basic functional groups in the same molecule helps to maintain the acid–base equilibria and stabilize the Cs LHP NCs with high durability. An ideal zwitterionic ligand should have a lower diffusion coefficient value and should possess higher binding energy. Moreover, zwitterionic ligands have a different electrokinetic potential value (ζ) to the OA/OAm ligand pair, which derives interesting results in surface passivation [25]. For example, OAm/OA-capped CsPbBr3 QDs show a ζ potential of 9.17 meV, whereas a zwitterionic ligand betaine (BT)-capped QDs show 25.8 meV [142].
Krieg et al. first used different kinds of zwitterionic ligands (sulfobetaine, phosphocholine, soy lecithin and γ-aminoacids) to stabilize orthorhombic CsPbBr3 NCs [140,143]. It is also possible to synthesize zwitterionic ligands through an in-situ chemical reaction in the presence of OAm [144]. Among these, sulfobetaine is the most used zwitterionic ligand and is characterized by both a quaternary ammonium group (positively charged) and a sulfonate group (negatively charged). The sulfobetaine ligand is quite efficient in passivating Mn2+-doped alloyed Cs LHP NCs with a high PLQY and could minimize the reabsorption effect [145]. Lee et al. used a sulfobetaine ligand, 3-(dodecyldimethyl-ammonio)-propane-1-sulfonate (12-SBE), to passivate CsPbBr3 NCs [146]. In this case, the 12-SBE modified NCs showed a near-unity PLQY, long PL lifetime, less agglomeration, and high stability. Although sulfobetaine efficiently binds to the Cs LHP NCs, TOPO is necessary to achieve monodispersed, highly stable CsPbBr3 NCs [147]. Unfortunately, the sulfobetaine-capped CsPbBr3 NCs show worse charge transport properties and a lack of electronic coupling compared to the DDAB-capped CsPbBr3 NCs [139,148,149]. This highlights the need for the molecular engineering of zwitterionic ligands. Various kinds of zwitterionic ligands such as aromatic [141], polymeric [150,151,152], photo-crosslinkable [153], amino acids [154], modular zwitterion functionalized polymers [155], carbazole [156], and quaternary ammonium halide-based zwitterionic ligands [157] show an important role in stabilizing highly luminescent Cs LHP NCs are explored. Interestingly, Zhu et al. have proposed that both surface-selective and non-selective ligands are necessary to control the morphology and stability of CsPbBr3 NCs [157]. The authors have demonstrated that zwitterionic ligands are surface-selective to achieve anisotropic nanostructures (nanocubes, NPLs and NRs), and didecylamine (DDA)-based surface non-selective ligands could assist for it along with stability. In most cases, the OA/OAm ligands of Cs LHP NCs are post-synthetically exchanged by zwitterionic ligands via a partial ligand exchange, and the resultant NC properties are analyzed. Figure 9 provides a schematic overview of the interaction between OA/OAm and zwitterionic ligands with the perovskite NC surface. It also highlights the properties of zwitterionic ligands in solution and in the solid state. Additionally, the figure offers a collective representation of the chemical structure, synthesis process, and key findings related to the PLQY and the ligand-exchange ratio for 12-SBE-capped CsPbBr3 NCs.
Similar to zwitterionic ligands, multidentate ligands are also helpful in stabilizing Cs LHP NCs with improved robustness. Multidentate ligands strongly coordinate with the Cs LHP NC surface, forming a shell to protect the moisture-sensitive Cs LHP NCs. For example, a silane-based branched ligand such as n-propyltrimethoxysilane-dimethyloctadeyclammonium bromine (PDB) was found to polymerize on the surface of the CsPbBr3 QDs and impart high stability [158]. Similarly, 3-mercaptopropyltrimethoxysilane (MPTMS) in 1-octanethiol could induce a hybrid polymerization for the stabilization of CsPbBr3 NCs [159]. Also, growing CsPbBr3 NCs in the micelles of a polymeric multidentate ligand, polystryrene-block-poly-2-vinyl pyridine (PS-b-P2VP), was found to be useful in reducing the polar solvent diffusion into the Cs LHP NCs [160]. Multidentate ligands such as ethylenediamine tetraacetic acid (EDTA) [161,162], trithiocarbonate terminated polymeric ligand [163], poly-ethylene glycol (PEG)-based polymeric molecule [151,164], polyethyleneimine [165], quaternary ammonium- or imidazolium-based polysalt ligand [166,167], picolinate [168], sulphur-tributylphosphine (S-TBP) bidentate ligand [169], tetramethylthiuram disulfide (TMTD) [170], etc., have proven their strong binding ability with the surface of Cs LHP NCs. Zeng et al. used bis(2,4,4-trimethylpentyl) phosphonic acid to protonate a multidentate amine ligand, N′-(2-aminoethyl)-N′-hexadecylethane-1,2-diamine (AHDA) [171]. These protonated AHDA-capped CsPbI3 NCs showed lower diffusion coefficient (66 ± 3.4 μm2 s−1) and higher binding energy value (2.36 eV) over traditional OAm-capped NCs. This shows that novel ligands with a higher binding energy value could be resistant to traditional washing solvents of Cs LHP NCs and retain long-term optical properties. From the application point of view, the surface-modified Cs LHP NCs using zwitterionic and multidentate ligands are beneficial for photocatalysis applications, as evidenced by their high stereoselectivity in C-C oxidative coupling reactions [152,172] and also in the fabrication of light-emitting diodes (LEDs) [146,165,170].
Other than these systems, several kinds of post-treatment approaches using a wide variety of compounds have been demonstrated to strip the native OA/OAm ligand pair and/or to passivate the surface defects of Cs LHP NCs [11,46,173]. These surface treatment processes impressively enhance the PLQY of Cs LHP NCs. All these findings clearly indicate the emergence of potential, alternative anchoring ligands for designing surface-tailored Cs LHP NCs by replacing the OA/OAm ligand pair. Figure 10 presents three key aspects: the surface ligand content of AHDA and OA/OAm-capped CsPbI3 NCs with respect to purification cycles; a schematic illustration depicting the interaction of capping ligands with the surface of perovskite NCs; and diffusion coefficient values for OA/OAm- and AHDA-capped CsPbI3 NCs. Additionally, it shows the chemical structure and a schematic representation of picolinate capping on the surface of perovskite NCs.
In general, improving the photostability and thermal stability of perovskite NCs is indeed a crucial area of research and development. However, their stability issues have been a significant challenge that needs to be addressed for widespread commercialization. Apart from the use of ligands, there are some strategies and recommendations to enhance the stability of Cs LHP NCs:
(i)
Encapsulation and passivation: One of the primary reasons for perovskite instability is exposure to moisture, oxygen, and light. Encapsulation techniques involving protective coatings or capping ligands can shield Cs LHP NCs from these environmental factors. The passivation of surface defects can also improve stability [22].
(ii)
Composition engineering: Modifying the chemical composition of Cs LHP NCs by incorporating elements with higher stability, such as lead-free perovskites or mixed-halide perovskites, can enhance their resistance to degradation.
(iii)
Structural modifications: Researchers have explored alternative perovskite structures, such as double perovskites and 2D perovskites, which may exhibit improved stability compared to traditional 3D perovskites.
(iv)
Doping: Introducing suitable dopants into Cs LHP NCs can improve their stability and performance. For example, incorporating small amounts of certain metal ions can passivate defects and reduce degradation.
(v)
Stabilizing additives: The use of stabilizing additives or polymer matrices in perovskite NC films can improve their resistance to environmental factors and enhance long-term stability [174].
Overall, addressing the stability challenges of perovskite NCs is a critical step toward unlocking their full potential for various applications. Research and development efforts focused on stability enhancement will play a pivotal role in the successful commercialization of perovskite-based technologies.

4. Conclusions and Summary

This discussion narrates the crucial role of OA/OAm in synthesizing Cs LHP NCs. These binary ligands help to fabricate highly crystalline, highly ordered nanostructures and films of Cs LHP NCs. All the existing analyses show that it is essential to understand the reaction chemistry to explore other possible roles of these binary ligands. It is possible to obtain all kinds of phases and morphologies of Cs LHP NCs by cleverly tuning the concentration and ratio of the OA/OAm in the reaction. The experimental findings also reveal the critical role of OAm in the structural framework of perovskite NCs, a decisive factor in achieving highly luminescent characteristics. Although OA/OAm plays an essential role in the synthesis, a clear understanding should be established to invest the knowledge of this into other types of perovskite NCs. Also, the active role of these binary ligands is not much studied for lead-free perovskite NCs and requires timely exploration.
Furthermore, multidentate ligands that could couple with OA and OAm to contribute strong binding ability should be developed. Also, promising ligands that help to avoid morphology and phase transformation in the presence of OA/OAm are helpful for long-term applications. In this view, zwitterionic ligands can be considered to develop Cs LHP NCs with excellent stability and superior optical properties. Replacing OA and OAm with alternative bidentate ligands is a relatively reasonable effort in stabilizing highly luminescent perovskite NCs. Moreover, it is essential to analyze the spectroscopic characteristics of NCs synthesized using OA- and OAm-free solvents to understand the nature of bonding. Still, the role of OA and OAm impart better control of the morphology and phase, as seen in this discussion. Future accomplishments in this view would reveal many interesting findings in determining the structure and phase of perovskite NCs, which is helpful for several promising applications.

Author Contributions

Conceptualization, A.S. (Ananthakumar Soosaimanickam) and R.A.; validation, A.S. (Alejandro Saura) and N.F. (Noemi Farinos); writing—original draft preparation, A.S. (Ananthakumar Soosaimanickam), A.S. (Alejandro Saura) and N.F.; writing—review and editing, A.S. (Ananthakumar Soosaimanickam), N.F. and R.A.; visualization, A.S. (Alejandro Saura) and R.A.; project administration, R.A.; funding acquisition, R.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the funding by MCIN/AEI/10.13039/501100011033 (Research projects no. PID2020-119628RB-C31, PID2020-119628RB-C32 and PID2019-104272RB-C53). This study forms part of the Advanced Materials Programme and was supported by MCIN with funding from the European Union Next Generation EU (PRTR-C17.I1) and by the Generalitat Valenciana (codes MFA/2022/040 and MFA/2022/007). The work was partially funded by MCIN/AEI through project TED2021-131600B-C32.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, Y.; Zhang, Y.; Kang, W.; Deng, N.; Yan, Z.; Sun, W.; Kang, X.; Ni, J. Progress in the preparation and application of CsPbX3 perovskites. Mater. Adv. 2022, 3, 4053–4068. [Google Scholar] [CrossRef]
  2. Aznar-Gadea, E.; Sanchez-Alarcon, I.; Soosaimanickam, A.; Rodriguez-Canto, P.J.; Perez-Pla, F.; Martinez-Pastor, J.P.; Abargues, R. Molecularly imprinted nanocomposites of CsPbBr3 nanocrystals: An approach towards fast and selective gas sensing of explosive taggants. J. Mater. Chem. C 2022, 10, 1754–1766. [Google Scholar] [CrossRef]
  3. Dong, H.; Ran, C.; Gao, W.; Li, M.; Xia, Y.; Huang, W. Metal halide perovskite for next-generation optoelectronics: Progresses and prospects. eLight 2023, 3, 3. [Google Scholar] [CrossRef]
  4. De Trizio, L.; Infante, I.; Manna, L. Surface chemistry of lead halide perovskite colloidal nanocrystals. Acc. Chem. Res. 2023, 56, 1815–1825. [Google Scholar] [CrossRef]
  5. Kirakosyan, A.; Kim, Y.; Sihn, M.R.; Jeon, M.-G.; Jeong, J.-R.; Choi, J. Solubility-controlled room-temperature synthesis of cesium lead halide perovskite nanocrystals. ChemNanoMat 2020, 6, 1863–1869. [Google Scholar] [CrossRef]
  6. Navarro-Arenas, J.; Suarez, K.; Chirvony, V.S.; Gualdron-Reyes, A.F.; Mora-Sero, I.; Martinez-Pastor, J. Sinlge-exciton amplified spontaneous emission in thin films of CsPbX3 (X = Br, I) perovskite nanocrystals. J. Phys. Chem. Lett. 2019, 10, 6389–6398. [Google Scholar] [CrossRef]
  7. Chen, Y.; Li, F.; Zhang, M.; Yang, Z. Recent progress on boosting the perovskite film quality of all-inorganic perovskite solar cells. Coatings 2023, 13, 281. [Google Scholar] [CrossRef]
  8. Ekanayaka, T.K.; Richmond, D.; McCormick, M.; Nandyala, S.R.; Helfrich, H.C.; Sinitskii, A.; Pikal, J.M.; Ilie, C.C.; Dowben, P.A.; Yost, A.J. Surface versus bulk state transistions in inkjet-printed all-inorganic perovskite quantum dot films. Nanomaterials 2022, 12, 3956. [Google Scholar] [CrossRef]
  9. Dutta, A.; Behera, R.K.; Pal, P.; Baitalik, S.; Pradhan, N. Near-unity photoluminescence quantum efficiency for all CsPbX3 (X = Cl, Br and I) perovskite nanocrystals: A generic synthesis approach. Angew. Chem. 2019, 58, 5552–5556. [Google Scholar] [CrossRef]
  10. Yang, D.; Li, X.; Wu, Y.; Wei, C.; Qin, Z.; Zhang, C.; Sun, Z.; Li, Y.; Wang, Y.; Zeng, H. Surface halogen compensation for robust performance enhancements of CsPbX3 perovskite quantum dots. Adv. Mater. 2019, 7, 1900276. [Google Scholar] [CrossRef]
  11. Gualdron-Reyes, A.; Masi, S.; Mora-Sero, I. Progress in halide-perovskite nanocrystals with near-unity photoluminescence quantum yield. Trends Chem. 2021, 3, 499–511. [Google Scholar] [CrossRef]
  12. Ibaceta-Jana, J.; Muydinov, R.; Rosado, P.; Mirhosseini, H.; Chug, M.; Nazarenko, O.; Dirin, D.N.; Heinrich, D.; Wagner, M.R.; Kuhne, T.D.; et al. Vibrational dynamics in lead halide hybrid perovskites investigated by Raman spectroscopy. Phys. Chem. Chem. Phys. 2020, 22, 5604–5614. [Google Scholar] [CrossRef]
  13. Jeong, D.-N.; Yang, J.-M.; Park, N.-G. Roadmap on halide perovskite and related devices. Nanotechnology 2020, 31, 152001. [Google Scholar] [CrossRef] [PubMed]
  14. Huang, Y.; Cohen, T.A.; Sperry, B.M.; Larson, H.; Nguyen, H.A.; Homer, M.K.; Dou, F.Y.; Jacoby, L.M.; Cossairt, B.M.; Gamelin, D.R.; et al. Organic building blocks at inorganic nanomaterial interfaces. Mater. Horiz. 2022, 9, 61–87. [Google Scholar] [CrossRef]
  15. Wang, X.; Liu, Y.; Liu, N.; Sun, R.; Zheng, W.; Liu, H.; Zhang, Y. Revisiting the nanocrystal formation process of zero-dimensional perovskite. J. Mater. Chem. A 2021, 9, 4658–4663. [Google Scholar] [CrossRef]
  16. Huang, H.; Feil, M.W.; Fuchs, S.; Debnath, T.; Richter, A.F.; Tong, Y.; Wu, L.; Wang, Y.; Doblinger, M.; Nickel, B. Growth of perovskite CsPbBr3 nanocrystals and their formed superstructures revealed by in situ spectroscopy. Chem. Mater. 2020, 32, 8877–8884. [Google Scholar] [CrossRef]
  17. Udayabhaskararao, T.; Kazes, M.; Houben, L.; Lin, H.; Oron, D. Nucleation, growth and structural transformations of perovskite nanocrystals. Chem. Mater. 2017, 29, 1302–1308. [Google Scholar] [CrossRef]
  18. Bodnarchuk, M.I.; Boehme, S.C.; ten Brinck, S.; Bernasconi, C.; Shynkarenko, Y.; Krieg, F.; Widmer, R.; Aeschlimann, B.; Gunther, D.; Kovalenko, M.V.; et al. Rationalizing and controlling the surface structure and electronic passivation of cesium lead halide nanocrystals. ACS Energy Lett. 2019, 4, 63–74. [Google Scholar] [CrossRef]
  19. Wei, S.; Yang, Y.; Kang, X.; Wang, L.; Huang, L.; Pan, D. Room temperature and gram-scale synthesis of CsPbX3 (X = Cl, Br, I) perovskite nanocrystals with 50-58% photoluminescence quantum yields. Chem. Commun. 2016, 52, 7265–7268. [Google Scholar] [CrossRef]
  20. Yu, W.; Wei, M.; Tang, Z.; Zou, H.; Li, L.; Zou, Y.; Yang, S.; Wang, Y.; Zhang, Y.; Li, X.; et al. Separating crystal growth from nucleation enables the in situ controllable synthesis of nanocrystals for efficient perovskite light-emitting diodes. Adv. Mater. 2023, 35, 2301114. [Google Scholar] [CrossRef]
  21. Abargues, R.; Navarro, J.; Rodriguez-Canto, P.; Maulu, A.; Sanchez-Royo, J.F.; Martinez-Pastor, J.P. Enhancing the photocatalytic properties of PbS QD solids: The ligand exchange approach. Nanoscale 2019, 11, 1978–1987. [Google Scholar] [CrossRef] [PubMed]
  22. Soosaimanickam, A.; Adl, H.P.; Chirvony, V.; Rodriguez-Canto, P.J.; Martinez-Pastor, J.P.; Abargues, R. Effect of alkali metal nitrate treatment on the optical properties of CsPbBr3 nanocrystal films. Mater. Lett. 2021, 305, 130835. [Google Scholar] [CrossRef]
  23. Arenas, J.N.; Soosaimanickam, A.; Adl, H.P.; Abargues, R.; Boix, P.P.; Rodriguez-Canto, P.J.; Martinez-Pastor, J.P. Ligand-length modification in CsPbBr3 perovskite nanocrystals and bilayers with PbS quantum dots for improved photodetection performance. Nanomaterials 2020, 10, 1297. [Google Scholar] [CrossRef] [PubMed]
  24. Adhikari, G.C.; Thapa, S.; Yue, Y.; Zhu, H.; Zhu, P. Near unity PLQY and high stability of barium thiocyanate based all-inorganic perovskites and their applications in white light-emitting diodes. Photonics 2021, 8, 209. [Google Scholar] [CrossRef]
  25. Dong, Y.; Wang, Y.-K.; Yuan, F.; Johnston, A.; Liu, Y.; Ma, D.; Choi, M.-J.; Chen, B.; Chekini, M.; Baek, S.-W.; et al. Bipolar-shell resurfacing for blue LEDs based on strongly confined perovskite quantum dots. Nat. Nanotechnol. 2020, 15, 668–674. [Google Scholar] [CrossRef]
  26. Akkerman, Q.A.; Park, S.; Radicchi, E.; Nunzi, F.; Mosconi, E.; De Angelis, F.; Brescia, R.; Rastogi, P.; Prato, M.; Manna, L. Nearly monodisperse insulator Cs4PbX6 (X = Cl, Br, I) nanocrystals, their mixed halide compositions, and their transformation into CsPbX3 nanocrystals. Nano Lett. 2017, 17, 1924–1930. [Google Scholar] [CrossRef]
  27. Almeida, G.; Goldoni, L.; Akkerman, Q.; Dang, Z.; Khan, A.H.; Marras, S.; Moreels, I.; Manna, L. Role of acid-base equilibria in the size, shape and phase control of cesium lead bromide nanocrystals. ACS Nano 2018, 12, 1704–1711. [Google Scholar] [CrossRef]
  28. Jana, A.; Meena, A.; Patil, S.A.; Jo, Y.; Cho, S.; Park, Y.; Sree, V.G.; Kim, H.; Im, H.; Taylor, R.A. Self-assembly of perovskite nanocrystals. Prog. Mater. Sci. 2022, 129, 100975. [Google Scholar] [CrossRef]
  29. Sanchez-Alarcon, R.I.; Noguera-Gomez, J.; Chirvony, V.S.; Pashei Adl, H.; Boix, P.P.; Alarcon-Flores, G.; Martinez-Pastor, J.P.; Abargues, R. Spray-driven halide exchange in solid-state CsPbX3 nanocrystal films. Nanoscale 2022, 14, 13214–13226. [Google Scholar] [CrossRef]
  30. De Roo, J.; Ibanez, M.; Geiregat, P.; Nedelcu, G.; Walravens, W.; Maes, J.; Martins, J.C.; Driessche, I.V.; Kovalenko, M.V.; Hens, Z. Highly dynamic ligand binding and light absorption coefficient of cesium lead bromide perovskite nanocrystals. ACS Nano 2016, 10, 2071–2081. [Google Scholar] [CrossRef]
  31. Pan, A.; He, B.; Fan, X.; Liu, Z.; Urban, J.J.; Alivisatos, A.P.; He, L.; Liu, Y. Insight into the ligand-mediated synthesis of colloidal CsPbBr3 perovskite nanocrystals: The role of organic acid, base, and cesium precursors. ACS Nano 2016, 10, 7943–7954. [Google Scholar] [CrossRef] [PubMed]
  32. Maulu, A.; Navarro-Arenas, J.; Rodriguez-Canto, P.J.; Sanchez-Royo, J.F.; Abargues, R.; Suarez, I.; Martinez-Pastor, J.P. Charge transport in trap-sensitized infrared PbS quantum-dot-based photoconductors: Pros and cons. Nanomaterials 2018, 8, 677. [Google Scholar] [CrossRef] [PubMed]
  33. Maulu, A.; Rodriguez-Canto, P.J.; Navarro-Arenas, J.; Abargues, R.; Sanchez-Royo, J.F.; Garcia-Calzada, R.; Martinez-Pastor, J.P. Strongly-coupled PbS QD solids by doctor blading for IR photodetection. RSC Adv. 2016, 6, 80201–80212. [Google Scholar] [CrossRef]
  34. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef]
  35. Shen, Z.; Zhao, S.; Song, D.; Xu, Z.; Qiao, B.; Song, P.; Bai, Q.; Cao, J.; Zhang, G.; Swelm, W. Improving the quality and luminescence performance of all-inorganic perovskite nanomaterials for light-emitting devices by surface engineering. Small 2020, 16, 1907089. [Google Scholar] [CrossRef]
  36. Toso, S.; Baranov, D.; Manna, L. Metamorphoses of cesium lead halide nanocrystals. Acc. Chem. Res. 2021, 54, 498–508. [Google Scholar] [CrossRef]
  37. Kumawat, N.K.; Liu, X.-K.; Kabra, D.; Gao, F. Blue perovskite light-emitting diodes: Progress, challenges and future directions. Nanoscale 2019, 11, 2109–2120. [Google Scholar] [CrossRef]
  38. Moyen, E.; Kanwat, A.; Cho, S.; Jun, H.; Aad, R.; Jang, J. Ligand removal and photo-activation of CsPbBr3 quantum dots for enhanced optoelectronic devices. Nanoscale 2018, 10, 8591–8599. [Google Scholar] [CrossRef]
  39. Zhang, Q.; Diao, F.; Xue, X.; Sheng, X.; Barba, D.; Wang, Y. Self-assembly of CsPbBr3 nanocubes into 2D nanosheets. ACS Appl. Mater. Interfaces 2021, 13, 44777–44785. [Google Scholar] [CrossRef]
  40. Huang, H.; Raith, J.; Kershaw, S.V.; Kalytchuk, S.; Tomanec, O.; Jing, L.; Susha, A.S.; Rogach, A.L. Growth mechanism of strongly emitting CH3NH3PbBr3 perovskite nanocrystals with a tunable bandgap. Nat. Comm. 2017, 8, 996. [Google Scholar] [CrossRef]
  41. Yin, J.; Yang, H.; Gutierrez-Arzaluz, L.; Zhou, Y.; Bredas, J.-L.; Bakr, O.M.; Mohammed, O.F. Luminescence and stability enhancement of inorganic perovskite nanocrystals via selective surface ligand binding. ACS Nano 2021, 15, 17998–18005. [Google Scholar] [CrossRef] [PubMed]
  42. Protesescu, L.; Yakunin, S.; Nazerenko, O.; Dirin, D.N.; Kovalenko, M.V. Low-cost synthesis of highly luminescent colloidal lead halide perovskite nanocrystals by wet ball milling. ACS Appl. Nano Mater. 2018, 1, 1300–1308. [Google Scholar] [CrossRef] [PubMed]
  43. Seth, S.; Samanta, A. A facile methodology for engineering the morphology of CsPbX3 perovskite nanocrystals under ambient condition. Sci. Rep. 2016, 6, 37693. [Google Scholar] [CrossRef]
  44. Yuan, Y.; Liu, Z.; Liu, Z.; Peng, L.; Li, Y.; Tang, A. Photoluminescence and self-assembly of cesium lead halide perovskite nanocrystals: Effects of chain length of organic amines and reaction temperature. Appl. Surf. Sci. 2017, 405, 280–288. [Google Scholar] [CrossRef]
  45. Stoffel, J.T.; Tsui, E.Y. Stability effects of selective anion abstraction from cesium lead halide perovskite nanocrystals. J. Phys. Chem. C 2023, 127, 2285–2293. [Google Scholar] [CrossRef]
  46. Dutt, V.G.V.; Akhil, S.; Mishra, N. Surface passivation strategies for improving photoluminescence and stability of cesium lead halide perovskite nanocrystals. ChemNanoMat 2020, 6, 1730–1742. [Google Scholar] [CrossRef]
  47. Chakrabarty, A.; Satija, S.; Gangwar, U.; Sapra, S. Precursor-mediated synthesis of shape-controlled colloidal CsPbBr3 perovskite nanocrystals and their nanofiber directed self-assembly. Chem. Mater. 2020, 32, 721–733. [Google Scholar] [CrossRef]
  48. Pradhan, N. Alkylammonium halides for facet reconstruction and shape modulation in lead halide perovskite nanocrystals. Acc. Chem. Res. 2021, 54, 1200–1208. [Google Scholar] [CrossRef]
  49. Kazes, M.; Udayabhaskararao, T.; Dey, S.; Oron, D. Effect of surface ligands in perovskite nanocrystals: Extending in and reaching out. Acc. Chem. Res. 2021, 54, 1409–1418. [Google Scholar] [CrossRef]
  50. Kirakosyan, A.; Chinh, N.D.; Sihn, M.R.; Jeon, M.-G.; Jeong, J.-R.; Kim, D.; Jang, J.H.; Choi, J. Mechanistic insight into surface defect control in perovskite nanocrystals: Ligands terminate the valence transition from Pb2+ to metallic Pb0. J. Phys. Chem. Lett. 2019, 10, 4222–4228. [Google Scholar] [CrossRef]
  51. Chen, T.; Xu, Y.; Xie, Z.; Jiang, W.; Wang, L.; Jiang, W. Ionic liquid assisted preparation and modulation of the photoluminescence kinetics for highly efficient CsPbX3 nanocrystals with improved stability. Nanoscale 2020, 12, 9569–9580. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, W.; Yun, R.; Liu, Y.; Zhang, X.; Yuan, M.; Zhang, L.; Li, X. Ligands in lead halide perovskite nanocrystals: From synthesis to optoelectronic applications. Small 2023, 19, 2205950. [Google Scholar] [CrossRef]
  53. Yassitepe, E.; Yang, Z.; Voznyy, O.; Kim, Y.; Walters, G.; Castaneda, J.A.; Kanjanaboos, P.; Yuan, M.; Gong, X.; Fan, F.; et al. Amine-free synthesis of cesium lead halide perovskite quantum dots for efficient light-emitting diodes. Adv. Funct. Mater. 2016, 26, 8757–8763. [Google Scholar] [CrossRef]
  54. Smock, S.R.; Chen, Y.; Rossini, A.J.; Brutchey, R.L. The surface chemistry and structure of colloidal lead halide perovskite nanocrystals. Acc. Chem. Res. 2021, 54, 707–718. [Google Scholar] [CrossRef] [PubMed]
  55. Ravi, V.K.; Santra, P.K.; Joshi, N.; Chugh, J.; Singh, S.K.; Rensmo, H.; Ghosh, P.; Nag, A. Origin of the substitution mechanism for the binding of organic ligands on the surface of CsPbBr3 perovskite nanocubes. J. Phys. Chem. Lett. 2017, 8, 4988–4994. [Google Scholar] [CrossRef]
  56. Bertolotti, F.; Nedelcu, G.; Vivani, A.; Cervellino, A.; Masciocchi, N.; Guagliardi, A.; Kovalenko, M.V. Crystal structure, morphology and surface termination of cyan-emissive, six-monolayers-thick CsPbBr3 nanoplatelets from X-ray total scattering. ACS Nano 2019, 13, 14294–14307. [Google Scholar] [CrossRef]
  57. Chen, Y.; Smock, S.R.; Flintgruber, A.H.; Perras, F.A.; Brutchey, R.L.; Rossini, A.J. Surface termination of CsPbBr3 perovskite quantum dots determined by solid-state NMR spectroscopy. J. Am. Chem. Soc. 2020, 142, 6117–6127. [Google Scholar] [CrossRef]
  58. Ng, C.K.; Wang, C.; Jasieniak, J.J. Synthetic evolution of colloidal metal halide perovskite nanocrystals. Langmuir 2019, 35, 11609–11628. [Google Scholar] [CrossRef]
  59. Park, J.; Park, S.; Cho, S.; Kim, Y.; Kim, H.; Jeong, S.; Woo, J.Y. Air-stable cesium lead bromide perovskite nanocrystals via post-synthetic treatment with oleylammonium bromides. New J. Chem. 2022, 46, 19514–19522. [Google Scholar] [CrossRef]
  60. Grisorio, R.; Di Clemente, M.E.; Fanizza, E.; Allegretta, I.; Altamura, D.; Striccoli, M.; Terzano, R.; Giannini, C.; Irimia-Vladu, M.; Suranna, G.P. Exploring the surface chemistry of cesium lead halide perovskite nanocrystals. Nanoscale 2019, 11, 986–999. [Google Scholar] [CrossRef]
  61. Qu, R.; Zou, L.; Qi, X.; Liu, C.; Zhao, W.; Yan, J.; Zhang, Z.; Wu, Y. Changing the shape and optical properties of CsPbBr3 perovskite nanocrystals with hydrohalic acids using a room-temperature synthesis process. CrystEngComm 2022, 22, 1853–1857. [Google Scholar] [CrossRef]
  62. Mourdikoudis, S.; Menelaou, M.; Fiuza-Maneiro, N.; Zheng, G.; Wei, S.; Perez-Juste, J.; Polavarapu, L.; Sofer, Z. Oleic acid/oleylamine ligand pair: A versatile combination in the synthesis of colloidal nanoparticles. Nanoscale Horiz. 2022, 7, 941–1015. [Google Scholar] [CrossRef]
  63. Amgar, D.; Stern, A.; Rotem, D.; Porath, D.; Etgar, L. Tunable length and optical properties of CsPbX3 (X = Cl, Br, I) nanowires with a few unit cells. Nano Lett. 2017, 17, 1007–1013. [Google Scholar] [CrossRef] [PubMed]
  64. Akkerman, Q.A.; Motti, S.G.; Kandada, A.R.S.; Mosconi, E.; D’Innocenzo, V.; Bertoni, G.; Marras, S.; Kamino, B.A.; Miranda, L.; De Angelis, F.; et al. Solution synthesis approach to colloidal cesium lead halide perovskite nanoplatelets with monolayer-level thickness control. J. Am. Chem. Soc. 2016, 138, 1010–1016. [Google Scholar] [CrossRef]
  65. Do, M.; Kim, I.; Kolaczkowski, M.A.; Kang, J.; Kamat, G.A.; Yuan, Z.; Barchi, N.S.; Wang, L.-W.; Liu, Y.; Jurow, M.J.; et al. Low-dimensional perovskite nanoplatelet synthesis using in situ photophysical monitoring to establish controlled growth. Nanoscale 2019, 11, 17262–17269. [Google Scholar] [CrossRef] [PubMed]
  66. Li, J.; Zhang, H.; Wang, S.; Long, D.; Li, M.; Wang, D.; Zhang, T. Inter-conversion between different compounds of ternary Cs-Pb-Br system. Materials 2018, 11, 717. [Google Scholar] [CrossRef] [PubMed]
  67. Li, M.; Zhang, X.; Dong, T.; Wang, P.; Matras-Postolek, K.; Yang, P. Evolution of morphology, phase composition and photoluminescence of cesium lead bromine nanocrystals with temperature and precursors. J. Phys. Chem. C 2018, 122, 28968–28976. [Google Scholar] [CrossRef]
  68. Ji, Y.; Wang, M.; Yang, Z.; Ji, S.; Qiu, H. Nanowire-assisted self-assembly of one-dimensional nanocrystal superlattice chains. J. Mater. Chem. C 2019, 7, 8471–8476. [Google Scholar] [CrossRef]
  69. Li, H.; Zhao, G.; Song, B.; Dong, W.; Han, G. Suppressing growth strategy on synthesizing stable and high-quality blue-emitting CsPbBr3 quantum dots. ACS Appl. Opt. Mater. 2023, 1, 1263–1271. [Google Scholar] [CrossRef]
  70. Liang, Z.; Zhao, S.; Xu, Z.; Qiao, B.; Song, P.; Xu, X. Shape-controlled synthesis of all-inorganic CsPbBr3 perovskite nanocrystals with bright blue emission. ACS Appl. Mater. Interfaces 2016, 8, 28824–28830. [Google Scholar] [CrossRef]
  71. Zhang, D.; Eaton, S.W.; Yu, Y.; Dou, L.; Yang, P. Solution-phase synthesis of cesium lead halide perovskite nanowires. J. Am. Chem. Soc. 2015, 137, 9230–9233. [Google Scholar] [CrossRef] [PubMed]
  72. Bi, C.; Hu, J.; Yao, Z.; Lu, Y.; Binks, D.; Sui, M.; Tian, J. Self-assembled perovskite nanowire clusters for high luminance red light-emitting diodes. Adv. Funct. Mater. 2020, 30, 2005990. [Google Scholar] [CrossRef]
  73. Tong, Y.; Bladt, E.; Ayguler, M.F.; Manzi, A.; Milowska, K.Z.; Hintermayr, V.A.; Docampo, P.; Bals, S.; Urban, A.S. Highly luminescent cesium lead halide perovskite nanocrystals with tunable composition and thickness by ultrasonication. Angew. Chem. 2016, 55, 13887–13892. [Google Scholar] [CrossRef]
  74. Zhang, Z.; Liu, Y.; Geng, C.; Shi, S.; Zhang, X.; Bi, W.; Xu, S. Rapid synthesis of quantum-confined CsPbBr3 perovskite nanowires using a microfluidic reactor. Nanoscale 2019, 11, 18790–18796. [Google Scholar] [CrossRef] [PubMed]
  75. Yang, H.; Cai, T.; Dube, L.; Hills-Kimball, K.; Chen, O. Synthesis of ultrathin perovskite nanowires via a postsynthetic transformation reaction of zero-dimensional perovskite nanocrystals. Cryst. Growth Des. 2021, 21, 1924–1930. [Google Scholar] [CrossRef]
  76. Fanizza, E.; Cascella, F.; Altamura, D.; Giannini, C.; Panniello, A.; Triggiani, L.; Panzarea, F.; Depalo, N.; Grisorio, R.; Suranna, G.P.; et al. Post-synthesis phase and shape evolution of CsPbBr3 colloidal nanocrystals: The role of ligands. Nano Res. 2019, 12, 1155–1166. [Google Scholar] [CrossRef]
  77. Paul, S.; Acharya, S. Postsynthesis transformation of halide perovskite nanocrystals. ACS Energy Lett. 2022, 7, 2136–2155. [Google Scholar] [CrossRef]
  78. Aamir, M.; Adhikari, T.; Sher, M.; Khan, M.D.; Akhtar, J.; Nunzi, J.-M. Cesium lead halide perovskite nanostructures: Tunable morphology and halide composition. Chem. Rec. 2018, 18, 230–238. [Google Scholar] [CrossRef]
  79. Yu, M.; Zhang, D.; Xu, Y.; Lin, J.; Yu, C.; Fang, Y.; Liu, Z.; Guo, Z.; Tang, C.; Huang, Y. Surface ligand engineering of CsPbBr3 perovskite nanowires for high-performance photodetectors. J. Colloid Interface Sci. 2022, 608, 2367–2376. [Google Scholar] [CrossRef]
  80. Imran, M.; Di Stasio, F.; Dang, Z.; Canale, C.; Khan, A.H.; Shamsi, J.; Brescia, R.; Prato, M.; Manna, L. Colloidal synthesis of strongly fluorescent CsPbBr3 nanowires with width tunable down to the quantum confinement regime. Chem. Mater. 2016, 28, 6450–6454. [Google Scholar] [CrossRef]
  81. Welyab, G.; Abebe, M.; Mani, D.; Thankappan, A.; Thomas, S.; Aga, A.G.; Kim, J.Y. All-inorganic CsPbBr3 perovskite nanocrystals synthesized with olive oil and oleylamine at room temperature. Micromachines 2023, 14, 1332. [Google Scholar] [CrossRef] [PubMed]
  82. Ghori, A.; Midya, A.; Ray, S.K. Surfactant-induced anion exchange and morphological evolution for composition-controlled caesium lead halide perovskites with tunable optical properties. ACS Omega 2019, 4, 12948–12954. [Google Scholar] [CrossRef] [PubMed]
  83. He, H.; Tang, B.; Ma, Y. Controlled synthesis of quantum confined CsPbBr3 perovskite nanocrystals under ambient conditions. Nanotechnology 2018, 29, 055601. [Google Scholar] [CrossRef] [PubMed]
  84. Hudait, B.; Dutta, S.K.; Patra, A.; Nasipuri, D.; Pradhan, N. Facets directed connecting perovskite nanocrystals. J. Am. Chem. Soc. 2020, 142, 7207–7217. [Google Scholar] [CrossRef]
  85. Bekenstein, Y.; Koscher, B.A.; Eaton, S.W.; Yang, P.; Alivisatos, A.P. Highly luminescent colloidal nanoplates of perovskite cesium lead halide and their oriented assemblies. J. Am. Chem. Soc. 2015, 137, 16008–16011. [Google Scholar] [CrossRef]
  86. Xiao, X.; Li, Y.; Xie, R.-J. Blue-emitting and self-assembled thinner perovskite CsPbBr3 nanoplates: Synthesis and formation mechanism. Nanoscale 2020, 12, 9231–9239. [Google Scholar] [CrossRef]
  87. Pramanik, A.; Sinha, S.S.; Gates, K.; Nie, J.; Han, F.X.; Ray, P.C. Light-induced wavelength dependent self assembly process for targeted synthesis of phase stable 1D nanobelts and 2D nanoplatelets of CsPbI3 perovskites. ACS Omega 2023, 8, 13202–13212. [Google Scholar] [CrossRef]
  88. Liu, J.; Song, K.; Shin, Y.; Liu, X.; Chen, J.; Yao, K.X.; Pan, J.; Yang, C.; Yin, J.; Xu, L.-J.; et al. Light-induced self-assembly of cubic CsPbBr3 perovskite nanocrystals into nanowires. Chem. Mater. 2019, 31, 6642–6649. [Google Scholar] [CrossRef]
  89. Li, H.; Lu, W.; Zhao, G.; Song, B.; Dong, W.; Han, G. Synthesis of size-controllable self-assembled CsPbBr3 nanowires: Implications for photoemission with less recombination. ACS Appl. Nano Mater. 2022, 5, 5527–5534. [Google Scholar] [CrossRef]
  90. Mehetor, S.K.; Ghosh, H.; Pradhan, N. Blue-emitting CsPbBr3 perovskite quantum rods and their wide-area 2D self-assembly. ACS Energy Lett. 2019, 4, 1437–1442. [Google Scholar] [CrossRef]
  91. Patra, B.K.; Agrawal, H.; Zheng, J.-Y.; Zha, X.; Travesset, A.; Garnett, E.C. Close-packed ultrasmooth self-assembled monolayer of CsPbBr3 perovskite nanocubes. ACS Appl. Mater. Interfaces 2020, 12, 31764–31769. [Google Scholar] [CrossRef]
  92. van der Burgt, J.S.; Geuchies, J.J.; van der Meer, B.; Vanrompay, H.; Zanaga, D.; Zhang, Y.; Albrecht, W.; Petukhov, A.V.; Filion, L.; Bals, S.; et al. Cuboidal supraparticles self-assembled from cubic CsPbBr3 perovskite nanocrystals. J. Phys. Chem. C 2018, 122, 15706–15712. [Google Scholar] [CrossRef] [PubMed]
  93. Li, H.; Liu, X.; Ying, Q.; Wang, C.; Jia, W.; Xing, X.; Yin, L.; Lu, Z.; Zhang, K.; Pan, Y.; et al. Self-assembly of perovskite CsPbBr3 quantum dots driven by a photo-induced alkynyl homocoupling reaction. Angew. Chem. 2020, 59, 17207–17213. [Google Scholar] [CrossRef]
  94. Pan, J.; Huang, H.; Zhao, W.; Lin, D.; Nie, Z.; Zhong, J.; Ma, L.; Zhang, X. Efficient and stable self-assembly blue-emitting CsPbBr3 nanoplatelets with self-repaired surface defects. ACS Appl. Nano Mater. 2022, 5, 15062–15069. [Google Scholar] [CrossRef]
  95. Mehetor, S.K.; Ghosh, H.; Pradhan, N. Acid-amine equilibria for formation and long-range self-organization of ultrathin CsPbBr3 perovskite platelets. J. Phys. Chem. Lett. 2019, 10, 1300–1305. [Google Scholar] [CrossRef] [PubMed]
  96. Kostopoulou, A.; Sygletou, M.; Brintakis, K.; Lappas, A.; Stratakis, E. Low-temperature benchtop-synthesis of all-inorganic perovskite nanowires. Nanoscale 2017, 9, 18202–18207. [Google Scholar] [CrossRef] [PubMed]
  97. Li, H.; Lu, W.; Zhao, G.; Song, B.; Dong, W.; Han, G. Evolution and mechanism of cesium lead bromide nanostructures in oleylamine-rich system by hot-injection method. Adv. Mater. Interfaces 2023, 10, 2201916. [Google Scholar] [CrossRef]
  98. Pan, J.; Li, X.; Gong, X.; Yin, J.; Zhou, D.; Sinatra, L.; Huang, R.; Liu, J.; Chen, J.; Dursun, I.; et al. Halogen vacancies enable ligand-assisted self-assembly of perovskite quantum dots into nanowires. Angew. Chem. 2019, 131, 16223–16227. [Google Scholar] [CrossRef]
  99. Baranov, D.; Fieramosca, A.; Yang, R.X.; Polimeno, L.; Lerario, G.; Toso, S.; Giansante, C.; De Giorgi, M.; Tan, L.Z.; Sanvitto, D.; et al. Aging of self-assembled lead halide perovskite nanocrystal superlattices: Effects on photoluminescence and energy transfer. ACS Nano 2021, 15, 650–664. [Google Scholar] [CrossRef]
  100. Pradhan, B.; Mushtaq, A.; Roy, D.; Sain, S.; Das, B.; Ghorai, U.K.; Pal, S.K.; Acharya, S. Postsynthesis spontaneous coalescence of mixed-halide perovskite nanocubes into phase-stable single-crystalline uniform luminescent nanowires. J. Phys. Chem. Lett. 2019, 10, 1805–1812. [Google Scholar] [CrossRef]
  101. Soetan, N.; Erwin, W.R.; Tonigan, A.M.; Walker, D.G.; Bardhan, R. Solvent-assisted self-assembly of CsPbBr3 perovskite nanocrystals into one-dimensional super-lattice. J. Phys. Chem. C 2017, 121, 18186–18194. [Google Scholar] [CrossRef]
  102. Chen, Y.-L.; Hu, Y.-H.; Ma, L.; Zhang, X.-Y.; Zhao, N.-X.; Yang, X.; Zhang, Y.-S.; Gu, Y.-L.; Xu, S.-L.; Dong, X.; et al. Self-assembled CsPbBr3 quantum dots with wavelength-tunable photoluminescence for efficient active jamming. Nanoscale 2022, 14, 17900–17907. [Google Scholar] [CrossRef]
  103. Leng, J.; Wang, T.; Zhao, X.; Ong, E.W.Y.; Zhu, B.; de Andrew Ng, J.; Wong, Y.-C.; Khoo, K.H.; Tamada, K.; Tan, Z.-K. Thermodynamic control in the synthesis of quantum-confined blue-emitting CsPbBr3 perovskite nanostrips. J. Phys. Chem. Lett. 2020, 11, 2036–2043. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, L.; Zhang, Y.; He, W.; Peng, H.; Dai, Q. CsPbBr3 perovskite nanowires and their optical properties. Opt. Mater. 2020, 109, 110399. [Google Scholar] [CrossRef]
  105. Zhou, Z.-R.; Liao, Z.-H.; Wang, F. Shape-controlled synthesis of one-dimensional cesium lead halide perovskite nanocrystals: Methods and advances. J. Mater. Chem. C 2023, 11, 3409–3427. [Google Scholar] [CrossRef]
  106. Krajewska, C.J.; Kaplan, A.E.K.; Kick, M.; Berkinsky, D.B.; Zhu, H.; Sverko, T.; Voorhis, T.V.; Bawendi, M.G. Controlled assembly and anomalous thermal expansion of ultrathin cesium lead bromide nanoplatelets. Nano Lett. 2023, 23, 2148–2157. [Google Scholar] [CrossRef]
  107. Dang, Z.; Dhanabalan, B.; Castelli, A.; Dhall, R.; Bustillo, K.C.; Marchelli, D.; Spirito, D.; Petralanda, U.; Shamsi, J.; Manna, L.; et al. Temperature-driven transformation of CsPbBr3 nanoplatelets into mosaic nanotiles in solution through self-assembly. Nano Lett. 2020, 20, 1808–1818. [Google Scholar] [CrossRef]
  108. Montanarella, F.; Akkerman, Q.A.; Bonatz, D.; van der Sluijs, M.M.; van der Bok, J.C.; Prins, P.T.; Aebli, M.; Mews, A.; Vanmaekelbergh, D.; Kovalenko, M.V. Growth and self-assembly of CsPbBr3 nanocrystals in the TOPO/PbBr2 synthesis as seen with X-ray scattering. Nano Lett. 2023, 23, 667–676. [Google Scholar] [CrossRef]
  109. Liu, Z.; Bekenstein, Y.; Ye, X.; Nguyen, S.C.; Swabeck, J.; Zhang, D.; Lee, S.-T.; Yang, P.; Ma, W.; Alivisatos, A.P. Ligand mediated transformation of cesium lead bromide perovskite nanocrystals to lead depleted Cs4PbBr6 nanocrystals. J. Am. Chem. Soc. 2017, 139, 5309–5312. [Google Scholar] [CrossRef]
  110. Udayabhaskararao, T.; Houben, L.; Cohen, H.; Menahem, M.; Pinkas, I.; Avram, L.; Wolf, T.; Teitelboim, A.; Leskes, M.; Yaffe, O.; et al. A mechanistic study of phase transformation in perovskite nanocrystals driven by ligand Passivation. Chem. Mater. 2018, 30, 84–93. [Google Scholar] [CrossRef]
  111. Dahl, J.C.; Wang, X.; Huang, X.; Chan, E.M.; Alivisatos, A.P. Elucidating the weakly reversible Cs-Pb-Br perovskite nanocrystal reaction network with high-throughput maps and transformations. J. Am. Chem. Soc. 2020, 147, 11915. [Google Scholar] [CrossRef] [PubMed]
  112. Bao, Z.; Chiu, H.-D.; Wang, W.; Su, Q.; Yamada, T.; Chang, Y.-C.; Chen, S.; Kanemitsu, Y.; Chung, R.-J.; Liu, R.-S. Highly luminescent CsPbBr3@Cs4PbBr6 nanocrystals and their application in electroluminescent emitters. J. Phys. Chem. Lett. 2020, 11, 10196–10202. [Google Scholar] [CrossRef] [PubMed]
  113. Qin, Z.; Dai, S.; Hadjiev, V.G.; Wang, C.; Xie, L.; Ni, Y.; Wu, C.; Yang, G.; Chen, S.; Deng, L.; et al. Revealing the origin of luminescence center in 0D Cs4PbBr6 perovskite. Chem. Mater. 2019, 31, 9098–9104. [Google Scholar] [CrossRef]
  114. Ray, A.; Maggioni, D.; Baranov, D.; Dang, Z.; Prato, M.; Akkerman, Q.A.; Goldoni, L.; Caneva, E.; Manna, L.; Abdelhady, A.L. Green-emitting powders of zero-dimensional Cs4PbBr6: Delineating the intricacies of the synthesis and the origin of photoluminescence. Chem. Mater. 2019, 31, 7761–7769. [Google Scholar] [CrossRef]
  115. Jing, Q.; Xu, Y.; Su, Y.; Xing, X.; Lu, Z. A systematic study of the synthesis of cesium lead halide nanocrystals: Does Cs4PbBr6 or CsPbBr3 form? Nanoscale 2019, 11, 1784–1789. [Google Scholar] [CrossRef]
  116. Cho, J.; Banerjee, S. Ligand-directed stabilization of ternary phases: Synthetic control of structural dimensionality in solution-grown cesium lead bromide nanocrystals. Chem. Mater. 2018, 30, 6144–6155. [Google Scholar] [CrossRef]
  117. Li, Y.; Huang, H.; Xiong, Y.; Kershaw, S.V.; Rogach, A.L. Revealing the formation mechanism of CsPbBr3 perovskite nanocrystals produced via a slowed-down microwave-assisted synthesis. Angew. Chem. 2018, 57, 5833–5837. [Google Scholar] [CrossRef]
  118. Xiao, M.; Hao, M.; Lyu, M.; Moore, E.G.; Zhang, C.; Zhang, C.; Luo, B.; Hou, J.; Lipton-Duffin, J.; Wang, L. Surface ligands stabilized lead halide perovskite quantum dot photocatalyst for visible light-driven hydrogen generation. Adv. Funct. Mater. 2019, 29, 1905683. [Google Scholar] [CrossRef]
  119. Chen, L.-C.; Tien, C.-H.; Tseng, P.-W.; Tseng, Z.-L.; Huang, W.-L.; Xu, Y.-X.; Kuo, H.-C. Effect of washing solvents on the properties of air-synthesized perovskite CsPbBr3 quantum dots for quantum dot-based light-emitting devices. IEEE Access 2020, 8, 159415–159423. [Google Scholar] [CrossRef]
  120. Li, F.; Lin, F.; Huang, Y.; Cai, Z.; Qiu, L.; Zhu, Y.; Wang, Y.; Chen, X. Bromobezene aliphatic nucleophilic substitution guided controllable and reproducible synthesis of high quality cesium lead bromide perovskite nanocrystals. Inorg. Chem. Front. 2019, 6, 3577–3582. [Google Scholar] [CrossRef]
  121. Quarta, D.; Imran, M.; Capodilupo, A.-L.; Petralanda, U.; van Beek, B.; De Angelis, F.; Manna, L.; Infante, I.; De Trizio, L.; Giansante, C. Stable ligand coordination at the surface of colloidal CsPbBr3 nanocrystals. J. Phys. Chem. Lett. 2019, 10, 3715–3726. [Google Scholar] [CrossRef] [PubMed]
  122. Dong, Y.; Zhao, Y.; Zhang, S.; Dai, Y.; Liu, L.; Li, Y.; Chen, Q. Recent advances toward practical use of halide perovskite nanocrystals. J. Mater. Chem. A 2018, 6, 21729–21746. [Google Scholar] [CrossRef]
  123. Yang, D.; Cao, M.; Zhong, Q.; Li, P.; Zhang, X.; Zhang, Q. All-inorganic cesium lead halide perovskite nanocrystals: Synthesis, surface engineering and applications. J. Mater. Chem. C 2019, 7, 757–789. [Google Scholar] [CrossRef]
  124. McGrath, F.; Ghorpade, U.V.; Ryan, K.M. Synthesis and dimensional control of CsPbBr3 perovskite nanocrystals using phosphorous based ligands. J. Chem. Phys. 2020, 152, 174702. [Google Scholar] [CrossRef] [PubMed]
  125. Fiuza-Maneiro, N.; Sun, K.; Lopez-Fernandez, I.; Gomez-Grana, S.; Muller-Buschbaum, P.; Polavarapu, L. Ligand chemistry of inorganic lead halide perovskite nanocrystals. ACS Energy Lett. 2023, 8, 1152–1191. [Google Scholar] [CrossRef]
  126. Smock, S.R.; Williams, T.J.; Brutchey, R.L. Quantifying the thermodynamics of ligand binding to CsPbBr3 quantum dots. Angew. Chem. 2018, 57, 11711–11715. [Google Scholar] [CrossRef]
  127. Zaccaria, F.; Zhang, B.; Goldoni, L.; Imran, M.; Zito, J.; van Beek, B.; Lauciello, S.; De Trizio, L.; Manna, L.; Infante, I. The reactivity of CsPbBr3 nanocrystals toward acid/base ligands. ACS Nano 2022, 16, 1444–1455. [Google Scholar] [CrossRef]
  128. Akhil, S.; Dutt, V.; Mishra, N. Bromopropane as a novel bromine precursor for the completely amine free colloidal synthesis of ultrastable and highly luminescent green-emitting cesium lead bromide (CsPbBr3) perovskite nanocrystals. Nanoscale 2021, 13, 13142–13151. [Google Scholar] [CrossRef]
  129. Akhil, S.; Biswas, S.; Palabathuni, M.; Singh, R.; Mishra, N. Amine-free synthetic route: An emerging approach to making high-quality perovskite nanocrystals for futuristic applications. J. Phys. Chem. Lett. 2022, 13, 9480–9493. [Google Scholar] [CrossRef]
  130. Akhil, S.; Dutt, V.G.V.; Singh, R.; Mishra, N. Surface-state-mediated interfacial hole transfer dynamics between CsPbBr3 perovskite nanocrystals and phenothiazine redox couple. J. Phys. Chem. C. 2021, 125, 22133–22141. [Google Scholar] [CrossRef]
  131. Chen, W.-C.; Fang, Y.-H.; Chen, L.-G.; Liang, F.-C.; Yan, Z.-L.; Ebe, H.; Takahashi, Y.; Chiba, T.; Kido, J.; Kuo, C.-C. High luminescence and external quantum efficiency in perovskite quantum-dots light-emitting diodes featuring bilateral affinity to silver and short alkyl ligands. Chem. Eng. J. 2021, 414, 128866. [Google Scholar] [CrossRef]
  132. Zheng, W.; Wan, Q.; Liu, M.; Zhang, Q.; Zhang, C.; Yan, R.; Feng, X.; Kong, L.; Li, L. CsPbBr3 nanocrystal light-emitting diodes with efficiency up to 13.4% achieved by careful surface engineering and device engineering. J. Phys. Chem. C 2021, 125, 3110–3118. [Google Scholar] [CrossRef]
  133. Chiba, T.; Hoshi, K.; Pu, Y.-J.; Takeda, Y.; Hayashi, Y.; Ohisa, S.; Kawata, S.; Kido, J. High-efficiency perovskite quantum-dot light-emitting devices by effective washing process and interfacial energy level alignment. ACS Appl. Mater. Interfaces 2017, 9, 18054–18060. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, Y.; Hou, G.; Wu, Y.; Chen, M.; Dai, Y.; Liu, S.; Zhao, Q.; Lin, H.; Fang, J.; Jing, C.; et al. Surface reconstruction of CsPbBr3 nanocrystals by the ligand engineering approach for achieving high quantum yield and improved stability. Langmuir 2023, 39, 6222–6230. [Google Scholar] [CrossRef]
  135. Abiodun, S.L.; Pellechia, P.J.; Greytak, A.B. Effective purification of CsPbBr3 nanocrystals with high quantum yield and high colloidal stability via gel permeation chromatography. J. Phys. Chem. C 2021, 125, 3463–3471. [Google Scholar] [CrossRef]
  136. Abiodun, S.L.; Gee, M.Y.; Greytak, A.B. Combined NMR and isothermal titration calorimetry investigation resolves conditions for ligand exchange and phase transformation in CsPbBr3 nanocrystals. J. Phys. Chem. C 2021, 125, 17897–17905. [Google Scholar] [CrossRef]
  137. Balakrishnan, S.K.; Kamat, P.V. Ligand assisted transformation of cubic CsPbBr3 nanocrystals into two-dimensional CsPb2Br5 nanosheets. Chem. Mater. 2018, 20, 74–78. [Google Scholar] [CrossRef]
  138. Akhil, S.; Dutt, V.G.; Singh, R.; Mishra, N. Post-synthesis treatment with lead bromide for obtaining near-unity photoluminescence quantum yield and ultra-stable amine-free CsPbBr3 perovskite nanocrystals. J. Phys. Chem. C 2022, 126, 10742–10751. [Google Scholar] [CrossRef]
  139. VanOrman, Z.A.; Weiss, R.; Medina, M.; Nienhaus, L. Scratching the surface: Passivating perovskite nanocrystals for future device integration. J. Phys. Chem. Lett. 2022, 13, 982–990. [Google Scholar] [CrossRef]
  140. Krieg, F.; Ochsenbein, S.T.; Yakunin, S.; ten Brinck, S.; Aellen, P.; Suess, A.; Clerc, B.; Guggisberg, D.; Nazarenko, O.; Shynkarenko, Y.; et al. Colloidal CsPbX3 (X = Cl, Br, I) nanocrystals 2.0: Zwitterionic capping ligands for improved durability and stability. ACS Energy Lett. 2018, 3, 641–646. [Google Scholar] [CrossRef]
  141. Wang, C.; Malinoski, A.; Yuan, J.; Hu, G. A surface engineering approach for promoting dexter energy transfer from lead halide perovskite nanocrystals. J. Phys. Chem. C 2023, 127, 1135–1144. [Google Scholar] [CrossRef]
  142. Mei, X.; He, K.; Zhuang, R.; Yu, M.; Hua, Y.; Zhang, X. Stabilizing dynamic surface of highly luminescent perovskite quantum dots for light-emitting diodes. Chem. Eng. J. 2023, 453, 139909. [Google Scholar] [CrossRef]
  143. Kreig, F.; Ong, Q.K.; Burian, M.; Raino, G.; Naumenko, D.; Amenitsch, H.; Suess, A.; Grotevent, M.J.; Krumeich, F.; Bodnarchuk, M.I.; et al. Stable ultraconcentrated and ultradilute colloids of CsPbX3 (X = Cl, Br) nanocrystals using natural lecithin as a capping ligand. J. Am. Chem. Soc. 2019, 141, 19839–19849. [Google Scholar] [CrossRef] [PubMed]
  144. Grisorio, R.; Fasulo, F.; Munoz-Garcia, A.B.; Pavone, M.; Conelli, D.; Fanizza, E.; Striccoli, M.; Allegretta, I.; Terzano, R.; Margiotta, N.; et al. In situ formation of zwitterionic ligands: Changing the passivation paradigms of CsPbBr3 nanocrystals. Nano Lett. 2022, 22, 4437–4444. [Google Scholar] [CrossRef] [PubMed]
  145. Montanarella, F.; McCall, K.M.; Sakhatskyi, K.; Yakunin, S.; Trtik, P.; Bernasconi, C.; Cherniukh, I.; Mannes, D.; Bodnarchuk, M.I.; Strobl, M.; et al. Highly concentrated, zwitterionic ligand-capped Mn2+:CsPb(BrxCl1−x)3 nanocrystals as bright scintillators for fast neutron imaging. ACS Energy Lett. 2021, 6, 4365–4373. [Google Scholar] [CrossRef]
  146. Lee, A.-Y.; Kim, J.; Lee, D.; Song, M.H. Efficient and stable perovskite nanocrystal light-emitting diodes with sulfobetaine-based ligand treatment. ACS Appl. Electron. Mater. 2023, 5, 5325–5331. [Google Scholar] [CrossRef]
  147. Guo, S.; Liu, H.; He, H.; Wang, W.; Jiang, L.; Xiong, X.; Wang, L. Eco-friendly strategy to improve durability and stability of zwitterionic capping ligand colloidal CsPbBr3 nanocrystals. Langmuir 2020, 36, 6775–6781. [Google Scholar] [CrossRef]
  148. Manoli, A.; Papagiorgis, P.; Sergides, M.; Bernasconi, C.; Athanasiou, M.; Pozov, S.; Choulis, S.A.; Bodnarchuk, M.I.; Kovalenko, M.V.; Othonos, A.; et al. Surface functionalization of CsPbBr3 nanocrystals for photonic applications. ACS Appl. Nano Mater. 2021, 4, 5084–5097. [Google Scholar] [CrossRef]
  149. Shcherbakov-Wu, W.; Sercel, P.C.; Krieg, F.; Kovalenko, M.V.; Tisdale, W.A. Temperature-independent dielectric constant in CsPbBr3 nanocrystals revealed by linear absorption spectroscopy. J. Phys. Chem. Lett. 2021, 12, 8088–8095. [Google Scholar] [CrossRef]
  150. Wang, S.; Du, L.; Jin, Z.; Xin, Y.; Mattoussi, H. Enhanced stabilization and easy phase transfer of CsPbBr3 perovskite quantum dots promoted by high-affinity polyzwitterionic ligands. J. Am. Chem. Soc. 2020, 142, 12669–12680. [Google Scholar] [CrossRef]
  151. Jin, X.; Ma, K.; Gao, H. Tunable luminescence and enhanced polar solvent resistance of perovskite nanocrystals achieved by surface-initiated photopolymerization. J. Am. Chem. Soc. 2022, 144, 20411–20420. [Google Scholar] [CrossRef] [PubMed]
  152. Jin, H.; Park, G.Y.; Kim, M.K.; Cha, J.; Ham, D.S.; Kim, M. Eco-friendly solvent-processible and highly luminescent perovskite nanocrystals with polymer zwitterions for air-stable optoelectronics. Chem. Eng. J. 2023, 459, 141531. [Google Scholar] [CrossRef]
  153. Noh, S.H.; Jeong, W.; Lee, K.H.; Yang, H.S.; Suh, E.H.; Jung, J.; Park, S.C.; Lee, D.; Jung, I.H.; Jeong, Y.J.; et al. Photocrosslinkable zwitterionic ligands for perovskite nanocrystals: Self-assembly and high-resolution direct patterning. Adv. Funct. Mater. 2023, 33, 2304004. [Google Scholar] [CrossRef]
  154. Zhang, M.; Zhang, J.; Zhao, G.; Wang, G.; Liu, P.; Li, H.; Hou, X.; Qiang, P.; Yang, Y.; Su, Q.; et al. Efficient blue CsPbBr3 perovskite nanocrystals synthesis with the assistance of zwitterionic straight chain amino acids. Colloids Surf. A Physicochem. Eng. Asp. 2023, 673, 131793. [Google Scholar] [CrossRef]
  155. Cohen, T.A.; Huang, Y.; Bricker, N.A.; Juhl, C.S.; Milstein, T.J.; MacKenzie, J.D.; Luscombe, C.K.; Gamelin, D.R. Modular zwitterion-functionalized poly(isopropyl methacrylate) polymers for hosting luminescent lead halide perovskite nanocrystals. Chem. Mater. 2021, 33, 3779–3790. [Google Scholar] [CrossRef]
  156. Kirsch, C.; Naujoks, T.; Haizmann, P.; Frech, P.; Peisert, H.; Chasse, T.; Brutting, W.; Scheele, M. Zwitterionic carbazole ligands enhance the stability and performance of perovskite nanocrystals in light-emitting diode. ACS Appl. Mater. Interfaces 2023, 15, 32744–32752. [Google Scholar] [CrossRef]
  157. Zhu, H.; Kick, M.; Ginterseder, M.; Krajewska, C.J.; Sverko, T.; Li, R.; Lu, R.; Shih, M.-C.; Voorhis, T.V.; Bawendi, M.G. Synthesis of zwitterionic CsPbBr3 nanocrystals with controlled anisotropy using surface-selective ligand pairs. Adv. Mater. 2023, 35, 2304069. [Google Scholar] [CrossRef]
  158. Xie, Q.; Wu, D.; Wang, X.; Li, Y.; Fang, F.; Wang, Z.; Ma, Y.; Su, M.; Peng, S.; Liu, H.; et al. Branched capping ligands improve the stability of cesium lead halide (CsPbBr3) perovskite quantum dots. J. Mater. Chem. C 2019, 7, 11251–11257. [Google Scholar] [CrossRef]
  159. Wang, X.; Zhu, S.; Fu, J.; Li, X.; Zhao, X.; Jiang, H.; Lv, G.; Li, P.; Li, J.; Zhang, W.-H.; et al. Hybrid ligand polymerization for weakly confined lead halide perovskite quantum dots. ACS Appl. Mater. Interfaces 2023, 15, 20208–20218. [Google Scholar] [CrossRef]
  160. Hou, S.; Guo, Y.; Tang, Y.; Quan, Q. Synthesis and stabilization of colloidal perovskite nanocrystals by multidentate polymer micelles. ACS Appl. Mater. Interfaces 2017, 9, 18417–18422. [Google Scholar] [CrossRef]
  161. Sanjayan, C.G.; Jyothi, M.S.; Sakar, M.; Balakrishna, R.G. Multidentate ligand approach for conjugation of perovskite quantum dots to biomolecules. J. Colloid Interface Sci. 2021, 603, 758–770. [Google Scholar] [CrossRef] [PubMed]
  162. Chen, J.; Jia, D.; Qiu, J.; Zhuang, R.; Hua, Y.; Zhang, X. Multidentate passivation crosslinking perovskite quantum dots for efficient solar cells. Nano Energy 2022, 96, 107140. [Google Scholar] [CrossRef]
  163. Jin, X.; Ma, K.; Chakkamalayath, J.; Morsby, J.; Gao, H. In situ photocatalyzed polymerization to stabilize perovskite nanocrystals in protic solvents. ACS Energy Lett. 2022, 7, 610–616. [Google Scholar] [CrossRef]
  164. Pan, Q.; Hu, J.; Fu, J.; Lin, Y.; Zou, C.; Di, D.; Wang, Y.; Zhang, Q.; Cao, M. Ultrahigh stability of perovskite nanocrystals by using semiconducting molecular species for displays. ACS Nano 2022, 16, 12253–12261. [Google Scholar] [CrossRef]
  165. Yin, W.; Li, M.; Dong, W.; Luo, Z.; Li, Y.; Qian, J.; Zhang, J.; Zhang, W.; Zhang, Y.; Kershaw, S.V.; et al. Multidentate ligand polyethyleneimine enables bright color-saturated blue light-emitting diodes based on CsPbBr3 nanoplatelets. ACS Energy Lett. 2021, 6, 477–484. [Google Scholar] [CrossRef]
  166. Wang, S.; Du, L.; Donmez, S.; Xin, Y.; Mattoussi, H. Polysalt ligands achieve higher quantum yield and improved colloidal stability for CsPbBr3 quantum dots. Nanoscale 2021, 13, 16705–16718. [Google Scholar] [CrossRef]
  167. Wang, S.; Wang, W.; Donmez, S.; Xin, Y.; Mattoussi, H. Engineering highly fluorescent and colloidally stable blue-emitting CsPbBr3 nanoplatelets using polysalt/PbBr2 ligands. Chem. Mater. 2022, 34, 4924–4936. [Google Scholar] [CrossRef]
  168. Malinoski, A.; Hu, G.; Wang, C. Strong bidentate coordination for surface passivation and ligand-shell engineering of lead halide perovskite nanocrystals in the strongly quantum-confined regime. J. Phys. Chem. C 2021, 125, 24521–24530. [Google Scholar] [CrossRef]
  169. Yin, W.; Li, M.; Dong, W.; Zhang, X.; Zheng, W. Overcoming the ambient manufacturability-performance bottleneck in perovskite nanocrystal emitters for efficient light-emitting diodes. Angew. Chem. 2023, 62, e202303462. [Google Scholar] [CrossRef]
  170. Zhang, J.; Shen, W.; Chen, S.; Zhang, Z.; Cai, B.; Qiu, Y.; Jiang, J.; He, Y.; Nan, M.; Chen, Y.; et al. Multidentate ligand-passivated CsPbI3 perovskite nanocrystals for stable and efficient red-light-emitting diodes. J. Phys. Chem. Lett. 2023, 14, 6639–6646. [Google Scholar] [CrossRef]
  171. Zeng, Q.; Zhang, X.; Bing, Q.; Xiong, Y.; Yang, F.; Liu, H.; Liu, J.-Y.; Zhang, H.; Zheng, W.; Rogach, A.L.; et al. Surface stabilization of colloidal perovskite nanocrystals via multi-amine chelating ligands. ACS Energy Lett. 2022, 7, 1963–1970. [Google Scholar] [CrossRef]
  172. Yuan, Y.; Zhu, H.; Hills-Kimball, K.; Cai, T.; Shi, W.; Wei, Z.; Yang, H.; Candler, Y.; Wang, P.; He, J.; et al. Stereoselective C-C oxidative coupling reactions photocatalyzed by zwitterionic ligand capped CsPbBr3 perovskite quantum dots. Angew. Chem. 2020, 59, 22563–22569. [Google Scholar] [CrossRef] [PubMed]
  173. Ye, J.; Byranvand, M.M.; Martinez, C.O.; Hoye, R.L.Z.; Saliba, M.; Polavarapu, L. Defect passivation in lead-halide perovskite nanocrystals and thin films: Toward efficient LEDs and solar cells. Angew. Chem. 2021, 133, 21804–21828. [Google Scholar] [CrossRef]
  174. Liu, M.; Matuhina, A.; Zhang, H.; Vivo, P. Advances in the stability of halide perovskite nanocrystals. Materials 2019, 12, 3733. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic representation of surface termination of a CsPbBr3 NC lattice. Reprinted from ref. [4]; (b) CsPbX3 NCs with different emission colors under UV-light illumination; and (c) PL emission spectra of CsPbX3 NCs with respect to their composition. Reprinted with permission from ref. [15] Copyright@ Royal Society of Chemistry.
Figure 1. (a) Schematic representation of surface termination of a CsPbBr3 NC lattice. Reprinted from ref. [4]; (b) CsPbX3 NCs with different emission colors under UV-light illumination; and (c) PL emission spectra of CsPbX3 NCs with respect to their composition. Reprinted with permission from ref. [15] Copyright@ Royal Society of Chemistry.
Nanoenergyadv 03 00019 g001
Figure 2. Schematic representation of CsPb(Cl, Br)3 NC synthesis using OAm and OA through (a) hot–injection and (b) the LARP method. Reprinted with permission from ref. [37] Copyright@ Royal Society of Chemistry.
Figure 2. Schematic representation of CsPb(Cl, Br)3 NC synthesis using OAm and OA through (a) hot–injection and (b) the LARP method. Reprinted with permission from ref. [37] Copyright@ Royal Society of Chemistry.
Nanoenergyadv 03 00019 g002
Figure 3. (a) Schematic representation of dynamic equilibria between carboxylic acid and amines on the CsPbX3 NCs surface. (b) Schematic representation of equilibria of ligands during passivation. Reprinted with permission from ref. [60]. Copyright@ Royal Society of Chemistry.
Figure 3. (a) Schematic representation of dynamic equilibria between carboxylic acid and amines on the CsPbX3 NCs surface. (b) Schematic representation of equilibria of ligands during passivation. Reprinted with permission from ref. [60]. Copyright@ Royal Society of Chemistry.
Nanoenergyadv 03 00019 g003
Figure 4. (a) Schematic diagram of the interaction of OAmBr with CsPbBr3 NCs. Reprinted with permission from ref. [59]; (b) 1H NMR spectrum and (c) Nuclear Overhauser effect spectrum (NOESY) of CsPbBr3 NCs dispersed in deuterated chloroform; and (d) Schematic diagram of possible binding modes of OAm+ via hydrogen bond. Reprinted with permission from ref. [55] Copyright 2017 American Chemical Society.
Figure 4. (a) Schematic diagram of the interaction of OAmBr with CsPbBr3 NCs. Reprinted with permission from ref. [59]; (b) 1H NMR spectrum and (c) Nuclear Overhauser effect spectrum (NOESY) of CsPbBr3 NCs dispersed in deuterated chloroform; and (d) Schematic diagram of possible binding modes of OAm+ via hydrogen bond. Reprinted with permission from ref. [55] Copyright 2017 American Chemical Society.
Nanoenergyadv 03 00019 g004
Figure 5. Schematic representation of post-synthetic modification of CsPbBr3 NCs in the presence of OA and OAm. Reprinted with permission from ref. [77] Copyright 2022 American Chemical Society (The results in the image are based on ref. [76]).
Figure 5. Schematic representation of post-synthetic modification of CsPbBr3 NCs in the presence of OA and OAm. Reprinted with permission from ref. [77] Copyright 2022 American Chemical Society (The results in the image are based on ref. [76]).
Nanoenergyadv 03 00019 g005
Figure 6. (ac) Formation of CsPbX3 nanowires in the presence of OA/OAm with increasing amounts of HCl, HBr and HI. Reprinted with permission from ref. [63] Copyright 2017 American Chemical Society, and (d) Correlation of PL wavelength with respect to the reaction temperature under different concentrations of OA/OAm. Reprinted with permission from ref. [65]. Copyright@ Royal Society of Chemistry.
Figure 6. (ac) Formation of CsPbX3 nanowires in the presence of OA/OAm with increasing amounts of HCl, HBr and HI. Reprinted with permission from ref. [63] Copyright 2017 American Chemical Society, and (d) Correlation of PL wavelength with respect to the reaction temperature under different concentrations of OA/OAm. Reprinted with permission from ref. [65]. Copyright@ Royal Society of Chemistry.
Nanoenergyadv 03 00019 g006
Figure 7. (ad) Schematic representation of formation mechanism of one-dimensional OA/OAm-capped CsPbBr3 NC super-lattice nanostructures. Reprinted with permission from ref. [101] Copyright 2017 American Chemical Society. (e,f) Schematic representation of formation mechanism of long-range CsPbBr3 NPLs assembly in the presence of OA. Reprinted with permission from ref. [95] Copyright 2019 American Chemical Society.
Figure 7. (ad) Schematic representation of formation mechanism of one-dimensional OA/OAm-capped CsPbBr3 NC super-lattice nanostructures. Reprinted with permission from ref. [101] Copyright 2017 American Chemical Society. (e,f) Schematic representation of formation mechanism of long-range CsPbBr3 NPLs assembly in the presence of OA. Reprinted with permission from ref. [95] Copyright 2019 American Chemical Society.
Nanoenergyadv 03 00019 g007
Figure 8. (a) Ligand control of the dynamic reversibility between CsPbBr3 and Cs4PbBr6. Reprinted with permission from ref. [111] Copyright 2020 American Chemical Society. (b) Transmission electron microscopy images of cubic CsPbBr3 to Cs4PbBr6 NCs (from left to right, respectively) using the addition of an excess of OA/OAm. (c) UV-visible spectra of CsPbBr3 NCs and Cs4PbBr6 NCs under the influence of OAm/OA. (d) X-ray diffraction (XRD) of the conversion of CsPbBr3 (red) to Cs4PbBr6 NCs (blue) and back along with the intermediates (green). Bulk XRD spectra for cubic CsPbBr3 and Cs4PbBr6 are given as a bar plot. (e) Two complete CsPbBr3–Cs4PbBr6 NC cycles as followed by UV–visible spectroscopy. Plotted is the wavelength of the first excitonic absorption peak. Reprinted with permission from ref. [110] Copyright 2018 American Chemical Society.
Figure 8. (a) Ligand control of the dynamic reversibility between CsPbBr3 and Cs4PbBr6. Reprinted with permission from ref. [111] Copyright 2020 American Chemical Society. (b) Transmission electron microscopy images of cubic CsPbBr3 to Cs4PbBr6 NCs (from left to right, respectively) using the addition of an excess of OA/OAm. (c) UV-visible spectra of CsPbBr3 NCs and Cs4PbBr6 NCs under the influence of OAm/OA. (d) X-ray diffraction (XRD) of the conversion of CsPbBr3 (red) to Cs4PbBr6 NCs (blue) and back along with the intermediates (green). Bulk XRD spectra for cubic CsPbBr3 and Cs4PbBr6 are given as a bar plot. (e) Two complete CsPbBr3–Cs4PbBr6 NC cycles as followed by UV–visible spectroscopy. Plotted is the wavelength of the first excitonic absorption peak. Reprinted with permission from ref. [110] Copyright 2018 American Chemical Society.
Nanoenergyadv 03 00019 g008
Figure 9. (a) Schematic representation of dynamic binding (OA/OAm binary ligands capping) and tight binding (zwitterionic ligand capping) of Cs LHP NCs. (b) Schematic representation of properties of tight binding long-chain ligands in solution and in solid state. Reprinted with permission from ref. [139] Copyright 2022 American Chemical Society. (c,d) Chemical structure and scheme of synthesis and redispersion of 12-SBE-capped CsPbBr3 NCs. (e,f) Comparative plots of PLQY and ratio of ligand exchange with different volume additions of 12-SBE zwitterionic ligand. Reprinted with permission from ref. [146] Copyright 2023 American Chemical Society.
Figure 9. (a) Schematic representation of dynamic binding (OA/OAm binary ligands capping) and tight binding (zwitterionic ligand capping) of Cs LHP NCs. (b) Schematic representation of properties of tight binding long-chain ligands in solution and in solid state. Reprinted with permission from ref. [139] Copyright 2022 American Chemical Society. (c,d) Chemical structure and scheme of synthesis and redispersion of 12-SBE-capped CsPbBr3 NCs. (e,f) Comparative plots of PLQY and ratio of ligand exchange with different volume additions of 12-SBE zwitterionic ligand. Reprinted with permission from ref. [146] Copyright 2023 American Chemical Society.
Nanoenergyadv 03 00019 g009
Figure 10. (a) Comparison graph of surface ligand content of AHDA- and OA/OAm-capped CsPbI3 NCs with respect to purification cycles. (b) Schematic representation of AHDA and OA/OAm stabilization of CsPbI3 NCs. VA represents a ligand vacancy. (c) Diffusion coefficient values of OA/OAm- and AHDA-capped CsPbI3 NCs, determined by diffusion-ordered spectroscopy (DOSY). Reprinted with permission from ref. [172] Copyright 2022 American Chemical Society. (d) Crystal structure of strongly bound picolinate ligand and schematic representation of its passivation on the perovskite NC surface. Reprinted with permission from ref. [169] Copyright 2021 American Chemical Society.
Figure 10. (a) Comparison graph of surface ligand content of AHDA- and OA/OAm-capped CsPbI3 NCs with respect to purification cycles. (b) Schematic representation of AHDA and OA/OAm stabilization of CsPbI3 NCs. VA represents a ligand vacancy. (c) Diffusion coefficient values of OA/OAm- and AHDA-capped CsPbI3 NCs, determined by diffusion-ordered spectroscopy (DOSY). Reprinted with permission from ref. [172] Copyright 2022 American Chemical Society. (d) Crystal structure of strongly bound picolinate ligand and schematic representation of its passivation on the perovskite NC surface. Reprinted with permission from ref. [169] Copyright 2021 American Chemical Society.
Nanoenergyadv 03 00019 g010
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

Soosaimanickam, A.; Saura, A.; Farinós, N.; Abargues, R. Influence of Binary Ligands in Designing Cesium Lead Halide (CsPbX3, X = Cl, Br, I) Perovskite Nanocrystals-Oleic Acid and Oleylamine. Nanoenergy Adv. 2023, 3, 376-400. https://doi.org/10.3390/nanoenergyadv3040019

AMA Style

Soosaimanickam A, Saura A, Farinós N, Abargues R. Influence of Binary Ligands in Designing Cesium Lead Halide (CsPbX3, X = Cl, Br, I) Perovskite Nanocrystals-Oleic Acid and Oleylamine. Nanoenergy Advances. 2023; 3(4):376-400. https://doi.org/10.3390/nanoenergyadv3040019

Chicago/Turabian Style

Soosaimanickam, Ananthakumar, Alejandro Saura, Noemi Farinós, and Rafael Abargues. 2023. "Influence of Binary Ligands in Designing Cesium Lead Halide (CsPbX3, X = Cl, Br, I) Perovskite Nanocrystals-Oleic Acid and Oleylamine" Nanoenergy Advances 3, no. 4: 376-400. https://doi.org/10.3390/nanoenergyadv3040019

Article Metrics

Back to TopTop