Next Article in Journal
Effect of Film Morphology on Electrical Conductivity of PEDOT:PSS
Previous Article in Journal
In Vivo Evaluation of Bone Regenerative Capacity of the Novel Nanobiomaterial: β-Tricalcium Phosphate Polylactic Acid-co-Glycolide (β-TCP/PLLA/PGA) for Use in Maxillofacial Bone Defects
Previous Article in Special Issue
Ytterbium-Doped Lead–Halide Perovskite Nanocrystals: Synthesis, Near-Infrared Emission, and Open-Source Machine Learning Model for Prediction of Optical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalysis Based on Metal Halide Perovskites for Organic Chemical Transformations

by
Metikoti Jagadeeswararao
,
Raquel E. Galian
* and
Julia Pérez-Prieto
*
Institute of Molecular Science, University of Valencia, C/Catedrático José Beltrán 2, 46980 Paterna, Valencia, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(1), 94; https://doi.org/10.3390/nano14010094
Submission received: 21 November 2023 / Revised: 19 December 2023 / Accepted: 23 December 2023 / Published: 28 December 2023
(This article belongs to the Special Issue Design and Synthesis of New Photoactive Perovskite Nanomaterials)

Abstract

:
Heterogeneous photocatalysts incorporating metal halide perovskites (MHPs) have garnered significant attention due to their remarkable attributes: strong visible-light absorption, tuneable band energy levels, rapid charge transfer, and defect tolerance. Additionally, the promising optical and electronic properties of MHP nanocrystals can be harnessed for photocatalytic applications through controlled crystal structure engineering, involving composition tuning via metal ion and halide ion variations, dimensional tuning, and surface chemistry modifications. Combination of perovskites with other materials can improve the photoinduced charge separation and charge transfer, building heterostructures with different band alignments, such as type-II, Z-scheme, and Schottky heterojunctions, which can fine-tune redox potentials of the perovskite for photocatalytic organic reactions. This review delves into the activation of organic molecules through charge and energy transfer mechanisms. The review further investigates the impact of crystal engineering on photocatalytic activity, spanning a diverse array of organic transformations, such as C–X bond formation (X = C, N, and O), [2 + 2] and [4 + 2] cycloadditions, substrate isomerization, and asymmetric catalysis. This study provides insights to propel the advancement of metal halide perovskite-based photocatalysts, thereby fostering innovation in organic chemical transformations.

1. Introduction

Solar energy, with its abundant availability, serves as the quintessential source of renewable power, offering a clean and accessible solution to the environmental challenges posed by fossil fuels and rising pollution levels [1]. Harnessing this energy through photocatalysis, a pivotal concept in the quest for global environmental sustainability, holds remarkable potential for effortlessly generating, storing, and releasing renewable energy on demand, contributing to a cleaner ecosystem and alternative energy solutions. Photocatalysis centres around specialized materials called photocatalysts, which generate electron–hole pairs (excitons) upon absorbing photons, thereby creating free radicals that participate in chemical reactions, as shown in Figure 1a, without the need for high temperatures or harsh conditions.
There are two types of photocatalysis: homogeneous and heterogeneous photocatalysis. The first one involves soluble photocatalysts, such as polypyridyl complexes of elements, such as ruthenium and iridium, as well as organic dyes like eosin Y and 9,10-dicyanoanthracene salts. These compounds smoothly interact with reactants (substrates) in the same phase, offering a wealth of opportunities for organic synthesis. The excited states induced by these photocatalysts initiate reactions that facilitate the formation of valuable carbon–carbon and carbon–heteroatom bonds, including chiral compounds via a stereo-selective process [2,3]. Moreover, photocatalysis has simplified the synthesis process by requiring a smaller amount of catalyst (1 mole% or less), using environmentally benign reagents, and operating under mild reaction conditions. Additionally, photocatalysis enables the formation of novel compounds that would otherwise be challenging or impossible to synthesize using conventional methods; for instance, C–H bond activation in saturated alkanes [3,4]. Nonetheless, the challenge lies in the intricate separation of homogeneous photocatalysts, which obstruct their reusability. Moreover, molecular photocatalysts are costly, thus imposing limitations on large-scale production in industries. Hence, there is a pressing need to innovate and advance novel heterogeneous photocatalysts that can demonstrate outstanding activity, on par with that of the existing noble metal-based catalysts, and provide high reaction yields under open-air conditions while addressing these shortcomings.
On the contrary, heterogeneous photocatalysis employs solid materials that exist in distinct physical phases from that of the substrates. This approach holds promise for directly functionalizing organic substrates, thereby engendering the creation or transformation of valuable organic molecules. The inherent advantages of solid photocatalysts, such as recyclability and easy separation from the medium, make heterogeneous photocatalysis an appealing avenue for sustainable synthesis initiatives. Among the catalysts used in this type of photocatalysis, solution-processable nanoparticles, nanorods, and nanosheets of materials such as TiO2, g-C3N4, metal sulphides, and perovskite oxides have undergone extensive scrutiny due to their high surface-to-volume ratio, high absorption cross-section, and suitable valence and conduction band energy levels, as shown in Figure 1b [5,6]. However, these materials exhibit absorption in the UV region and slight absorption in the visible, thereby rendering them unsuitable for visible light-driven photocatalytic applications. In search of alternatives, Cd-based II-VI semiconductor nanocrystals (NCs) have been explored for photocatalytic applications in the oxidation/reduction of organic compounds [7,8,9] due to their strong absorption in the visible light spectrum, high surface-to-volume ratio, and tuneable valence band maximum (VBM) and conduction band minimum (CBM) due to the quantum confinement effect (QCE). However, the presence of intra-mid gap states stemming from the non-passivated surface states in these Cd-based NCs, as shown in the left side of Figure 1g, leads to undesirable, non-radiative recombination of excitons, thus diminishing their activity [10]. Rectifying this issue requires extensive surface passivation to get rid of the intra-mid gap states; this involves complex synthesis procedures, such as the addition of an inorganic shell that improves their stability but reduces the charge transfer and, consequently, limits their application in photocatalysis [11].
Figure 1. (a) Schematic demonstration of an organic reaction using perovskite NC under visible light illumination; (b) the energy land shape of various traditional semiconductors, lead-based and lead-free perovskites, and molecular photocatalysts; (c) AMIIX3 3D cubic perovskite; (d) A3M2IIX9 2D layered perovskite; (e) A2MIM2IIIX6 3D cubic double perovskite and (f). A4MIIM2IIIX12 layered double perovskite; top and bottom of the (cf) show the unit cells of different perovskite crystal structures and the corresponding crystal structure models viewed from the [010] zone axis; (g) electronic band structure of conventional semiconductors (left side) and perovskites (middle and right side); (h) the number of publications per year from 2017 to 2023. (cf) Reproduced with permission [12]. Copy right: 2020, American Chemical Society.
Figure 1. (a) Schematic demonstration of an organic reaction using perovskite NC under visible light illumination; (b) the energy land shape of various traditional semiconductors, lead-based and lead-free perovskites, and molecular photocatalysts; (c) AMIIX3 3D cubic perovskite; (d) A3M2IIX9 2D layered perovskite; (e) A2MIM2IIIX6 3D cubic double perovskite and (f). A4MIIM2IIIX12 layered double perovskite; top and bottom of the (cf) show the unit cells of different perovskite crystal structures and the corresponding crystal structure models viewed from the [010] zone axis; (g) electronic band structure of conventional semiconductors (left side) and perovskites (middle and right side); (h) the number of publications per year from 2017 to 2023. (cf) Reproduced with permission [12]. Copy right: 2020, American Chemical Society.
Nanomaterials 14 00094 g001
On the other hand, perovskite materials with the AMIIX3 stoichiometry, where A represents methylammonium (MA+), formamidinium (FA+), Cs+ (caesium); MII represents Pb2+, Sn2+; and X represents Cl, Br, and I, exhibit a unique three-dimensional (3D) crystal structure. In this structure, the MII-site cation is surrounded by six halide ions, forming octahedral units that share corners, thus resulting in a large void with cuboidal octahedral shapes. This void is occupied by the A-site cation. The crystal structure of perovskite (AMIIX3) is depicted in Figure 1c. By changing the composition of AMIIX3, the VBM and CBM can be adjusted, thus tuning the bandgap, as shown in Figure 1b. Furthermore, in stark contrast with conventional chalcogenide semiconductors, Pb-based perovskites exhibit a defect-tolerant electronic structure. This distinctive feature arises from the intricate interplay between a Pb lone pair s orbital and an iodine (halide) p orbital leading to anti-bonding coupling in the perovskite lattice. The outcome is the emergence of a bandgap between two antibonding bands, as illustrated on the right side of Figure 1g. Notably, this band structure directs structural defects, such as those originating from A-site cation vacancies and halide vacancies, to occupy energy levels within the VB and CB rather than lying within the bandgap [13]. The defect tolerance observed in lead halide perovskites is also, in part, ascribed to the ionicity of bonding [14]. Additionally, these materials exhibit outstanding characteristics, including excellent charge carrier transport [15,16], 3D connectivity, low excitonic binding energy, and simple solution processibility, making them ideal for various optoelectronic applications [15,17,18,19,20,21,22]. These materials have also shown promise in photocatalytic H2 generation [23], CO2 reduction [24], and organic transformations, including oxidation, C–C, C–N, C–O coupling reactions, and the formation of chiral organic compounds [25]. It is worth noting that perovskite-based photocatalysts also produced novel organic compounds, which would not have been possible by means of molecular photocatalysts [26]. This highlights the interest of the promising applications of MHP for photocatalytic organic transformations. However, concerns about the thermal and moisture stability of Pb-based perovskites have encouraged exploration of alternative materials.
Substituting Pb2+ with bi-valent metal ions, such as Sn2+ and Ge2+, is challenging due to their easy oxidation into Sn4+ and Ge4+, respectively. Incorporating metal ions with the ns2 electronic configuration, such as Bi3+ and Sb3+, results in an A3M2IIIX9 layered structure, as shown in the Figure 1d, in which the charge imbalance is addressed but at the expense of the disruption of the 3D perovskite structure.
To maintain both the structure and charge neutrality, A2MIM2IIIX6 double perovskites have been developed, as shown in the Figure 1e; they exhibit improved thermal stability but suffer from wide bandgaps and reduced photocatalytic performance. A recent breakthrough by Solis-Ibarra et al. introduced layered double perovskites (LDPs) composed of environmentally friendly elements exemplified by Cs4CuSb2Cl12 [27]. This innovative approach involves replacing MI with MII in 3D double perovskites, leading to the creation of LDPs with the stoichiometry A4MIIM2IIIX12, thereby generating a new formula (e.g., A = Cs+ and Rb+; MII = Ti2+, V2+, Cr2+, Mn2+, Fe2+, Co2+, Ni2+, Cu2+, Zn2+, and Cd2+; MIII = Sb3+ and Bi3+). These LDPs feature repeating layers of [MIIICl6]3−-[MIICl6]4−-[MIIICl6]3−-vacancy, as shown in Figure 1f. Therefore, Cs4CuSb2Cl12, Rb4CuSb2Cl12, and Cs4ZnSb2Cl12 LDPs exhibit an optimal direct bandgap (1–1.8 eV) that holds promise for photocatalytic applications, coupled with impressive stability against humidity, light, and heat [28,29].
The existing body of literature has mainly focused on reviewing the photocatalytic activity of H2 generation and CO2 reduction as well as dye degradation. Comparatively, less attention has been devoted to exploring organic transformations [23,30,31,32,33,34,35]. In addressing this gap, this review aims to focus on recent developments, underscoring a growing interest that is indicated by the rising number of publications on perovskite-based organic chemical transformations from 2017 onwards, as illustrated in Figure 1h. Specifically, the review delves into organic transformations utilizing both MHP bulk materials and NCs. Additionally, it extends its scope to heteronanocrystals resulting from combinations with other structures, such as TiO2, g-C3N4, CdS, Mxene, among others. By scrutinizing these particular areas, the review endeavours to offer valuable insights into the application of perovskite materials in catalysing organic transformations. The temporal focus from 2017 to 2023 enables an exploration of recent advancements, contributing to the fundamental understanding and unlocking the potential of perovskite-based photocatalysis in the realm of organic transformations.
Bulk perovskite materials have a smaller surface area than perovskite NCs (PNCs), and consequently, they exhibit lower photocatalytic activity, as demonstrated by Wu et al. For example, the yield of disulphide product formation in a thiophenol coupling reaction was <8% when using CsPbBr3 bulk powder, but it increased up to 98% when using the NCs [36]. The high surface-to-volume ratio of the NCs provides a large active surface area, thus increasing the chances of an interaction between the photocatalyst and reactants. Moreover, PNC composition and size, and therefore the conduction and valence band edges of perovskite NCs, can be readily adjusted [37], thereby enabling favourable redox processes. In addition, the capping of PNCs with organic ligand(s) allow for their dispersion in organic solvents, thus expanding their potential applications. Furthermore, one notable advantage of PNCs is their ease of separation from organic mixtures by centrifugation, which facilitates the recovery and purification of the PNCs. Notably, perovskite NCs can be easily synthesized at room temperature under ambient conditions [38,39] using simple synthetic procedures, such as ligand-assisted re-precipitation (LARP) and emulsion methods, to yield a large number of NCs and different morphologies.

2. Factors Effecting the Photocatalytic Performance

In the context of organic synthesis, the manipulation of the reaction condition is pivotal for achieving the desired product with an optimal yield. To unlock the full potential of PNCs for their photocatalytic activity, it is imperative to comprehensively assess various aspects, including the impact of PNC size and bandgap tuning on its photocatalytic performance, interaction with solvents, and considerations related to acidity, ions, and moisture.

2.1. Effect of Size of NCs on Photocatalytic Activity

The use of semiconductors as photocatalysts requires them to exhibit a high absorption coefficient and ensure the photogenerated charge carriers reach the surface promptly with a minimum recombination. In addition, a suitable band alignment is crucial to enable the efficient charge transfer for oxidation and reduction processes. The use of small NCs is pivotal to achieving these objectives; their diminutive size facilities the swift transport of the photogenerated charge carriers to the surface. Semiconductor NCs exhibiting QCE are particularly promising candidates for photocatalysis.

2.1.1. Size Dependent Photocatalytic Activity of Lead Halide PNCs

It is widely recognized that strong QCE occurs when the NC size is smaller than the Bohr radius of the material. The Bulk Bohr exciton radius has been calculated using the equation a b = ε r m 0 / µ a 0 , where a b is the exciton Bohr radius, ε r is the relative dielectric constant of the bulk semiconductor, m 0 is the mass of a free electron, µ is the reduced mass of the exciton, and a 0 is the Bohr exciton radius of the hydrogen atom (0.529 Å). The calculated Bohr radius for CsPbBr3 is 3.5 nm. The calculated Bohr exciton radius for CsPbBr3 is 3.5 nm [37,40]. Figure 2 shows the dependence of the UV–visible and photoluminescence (PL) spectra of different-sized CsPbBr3 NCs intended for Csp3-Csp3 and Csp3-Nsp3 bond formation and their respective yields.
CsPbBr3 NCs, under quantum confinement, exhibit size-dependent properties. Zhu et al. conducted a study on the catalytic performance of organic compounds using different sizes of CsPbBr3 NCs. They prepared octylamine-capped NCs with different edge lengths, which exhibited bulk-like NCs from 2 to 100 nm (P1) and quantum-confined NCs, such as oleylamine-capped NCs with sizes of 14, 9, 6, and 4 nm (P2, P3, P4, and P5, respectively). Figure 2a shows the size-dependent absorption and PL of the different-sized CsPbBr3 NCs and their use as photocatalysts for Csp3-Csp3 bond formation via C–H bond activation (see reaction 1 and 2 in Figure 1b).
The results showed that smaller PNCs exhibited a higher initial activity for C–H bond activation, leading to the formation of product 1a in Figure 1b. Interestingly, P2–P4 NCs achieved product yields of 54–64% within 40 min, but the prolonged reaction time did not increase the yield significantly. On the other hand, P1 initially showed a slower production of 1a in Figure 1b but improved the yield up to 85% with a prolonged reaction time. This disparity was attributed to the small NCs, whose higher surface area-to-volume ratio enabled a swifter reaction rate during the initial stages. Moreover, it was observed that moisture in the solvent interacted more substantially with the smaller NCs as opposed to the larger ones, this resulting in progressive perovskite degradation and, hence, activity decrease. In addition, P5 exhibited a yield of 8%, primarily owing to its limited absorption of visible light and its increased susceptibility to degradation. In contrast, PNCs were stable when the reactions were performed under a pre-dried, non-halide solvent as ethyl acetate. Product 2a in Figure 2b was produced in an 87% and 86% yield with P1 and P2 after 6 and 2 h of reaction, respectively. This underscores the pivotal role of smaller-sized NCs in achieving high yields in shorter time frames.

2.1.2. Size Dependent Photocatalytic Activity of Lead-Free Halide PNCs

To the best of our knowledge, the systematic study of quantum-confined, size-dependent photochemical activity has not yet been explored using A3M2IIIX9 and A2MIM2IIIX6 type NCs [41]. Moreover, A4MIIM2IIIX12 type NCs, specifically Cs4ZnSb2Cl12 LDP NCs, showed size-dependent optical properties. The calculated Bohr diameter of Cs4ZnSb2Cl12 was 5.5 nm or above (due to µ changes with different crystal facets). Caruso et al. reported different-sized organic capped Cs4ZnSb2Cl12 NCs, namely 6.2 nm, 8.3 nm, 11.0 nm, and 15.6 nm NCs in the form of spherical dots [29]. Figure 3a displays the size-dependent absorption features of Cs4ZnSb2Cl12 NCs and, consequently, the bandgap tuning with the QCE. The authors conducted calculations to determine the CBM and VBM for bulk and various-sized Cs4ZnSb2Cl12 NCs, as presented in Figure 3b. Importantly, QCE was observed to fine-tune the edges of the conduction and valence bands. Notably, the shift in the VBM proved to be more significant than that of the CBM, thus providing an appropriate band alignment conducive to driving the oxidation of toluene. As shown in Figure 3c, smaller-sized NCs exhibited a higher photocatalytic performance than bulk Cs4ZnSb2Cl12. Specifically, the NCs with sizes of 6.2 nm and 8.3 nm exhibited superior performance when compared to larger-sized NCs. Notably, the 6.2 nm NCs displayed the highest activity during the initial 6 h, but their activity declined over time when compared to that of the 8.3 nm NCs. This reduction in activity can be attributed to the gradual degradation of the 6.2 nm NCs when exposed to water released during the prolonged benzaldehyde formation process.
In general, both Pb and Pb-free PNC-based photocatalytic reactions have shown that smaller-sized NCs with suitable light absorption characteristics exhibit high photocatalytic activity. However, this increased activity often increases at the cost of compromising their stability over extended periods. Therefore, it is vital to enhance the stability of smaller NCs for performing chemical reactions under ambient/water conditions. One potential approach involves the application of a thin SiO2 layer to shield the perovskite NCs, as suggested in previous studies [42,43]. Another strategy entails the incorporation of these smaller PNCs into the pores of stable materials, such as melamine foam, to address the stability concerns [44]. Further investigation is warranted to comprehensively assess the photocatalytic performance and stability of these composite materials across a range of chemical reactions.

2.2. Stability and Reaction Condition Tolerance

In the context of organic synthesis using photocatalysis, the evaluation of perovskite materials becomes crucial, considering factors such as the co-catalyst, atmosphere, and type of solvent.

2.2.1. Role of a Co-Catalyst

CsPbBr3 NCs have been used in different reaction mediums to investigate the impact of ion effects and the role of solvents in perovskite photocatalysis. Reaction 1a in Figure 2b involves the use of (ClCH2CH2Cl)2NH2Cl as a co-catalyst with CsPbBr3 NCs photocatalyst, resulting in the formation of CsPbBrxCl3−x NCs. However, it is worth noting that the co-formation of excess Br ion during reaction 1a in Figure 2b could potentially lead to the subsequent Cl ion exchange with CsPbBrxCl3−x, thus resulting in the formation of the more stable CsPbBr3 perovskite. Conversely, reaction 1b in Figure 2b, which used Cl-substrates, demonstrated the complete transformation of CsPbBr3 NCs to CsPbCl3 NCs after the reaction. Furthermore, when non-halide solvents were employed in reaction 2a in Figure 2b, the overall stability of the PNCs (P1 and P4) was enhanced, thus improving the perovskite recyclability.

2.2.2. Role of Acidity

The role of acidity in perovskite photocatalysis is significant. When various organic acids, such as propionic acid, benzoic acid, or trifluoroacetic acid (TFA), are used as co-catalysts or substrates, an enhancement in PL was observed for P1 CsPbBr3 NCs up to an optimum concentration. This enhancement has been attributed to the carboxylic acid acting as a capping ligand through strong hydrogen bonding with the halide ions and to a robust interaction between the carboxylic acid and Pb atoms. Interestingly, the optimized concentration of TFA, which exhibits high PL intensity, also led to maximum product yields for specific reactions 3c and 3d in Figure 4. Therefore, non-halide organic acids not only stabilize the PNCs but also enhance the overall catalytic efficiency.

2.2.3. Role of Air

Air tolerance is an important aspect in chemical synthesis. While molecular photocatalysts typically require air-free conditions, PNCs do not. In fact, PNCs exhibit strong quenching with organic substrates when compared to oxygen (O2). In contrast, molecular photocatalysts often show significant and competitive O2 quenching, thus resulting in a poor catalytic performance. Consequently, PNCs demonstrate catalytic performance one to two orders of magnitude higher for reactions 1, 2, and 3 (in Figure 2b and Figure 4) than that of CdSe and molecular photocatalysts, making them highly efficient in air-tolerant reactions.

2.3. Role of Bandgap Tuning of Perovskite NCs on Photocatalytic Organic Reactions

Bandgap tuning plays a crucial role in selectively activating organic substrates during photocatalytic reactions. Three strategies can be used to tune the bandgap of metal halide PNCs (AMIIX3), specifically engineering the A- and MII-site cations and X-site anion. Out of these three strategies, halide exchange has been promising for tuning the bandgap [14]. Figure 5a shows the colour and PL tuning across the entire visible range of lead halide PNCs by treating the CsPbBr3 NCs with chloride and iodide precursors. Figure 5b shows the band alignment of CsPbX3 NCs (X: Cl, Br, and I). When the X changes from Cl to I from Br, different energy levels can be accessed, such as the well-known molecular catalysts. This allows for the activation of some interesting photochemical reactions, specifically for C–O bond formation. When CsPbBr3 NCs were used in combination with dtbbpyNiBr2 as a co-catalyst, only a minimal amount of product was generated, as shown in the reaction in Figure 5f. In contrast, when dtbbpyNiCl2 was present alongside CsPbBr3, a 78% yield was achieved. This indicates that the anion exchange process, wherein Br is replaced by Cl, shifts the PNC CBM to a higher energy, as shown in Figure 5b. This shift eases energetically favourable electron transfer processes during the reaction. Substantiating this mechanism, the study observed efficient triplet energy transfer from Ir(bpy)3 molecular photocatalysts to the substrates, a result achieved through the modification of substituent groups.
B-site engineering of lead halide PNCs (LHPNCs) leads to the tuning of the bandgap in a relatively lower range than halide exchange. For instance, in CsPbBr3 NCs, a partial substitution of Pb2+ with a smaller transition metal ions, such as Zn2+ and Cd2+, would increase the bandgap by ~40–60 nm, as shown in the Figure 5d, by contracting the lattice [45]. On the other hand, the introduction of lanthanide metal ions with a +3 oxidation state (e.g., Ce3+, Eu3+, Sm3+, Tb3+, Dy3+, Er3+, and Yb3+) [46,47,48] and a transitional Mn2+ ion [48,49] leads to multicolour emission peaks related to dopant and hosts emission peaks spanning from visible to the near infra-red (NIR) region by creating new luminescent centres but with slightly less or no changes in the bandgap of the host NCs [47]. In contrast, the complete replacement of Pb2+ with Sn2+ (CsSnX3) allow for the bandgap tuning of lead-free PNCs in the visible and NIR region, as shown in Figure 5e. However, the photocatalytic performance of these doped and CsSnX3 NCs for organic transformations are yet to be explored [50]. Furthermore, the halide exchange strategy can also be extended to other lead-free perovskites NCs A3M2IIIX9 and A2MIMIIIX6 to tune the bandgap [51,52].
The influence of the A-site cation on the bandgap in MHP is generally minimal, as the A-site cation does not directly participate in shaping the electronic band structure. However, slight adjustments in the bandgap can be achieved by considering the size differences of the A-site cation, which may lead to alterations in the octahedral tilting and/or lattice contraction or expansion [53,54]. Hazarika et al. demonstrated that introducing Cs into FAPbI3 NCs resulted in the fine-tuning of absorption and emission properties, shifting them from 800 nm to 650 nm [55].

3. Types of Heterojunctions for Photocatalysis

The superior photocatalytic performance of PNCs compared to bulk perovskite is evident. Nevertheless, the heightened Coulombic attraction force between opposite charge carriers within smaller NCs often results in a pronounced recombination, thereby hampering the migration of charge carriers. To ameliorate this issue and enhance the separation of electron–hole pairs, leading to an improved photocatalytic performance and increased stability of PNCs, diverse heterostructures have been innovatively engineered. These heterostructures incorporate varying band alignments, encompassing Schottky, type-II, and Z-scheme heterojunctions, thereby opening new avenues for advanced materials and enhanced photocatalysis. These heterostructures have been prepared by combining semiconductor–conductor and semiconductor (SC I)–semiconductor (SC II) with unequal band energy levels.
Various metals, such as Au, Ag, Pt, and carbon-based materials, such as graphene and reduced graphene, possess Fermi energy levels that are relatively positive in comparison to perovskite materials. When these conductive materials are in contact with the perovskite materials, electrons are transferred from the semiconductor to the metal until equilibrium is reached. This electron transfer leads to an upward shift in the energy bands of the semiconductor at the interface, thus resulting in band bending, as illustrated in Figure 6a. This type of junction is commonly referred to as a Schottky junction, and the band bending it creates gives rise to an energy barrier known as the Schottky barrier. The magnitude of the band bending and the Schottky barrier depends on the difference in work function between the metal and semiconductor [56]. It is important to note that the effect of band bending is less pronounced when dealing with semiconductor NCs in contrast to bulk semiconductors. Consequently, establishing Schottky junctions with NCs is considered a promising approach to enhance photocatalysis. Schottky heterojunctions serve a dual purpose in this context; they not only facilitate the extraction of charge carriers by promoting charge separation, but they also act as a co-catalyst, offering active catalytic sites essential for specific photocatalytic processes [57].
When two semiconductors SC I and SC II, each having differing CB and VB energies, come into contact, they can form a type-II heterojunction, as depicted in Figure 6b. This unique type-II band alignment enables the transfer of photogenerated electrons from SC I to SC II and photogenerated holes from SC II to SC I. Recent studies have extensively explored type-II band alignment through the creation of heterojunctions involving perovskites and various semiconductors, such as TiO2, graphitic carbon nitride (g-C3N4), CdS, etc. Depending on the specific characteristics of the semiconductors involved, some of these heterojunctions may even result in the formation of p-n heterojunctions. When p-type and n-type semiconductors with differing Fermi energy levels combine to form a junction, the flat bands of n-type and p-type undergo upward and downward band bending, respectively, at the interface. This thermodynamically favourable band alignment eases the separation of photogenerated electrons and holes, causing them to reside on different sides in distinct semiconductors, thus enhancing the electron and hole separation and thereby increasing photocatalytic performance.
To expedite the separation and transfer of photogenerated electrons and holes, a commonly employed approach is the utilization of a type-II band alignment. However, in type-II heterojunctions, the redox potential of the photocatalyst is often compromised due to oxidation and reduction processes occurring on the semiconductor with lower oxidation and reduction potentials, respectively [59]. In contrast, a Z-scheme heterojunction can maximize the redox potential of the heterostructures while maintaining efficient light absorption [60]. This proves particularly advantageous in cases where high reduction potential CB and oxidation potential VB are required, such as in a water splitting reaction. The band alignment in Z-scheme heterojunctions closely resembles that of type-II heterojunctions, with a key distinction being the direction of photogenerated electron transfer from SC II to SC I as illustrated in Figure 6c, d. Initially, all-solid Z-scheme heterojunctions where two different semiconductors were connected via an electron mediator, such as Ag, Au, and Pt, were reported. Under illumination, photoinduced electrons are transferred from the CB of SC II to VB of SC I through the electron mediator (Figure 6c). Consequently, the photogenerated holes and electrons in SC II and SC I are spatially separated and can be harnessed for various photocatalytic applications. However, the use of such expensive materials as electron mediators restricts their large-scale application in photocatalysis. Subsequently, direct Z-scheme heterojunctions were developed, where two different semiconductors with varying Fermi levels were in direct contact with each other. This results in charge distribution at the interface under dark conditions, causing band bending and the creation of an interfacial built-in electric field (IEF) in direct Z-scheme heterojunctions (Figure 6d). Moreover, a substantial difference in work function between SC I and SC II results in the formation of relatively large barriers that hinder electron transfer from SC I to SC II and hole transfer from SC II to SC I, defining the formation of a direct Z-scheme heterojunction. However, when the difference of work functions of SC I and SC II is smaller, a type-II band alignment can still co-exist with a Z-scheme heterojunction. Researchers have demonstrated the formation of both type-II and Z-scheme heterojunctions using the same perovskite materials. For instance, Tüysüz et al. [61] prepared a type-II heterojunction with CsPbBr3/TiO2, while Yu et al. established a Z-scheme heterojunction [62]. It appears that two different semiconductors with n-type characteristics can form a Z-scheme heterojunction.
The photocatalytic activity of MHP and their heterostructures for the most representative organic transformation are summarized in Table 1 and discussed in detail in Section 4, Section 5, Section 6 and Section 7.

4. Carbon–Carbon (C–C) Bond Formation

C–C bond formation is a foundational process in organic synthesis, crucial for creating complex molecular structures [63]. Traditional methods, including artificial photoredox catalysis, have advanced, but they often require noble metals or complex procedures [3,64]. Addressing this, NCs emerge as promising photocatalysts for C–C bond formation under visible light. Unlike conventional systems, PNCs offer simplified product separation, eliminating the need for costly noble metal catalysts. Their unique advantage lies in an efficient energy transfer to organic substrates, initiating photocatalytic reactions with long-lived molecular triplets. This introduces a tuneable and versatile photocatalyst, showcasing PNCs as a potential game-changer for sustainable and efficient C–C bond-forming reactions under mild conditions. In the following subsections, we delve into recent advancements in C–C coupling reactions utilizing PNCs as photocatalysts.

4.1. α-Alkylation of Aldehydes Using LHPNCs

Zhu et al. demonstrated the notable selectivity and high yield achieved in the α-alkylation of aldehydes. This process facilitated the formation of C–C bonds and reduction of C–Br using CsPbBr3 and MAPbBr3 nanoparticles (NPs) as photocatalysts under blue LED irradiation [65]. Impressively, they reported an exceptionally high turnover number (TON) exceeding 52,000, surpassing the efficiency of catalysts based on noble metals. In this method, octylamine-capped CsPbBr3 NPs (P1) were used. Under blue LED illumination in an open-air environment and in the presence of CsPbBr3 NPs, dicyclohexylamine as a co-catalyst, and 2,6-lutidine as a base, compounds 1 and 2 (Figure 7a) yielded compounds 3, 4, and 5 (Figure 7a). The formation of these three compounds was explained by a radical mediated mechanism involving three pathways. Upon perovskite photoexcitation, the electron is transferred from compound 1 and is reduced, and it eventually forms radical 7 (Figure 7b). In the presence of a sacrificial donor, N,N-diisopropylethylamine (DIPEA), radical 7 transforms into compound 3 (path I in Figure 7b). In the absence of a donor, radical 7 self-couples, resulting in the formation of compound 4 (path II in Figure 7b). The regeneration of the photocatalyst occurs through the oxidation of the electron donor and aldehyde in path I and path II, respectively. In path III (Figure 7b), oxidative quenching with in-situ generated enamine 8 (formed by the reaction between dicyclohexylamine and octanal with a loss of a water molecule) leads to radical cation 9, which, upon reaction with radical 7 followed by hydrolysis, generates the iminium cation 10, which releases product 5 (Path III in Figure 7b) and regenerates the co-catalyst. Additionally, radical trapping experiments were performed to confirm the formation of radical intermediates using TEMPO, as shown in Figure 7c.
To gain a deeper understanding of the reaction mechanism, Wang et al. employed transient absorption (TA) spectroscopy to investigate the charge transfer and reaction dynamics, focusing specifically on the selective formation of product 5 (Figure 7a) [66]. In this study, amine-free CsPbBr3 NCs were utilized as photocatalysts to avoid a potential charge transfer to the amine ligand. Figure 8a shows the results of the ground-state bleach dynamics for different systems: pure CsPbBr3NCs (A), CsPbBr3 NCs + 2-bromoacetophenone (B), CsPbBr3 NCs + octanal (C), CsPbBr3 NCs + octanal + dicyclohexylamine (D), and CsPbBr3 NCs + dicyclohexylamine (E). The analysis revealed that both B, D, and E (Figure 8a,b) exhibited decay times of ~50 ps, ~70 ps, and ~70 ps, respectively, followed by a slower decay pattern compared to pure CsPbBr3 NCs (A). Figure 8c provided insights into the reduction and oxidation potentials of the reactants, photocatalysts, and co-catalysts. The energy level of 2-bromoacetophenone was favourable for ultra-fast electron transfer, occurring within ~50 ps. On the other hand, dicyclohexylamine and the in situ-formed enamine facilitated the hole transfer, with a timescale of ~70 ps. However, it was determined that a hole transfer to dicyclohexylamine did not result in the desired oxidized enamine intermediate, eliminating this possibility from the mechanism. Consequently, following the ultrafast electron and hole transfer processes, charge-separated states were formed with a lifetime of approximately 0.8 µs. This 0.8 µs charge-separated state allows the photogenerated-charged radical intermediates to undergo C–C bond formation through a biradical pathway [66].

4.2. Benzyl Bromide Coupling Using LHP NCs

Pérez-Prieto and co-workers have demonstrated that the use of dodecylamine-capped CsPbBr3 NCs, combined with an electron donor, eventually results in photoredox-catalysed homo- and cross-coupling of benzyl bromides under visible light (447 nm) excitation, as shown in Figure 9a [67]. In this study, the authors found the thermodynamically unfavourable band alignment for electron transfer between benzyl bromide (−1.5 V vs. SHE) and CBM of CsPbBr3 NCs (−0.96 V vs. SHE). This, in principle, unexpected result was rationalized as due to substrate pre-concentration at the NC surface, facilitated by the van der Walls interaction between the benzyl bromide and alkyl chains of the organic ligand anchored to the NC surface, followed by an electron transfer from the NC to the benzyl bromide, thus eventually generating a high concentration of benzyl radicals close to or at the NC surface and driving the C–C coupling reaction forward.
Moreover, the hole in the valence band of CsPbBr3 refilled by an electron transfer from DIPEA facilitated by the favourable HOMO level (−5.76 V) relative to the VBM of CsPbBr3 (between −5.85 to −6.4 V), as shown in the Figure 9c. Notably, a TON of 17,500 with a >80% product yield was achieved for the C–C coupling of benzyl bromide. Additionally, the authors discovered that the product yield is influenced by substitution at the para position, with the yield being H, tBu, OMe > Cl > Br.

4.3. Stereo Selective C–C Oxidative Coupling Reaction

Chen and co-workers demonstrated the first photocatalytic stereoselective C–C oxidative coupling reaction using PNCs [68]. They used 2-(3,4-dimethoxyphenyl)-3-oxobutanenitrile (compound 1a in Figure 10a) as a model substrate and CsPbX3 (X = Br, I) NCs as photocatalyst under visible light illumination. This study revealed the critical role of surface ligands ensuring a stable reaction. In this context, CsPbX3 NCs capped with oleic acid/oleylamine (OA/OAm) exhibited impressive results, with a yield of 95% and excellent stereoselectivity of 99%. However, a significant drawback was the rapid degradation of OA/OAm-capped CsPbX3 NCs during the photocatalytic reaction. This issue was attributed to the instability of CsPbI3 NCs and possible ligand detachment during the post-reaction purification process. To address this challenge, the authors introduced surface modification to transform OA/OAm-CsPbBr3 NCs into zwitterionic (ZW) ligand-capped CsPbBr3 NCs. This innovation not only improved the recyclability of the reaction, extending its usability to at least three cycles, but also enhanced the reaction kinetics. This enhanced reactivity was explained by the reduction in the density of surface ligand coverage, decreasing from 5.4 nm2 for OA/OAm-capped CsPbBr3 NCs to 3.0/nm2 for ZW- capped CsPbBr3 NCs. This reduction facilitated increased substrate adsorption onto the NC surface and lowered the kinetic barrier for the reaction. It is worth noting that ZW ligands strongly adhered to the PNC surface, which enabled multiple washing steps during the NC purification process to remove excess ligands. This feature is advantageous for optoelectronic applications as well [69,70].
Furthermore, the scope of the reaction was expanded by introducing various substituents at different positions on the phenyl ring. They discovered that electron donating groups in the para position plus an extended conjugated π system were essential for the C–C coupling (dimerization) reaction to occur. Figure 10b illustrates the reaction mechanism, which begins with the formation of the substrate (α-aryl keto nitriles) radical due to the oxidation of photoexcited holes in the PNCs and follows with the C–C bond coupling of neighbouring radicals. Depending on the relative configuration of the two radicals on the NC surface, either dl- or meso-isomers can be formed. However, less steric hindrance between the aryl groups of the radicals in the trans-arrangement and an attractive interaction between the gauche cyano groups upon molecular structural relaxation drives the reaction kinetically and thermodynamically to the formation of the dl-isomer over the meso-isomer.

4.4. Synthesis of Gem-Dihaloenones by the Activation of C-Br Bonds of CBrX3 (X = Cl, Br)

Mal et al. employed CsPbBr3 NCs (PLQY of 93%) for the synthesis of gem-dihaloenones via the activation of C–Br bonds of CBrX3 (X = Cl, Br) as shown in Figure 11a, thus achieving a 94% yield of the enone under sunlight [71]. The same group reported chemo-divergent reactions (i.e., obtention of different products by changing the reaction conditions), such as the functionalization of N-methylalkanamides to form 6-endo-trig or 5-exo-trig mode of cyclization via C–Br bond activation of CBr4 using CsPbBr3 NCs obtained from an unprecedented bromide precursor dibromoisocyanuric acid as shown in Figure 11b. In both cases, an electron transfer takes place from the excited photocatalyst to the reactant to generate a radical anion intermediate.

4.5. Aminomethylation of Imidazo-Fused Heterocycles

Shi et al. have reported the direct aminomethylation of imidazo-fused heterocycles through a decarboxylative coupling reaction of N-phenyl glycine (2a in Figure 12a) and imidazo-fused heterocycles (1a in Figure 12a), using octylamine-capped CsPbBr3 NCs (P1) [72]. This reaction was carried out under open-air conditions at room temperature and with a high yield (>90%). The mechanism behind this reaction involves radical intermediates. Initially, an electron transfer from 2a (in Figure 12a) to the generated hole results in the formation of intermediate radical 5 (in Figure 12b), accompanied by the release of carbon dioxide and H+. Subsequently, the addition of radical 5 (in Figure 12b) to the C=C bond of 1a (in Figure 12a) leads to the formation of radical 6 (in Figure 12b), followed by oxidation to generate radical 7 (in Figure 12b). Eventually, radical 7 (in Figure 12b) undergoes deprotonation to yield the desired product 3a (in Figure 12b). Importantly, the CsPbBr3 NCs exhibited excellent recyclability, enabling the reproduction of the product with a high yield for up to five cycles.

4.6. Photothermal Suzuki Coupling CsPbBr3/Pd

In organic synthesis, the Suzuki coupling reaction is a pivotal method for C–C bond formation. Initially, thermally driven Suzuki coupling reactions utilized Pd complexes as homogeneous catalysts to perform these reactions. Over time, diverse heterogeneous catalysts have been developed to improve such coupling reaction. This was achieved through the integration of photocatalysts, combining Pd NPs with materials such as TiO2, SiC, C3N4, and WS2 [73,74,75,76,77]. In this context, Roeffaers et al. employed Pd nanocubes deposited onto CsPbBr3 NCs as an efficient photothermal catalyst for Suzuki coupling reaction of iodobenzene with phenylboronic acid as a model reaction under solar irradiation (AM 1.5 G) at 30 °C for 4 h [78]. The 3 wt% Pd/CsPbBr3 photocatalyst exhibited the highest conversion, being 11-fold higher than when using CsPbBr3 NCs; moreover, it occurred at a nearly 100% selectivity. Furthermore, the 3 wt% Pd/CsPbBr3 catalyst displayed an excellent performance for several reactions involving different aryl halides and electron-withdrawing and electron-donating substituents.
Notably, this catalytic performance has been attributed to the synergistic effects of both photoinduced and thermal contributions, with the photothermal activity of 3 wt% Pd/CsPbBr3 outperforming both the photoinduced activity of pristine CsPbBr3 and the photo-thermal activity of Pd nanocubes at 60 °C, underscoring the unique advantages of this composite. Furthermore, this study elucidated the activation energies of 40.4 and 30.1 kJ mol−1 for the thermal and photothermal catalysts, respectively. This information helps to explain the favourable reaction conditions observed in the Suzuki coupling reaction under photothermal catalysis. Stability tests confirmed the sustained activity of the 3 wt% Pd/CsPbBr3 composite for up to 24 h.
Mechanistic studies corroborated the key role of alkoxides, such as EtO-, in triggering the Suzuki coupling reaction. The plausible mechanism involves photoinduced electrons in CsPbBr3 transferring to Pd through the Schottky contact [79] as shown in the Figure 13A,B, thus resulting in an electron-rich Pd species that can activate the C–X bond of aryl halides, to produce Pd-adsorbed aryl intermediates (Figure 13C). Simultaneously, phenylboronic acids and alkoxides lead to a boron alkoxide complex, which interacts with photoinduced holes, weakening the carbon–boron bond as shown in the Figure 13C,D. As the activated phenylboronic acid complex migrates to the redox-activated aryl halide, the coupling reaction produces the corresponding biphenyl products (Figure 13A–D).

4.7. C–C Bond Cleavage via Decarboxylation and Dehydrogenation Reactions

Soo and co-workers conducted a study involving the C–C cleavage through decarboxylation and dehydrogenation reactions using stable, structurally defined 2D lead and tin iodide perovskites [80]. To enhance the stability of these iodide-based perovskites, they introduced hexadecyl ammonium (HDA) cations as the intercalating layers between the metal halide nanosheets, forming (HDA)2MI4 (M = Pb, Sn), as illustrated in Figure 14a. It is important to note that the inclusion of the hydrophobic, long HDA cation greatly improves the stability of these (HDA)2MI4 perovskites in a polar solvent, such as water. However, the stability of (HDA)2SnI4 is somewhat lower than that of (HDA)2PbI4. This disparity primarily arises from the susceptibility of Sn2+ to oxidize to Sn4+. Both 2D perovskites, (HDA)2PbI4 and (HDA)2SnI4, exhibit distinct bandgaps, specifically at 2.3 and 1.9 eV, respectively. These bandgap values make them suitable candidates for visible light-driven photochemical reactions.
The authors used these perovskites for decarboxylation and dehydrogenation of indoline-2-carboxylic acid to indoline and indole, respectively, in the presence of O2 (Figure 14b). Notably, (HDA)2PbI4 yielded higher quantities of indoline and indole compared to the lead-free (HDA)2SnI4. This difference in performance was attributed to the oxidative instability of Sn2+ and abundance of traps in the (HDA)2SnI4 material. Figure 14c provides an explanation of the proposed mechanism, suggesting that photooxidation is the rate determining step in both reactions, with the formation of a superoxide radical playing a crucial role in the dehydrogenation process.

5. Doped NCs

Doping serves as a highly effective and versatile strategy in photocatalysis, particularly notable in advancements with halide perovskite photocatalysts. The deliberate introduction of suitable transition metal dopants, through both bulk and surface doping, plays a pivotal role in optimizing photocatalytic processes, notably for CO2 reduction and organic synthetic reactions [81]. The success of this strategy lies in its ability to fine-tune material properties, enhancing stability, activity, and selectivity [81]. In the case of CsPbBr3, theoretical studies and experimental validations demonstrate that Co and Fe dopants with compatible lattice parameters significantly reduce energy barriers for key reaction steps, showcasing catalytic superiority [82,83]. This enhanced activity is attributed to the adsorption of CO2 at doping sites, optimizing the distance to the metal dopant. Experimental confirmation, such as with Fe-doped CsPbBr3, further underscores the efficacy of doping in activating photochemical reactions, promoting superior product yields compared to undoped counterparts [84].

Cu-Doped CsPbBr3 NCs for N–N Heterocyclization Reaction

Yan et al. developed a novel photocatalyst that employs transition metal Cu(I) ~ 1% doped CsPbBr3 NCs for a N–N heterocyclization reaction [85], which is a challenging reaction due to saturated sp3 C-based N–N rings, which are very labile in nature. The doping of Cu(I) was performed by replacing the Cs atoms on the surface of CsPbBr3 NCs. Thus, surface bound Cu(I) ions allow for the accumulation of holes after photoexcitation owing to the fact that their energy levels lie above the VB of CsPbBr3 of the NCs.
These Cu(I)-doped CsPbBr3 NC photocatalysts have been used to induce pyridazine formation from reactant 1a (in Figure 15a) with a 90% yield within 18 h, using the conditions mentioned in the reaction. In contrast with a single-carrier transfer in many catalytic reactions, in this reaction, photoinduced holes that were transferred to a diamine substrate form a biradical upon a consecutive inner sphere charge transfer from VB of the NC to surface bound Cu(I) orbitals. Therefore, a multi-charge transfer from the diamine substrate led to the formation of a radical intermediate, which is the key step for the intermolecular biradical formation. As a result, an oxidative intermolecular N–N coupling reaction takes place, as shown in the scheme in Figure 15b. It is worth mentioning that product 1b (in Figure 15a) cannot be formed when using a molecular photocatalyst or free Cu in a solution as a co-catalyst. Furthermore, this reaction can be applied to many other aliphatic and aromatic cyclic rings. This study highlights the potential applications of transition metal-doped PNCs in photocatalytic organic synthesis.

6. Photocatalytic Polymerization

Poly(3,4-ethylenedioxythiophene) (PEDOT), a widely used conducting polymer in optoelectronic devices, is typically synthesized through the chemical or electrochemical oxidation of 3,4-ethylenedioxythiophene. However, integrating PNCs into a PEDOT matrix using conventional methods poses challenges, risking contamination or damage. Photodeposition, leveraging halide perovskites attributes, offers a clean approach. This process, facilitated by photoinduced electron transfer, enhances optoelectronic properties by efficiently transferring carriers between perovskite QD and the formed polymer matrix.
In this context, Tüysüz et al. were the first to report the photocatalytic polymerization of 2,2’,5’,2”-ter-3,4-ethylenedioxythiophene (TerEDOT) using CsPbI3 NCs and visible light illumination [86]. The favourable band alignment between TerEDOT and CsPbI3 NCs enabled the transfer of photoexcited holes to TerEDOT, leading to the oxidation of TerEDOT and, ultimately, the formation of poly-(3,4-ethylenedioxythiophene) (PEDOT) in the presence of either benzoquinone or O2 as an electron acceptor (Figure 16a). It is noteworthy that, when this reaction is performed in the presence of O2, the cubic structure of CsPbI3 NCs transformed into an orthorhombic structure, which is unsuitable for optoelectronic applications.
Conversely, in the presence of 1,4-benzoquinone (BQ), the cubic structure of CsPbI3 NCs is maintained. Furthermore, the resulting polymer acted as a protective layer, encapsulating the CsPbI3 NCs and preserving their cubic structure (Figure 16b). This study emphasizes the utilization of narrow bandgap perovskites for photocatalytic applications. Additionally, Chen and co-workers investigated a similar reaction using CsPbBrxI3−x NCs and observed that CsPbI3 NCs outperformed CsPbBrxI3−x NCs in the photocatalytic polymerization of TerEDOT [87]. This superior performance was attributed to the lower binding energy of CsPbI3 NCs (152 meV) compared to CsPbBr1.5I1.5 NCs (472 meV) and CsPbBr3 NCs (520 meV). Furthermore, they demonstrated that post-treatment with methyl acetate of oleic acid/oleylamine-capped CsPbBrxI3−x NCs accelerates the photocatalytic polymerization of TerEDOT.

7. Coupling of Thiols to Disulfides

Disulfides play a crucial role in facilitating the folding of proteins into biologically active conformations [88]. Additionally, they find extensive applications in the fine chemical industry, serving as antioxidants, pesticides, pharmaceuticals, and vulcanization agents [89]. The conventional method for obtaining symmetric disulfides involves oxidative coupling of thiols with oxidants [90]. However, this process often presents challenges, including issues such as over-oxidation; the need for high catalyst-loading, excessive, and costly oxidants; reliance on strong acidic or basic media; and elevated temperatures. The quest for a mild, environmentally friendly, selective, and efficient thiol-coupling protocol remains a significant goal. In this regard, photocatalysis emerges as a gentler pathway to accomplish this crucial transformation [8,91].
Wu et al. demonstrated the activation of a S–H bond using CsPbX3 PNCs [36]. Figure 17 shows the scheme of the oxidative coupling of thiophenol to form the disulfide, using CsPbBr3 NCs under the conditions provided in the scheme. Notably, the choice of solvent plays a pivotal role in the reaction success. Changing from CH2Cl2 to CH2Br2/toluene/hexane significantly reduces the product yield, likely due to the conversion of the stable CsPbBr3 to CsPbCl3 (larger bandgap) and its relatively poor stability in various solvent systems. The most optimal solvent combination identified for this coupling reaction is CH2Cl2/cyclohexane and occurs with a 98% yield. In contrast, the reaction yields for CsPbI3 and CsPbCl3 are 58 and 12%, respectively. Furthermore, the versatility of this reaction extends to a wide range of substrates, including aromatic, aliphatic, and symmetric and unsymmetric compounds with electron-donating and -withdrawing groups. Remarkably, it consistently provides high yields. The authors proposed plausible mechanisms for this reaction. In this mechanism, thiol S and H coordinates with Pb and Br, respectively. Upon photoexcitation, electrons and holes transfer to H and S, respectively, eventually forming the corresponding radicals. As a result, two thiyl radicals combine to form disulfide, while two H radicals combine to generate H2.
Moreover, researchers have also demonstrated a fascinating cross-dehydrogenative coupling reaction involving tertiary amines, targeting C–H bonds adjacent to amino nitrogen atoms and phosphite esters, as illustrated in Figure 17b.
Table 1. Organic transformations photocatalyzed by MHP and their composites/heterostructures compared with conventional semiconductors. C% and S% represent conversion and selectivity, respectively.
Table 1. Organic transformations photocatalyzed by MHP and their composites/heterostructures compared with conventional semiconductors. C% and S% represent conversion and selectivity, respectively.
PhotocatalystIrradiationOrganic Transformation/
Photocatalytic Reaction
C (%)S (%)YearRef
TiO2simulated light irradiation, AM 1.5 Goxidation of benzylic alcohol15952018[92]
FAPbBr3simulated light irradiation, AM 1.5 Goxidation of benzylic alcohol15992018[92]
nano-FAPbBr3simulated light irradiation, AM 1.5 Goxidation of benzylic alcohol11992018[92]
15% FAPbBr3/TiO2simulated light irradiation, AM 1.5 Goxidation of benzylic alcohol63992018[92]
15% FAPbBr3/SiO2simulated light irradiation, AM 1.5 Goxidation of benzylic alcohol13992018[92]
15% FAPbBr3/TiO2-Msimulated light irradiation, AM 1.5 Goxidation of benzylic alcohol37992018[92]
15% FAPbBr3/TiO2λ ≥ 500 nmoxidation of benzylic alcohol13992018[92]
15% FAPbBr3/TiO2without light irradiationoxidation of benzylic alcohol002018[92]
CdS NCs2.6 mW 405 nm laseroxidation of benzylic alcohol-992019[93]
(Au@CdS)/Ni300 W Xe lamp
(λ > 420 nm)
oxidation of benzylic alcohol-992022[94]
Au2S/CdS6 W blue LED (445 nm)oxidation of benzylic alcohol991002022[95]
g-C3N4 oxidation of benzylic alcohol 99 [96]
5% NiOx/FAPbBr3/TiO2150 W Xe lamp, AM 1.5 G simulated light irradiationOxidation of cyclohexane0.016>992019[97]
5% NiOx/FAPbBr3/TiO2150 W Xe lamp, AM 1.5 G simulated light irradiationOxidation of cyclooctane0.032>992019[97]
CsPbI3Vis LED, 420–700 nmthiophenol coupled to disulfide58-2018[36]
CsPbBr3Vis LED, 420–700 nmthiophenol coupled to disulfide98-2018[36]
CsPbBr2ClVis LED, 420–700 nmthiophenol coupled to disulfide98-2018[36]
CsPbBr0.5Cl0.5Vis LED, 420–700 nmthiophenol coupled to disulfide68-2018[36]
CsPbBrCl2Vis LED, 420–700 nmthiophenol coupled to disulfide35-2018[36]
CsPbCl3Vis LED, 420–700 nmthiophenol coupled to disulfide12 2018[36]
CsPbCl3Vis LED, 420–700 nmthiophenol coupled to disulfide93-2018[36]
CsPbCl3 + Br2Vis LED, 420–700 nmthiophenol coupled to disulfide62-2018[36]
Cs3Bi2Br9visible light ≥ 420 nmalcoholysis of styrene oxide in isopropanol (IPA)>99-2019[98]
Cs3Bi2Br9visible light ≥ 420 nmalcoholysis of styrene oxide in IPA>99-2019[98]
CsPbBr3visible light ≥ 420 nmalcoholysis of styrene oxide in IPA1-2019[98]
CsPbI3visible light ≥ 495 nmpolymerization of 3,4-ethylenedioxythiophene32.6-2017[86]
CsPbBr3Blue LED 455 nmα-alkylation of aldehydes>99962019[65]
CsPbBr3 NCs12 W Blue LED, 455 nmsynthesis aldehyde85-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmsynthesis aldehyde52-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmsynthesis tertiary amines90-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmsynthesis tertiary amines79-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmcyclization of benzaldehyde phenylhydrazone88-2019[26]
MAPbBr312 W Blue LED, 455 nmcyclization of benzaldehyde phenylhydrazone75-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmcyclization of ethyl
(Z)-3-phenyl-3-(phenylamino)acrylate
93-2019[26]
MAPbBr312 W Blue LED, 455 nmcyclization of ethyl
(Z)-3-phenyl-3-(phenylamino)acrylate
65-2019[26]
CsPbBr3 NCs12 W Blue LED, 455 nmcoupling of benzoic acid with 4-bromotrifluorobenzene78-2019[26]
CsPbBr34.6 W Blue LEDphotopolymerized styrene12-2018[99]

8. Chiral Perovskite NCs

A material is considered chiral if its mirror image cannot be superimposed onto it. Chirality has been identified, for example, in carbohydrates and amino acids (except for glycine). Chiral catalysts have found extensive use in the asymmetric synthesis of pharmaceuticals. Recently, researchers have explored chiral MHPs to manipulate charge, light, and spin for applications in chiral optoelectronics, ferroelectrics, spin LEDs, spintronics, solar cells, and more [100]. Hybrid perovskites offer more exciting properties compared to inorganic- and organic-based perovskites, including exceptional diffusion lengths and fewer defects. Additionally, introducing chiral organic ligands into the perovskite crystal structure can induce chirality. The transfer of chirality can take place through the establishment of chemical bonds or even via spatial interactions between a chiral and an achiral system.
Reduced dimension perovskites can be represented by L2(APbBr3)n-1PbBr4 where L is surface organic chiral ligands, and A is a monovalent cation; n represents the number of inorganic layers between chiral organic ligands. Colloidal PNCs can be prepared with layers from n = 1 to ∞ by controlling the ration of L, A, and PbBr2, schematically shown in the Figure 18; L can be the R- or S-1-phenylethylammonium cation. Notably, the surface properties can be adjusted by the introduction of distinct chiral ligands.

Production of N–C Axial Chiral N-Heterocycles Using Chiral CsPbBr3 NCs

Asymmetric catalysis requires the presence of a chiral-NC surface. Earlier research suggests a specific set of chiral ligands (ammonium-based) for perovskites. These chiral ligands are introduced to the perovskite either through direct synthesis or via post-synthetic ligand exchange. In the case of direct synthesis, chiral ligands impact the crystal structure, while post-synthesis incorporation leads to chirality due to the distortion of the NC surface. Recently, Yan et al. have demonstrated asymmetric photochemical synthesis using R-PEA hybridized perovskite PEA/CsPbBr3 NCs as photocatalysts [101]. These NCs were able to oxidize N-arylamine and produce N–C axially chiral N-heterocycles (N-arylindoles) under visible light. A range of R-PEA/CsPbBr3 NCs were synthesized using emulsion synthesis, hot injection, or high-energy tip sonication methods. Among them, the tip sonication method exhibited the highest coverage of chiral ligands. The PNC surfaces were coated with both chiral ligands and non-chiral oleylammonium ligands, as depicted in Figure 18.
Typically, optically active compounds are assessed using circular dichroism (CD) spectroscopy. The chiral R- and S-forms of CsPbBr3 NCs displayed distinct CD signals at 515 nm and demonstrated anisotropy near the bandgap, as illustrated in Figure 19a. This CD response denotes the electronic interaction between the perovskite NC lattice and the chiral ligand. Chiral PEA/CsPbBr3 NCs were employed for the conversion of 1a to 2a (in Figure 19). Interestingly, a lower surface coverage with chiral ligands (9%, 17%, and 47%) yielded a higher enantiomeric excess (ee), specifically, 63%, 83%, and 99%, respectively. This suggests that a greater surface coverage enhances enantioselectivity.
Figure 19c presents the proposed mechanism for the formation of highly enantioselective N-arylindoles; the substrate 1a approaches the chiral NC surface via the weak hydrogen bonding between the N-arylindol 2a and the enantiomerically pure PEA. After light excitation, the photogenerated holes transfer to 1a, resulting in 1a+, which undergoes ring closing on the surface-bound prochiral substrate (in Figure 19c).
Density functional theory (DFT) calculations were conducted to differentiate the preferred attachment of R- and S-1a substrates to the R-PEA chiral surface of CsPbBr3 NCs. These calculations revealed a −2.4 kcal/mol preferred binding energy of prochiral N-arylamine (1a) with R-configuration compared to its enantiomeric S-configuration counterpart, binding weakly through H-bonding to the R-PEA chiral surface (in Figure 19d). This suggests that the discrimination step occurs at the beginning of the catalytic cycle. Additionally, differences in surface binding energy after a photoinduced hole transfer indicated a −5.5 kcal/mol preferred binding energy of prochiral-R-1a+ over the prochiral-S-1a+ intermediate (in Figure 19d). Interestingly, during the ring-closing step, both R- and S-forms exhibited the same transition energy. Overall, the binding preference of R-PEA chiral surface of the CsPbBr3 NCs governs the N–C axial chirality of the N-heterocycles.

9. Triplet Energy Transfer from Dyes and NCs to Substrates

Polycyclic aromatic hydrocarbons (PAHs) are organic compounds extensively explored for their singlet and triplet states in various applications, including photocatalysis, solar energy, light emission, photon up-conversion, and room temperature phosphorescence [102,103,104,105,106]. Nonetheless, the direct generation of their triplet state through optical excitation proves to be challenging due to their inherently dark nature and the inefficiency of intersystem crossing (ISC). Therefore, a sensitization approach has been adopted to generate PAH triplets, employing sensitizers, such as boron-dipyrromethane (BODIPY) dyes and fullerenes [107,108,109]. These sensitizers absorb light to populate singlet states, which then efficiently undergo ISC to populate triplets. Subsequently, these triplets can transfer their energy to PAHs (acceptor) through Dexter-type triplet energy transfer (TET), as illustrated on the left side of Figure 20. However, during the ISC process, there can be an intrinsic energy loss in the order of 0.5 eV or more, which leads to practical issues in visible-to-UV light up-conversion applications [110].
Semiconductor NCs present significant benefits compared to molecular photosensitizers due to their easily achievable preparation, robust photostability, size-dependent electronic and photophysical characteristics, high molar extinction coefficients, and simple post-synthesis functionalization [111]. Moreover, the minimal energy separation of 1 to 15 meV between bright and dark states, as opposed to the singlet–triplet splitting observed in organic molecules, renders NCs promising candidates for the sensitization process at the limit of energy conservation, driven by visible photons [111,112]. Among the diverse arrays of NCs, PNCs have excellent optical properties, such as tuneable bandgaps in the UV–visible region, near unity PL QY, and potential as a triplet photosensitizer [113]. These attributes position PNCs as promising candidates for generating triplet states in target substrates. On the right side of Figure 20, an illustration shows the sensitization of naphthalene using confined CsPbBr3 NCs, which possess a bandgap exceeding 2.6 eV, effectively aligning with the triplet energy (ET) of approximately 2.6 eV [114]. To achieve efficient TET from NCs to acceptors, the NC bandgap should be greater than the acceptor triplet energy level to be in resonance.

9.1. Demonstration of TET from Quantum Confined PNCs to Substrates

The utilization of visible light captured by long-lived triplet excited states has found widespread application in photocatalytic organic synthesis and photon up-conversion across various spectral ranges. An effective strategy involves storing the excitation energy of NCs in long-lived molecular triplet states through TET to PAHs. In this context, the combination of PNCs and PAHs allows for efficient TET, leveraging the strong light-harvesting capability of PNCs and surpassing the short emission lifetimes. This combination facilitates effective energy transfer, enabling the utilization of long-lived triplet states for desired applications in photon up-conversion and other related fields [115]. For example, in a study conducted by Wu et al., the investigation focused on the size-dependent carrier density probability at the surface of CsPbBr3 NC and its linear correlation with TET rates [116]. They employed CsPbBr3 NCs with varying edge lengths (L), ranging from 11.2 to 3.5 nm, resulting in bandgap variations from 2.43 to 2.73 eV. Triplet acceptor 1-pyrenecarboxylic acid (PCA) was introduced due to its triplet energy (ET) of 2 eV and favourable band alignment with CsPbBr3.
All CsPbBr3 NCs in the size range of 11.2 to 3.5 nm possessed bandgap energies higher than the ET of PCA, theoretically enabling TET. To explore size-dependent TET, PCA was grafted onto various-sized CsPbBr3 NCs via a carboxylic group. Figure 21a shows the absorption spectrum of the CsPbBr3-PCA complex (1:1 ratio of different-sized CsPbBr3 NCs and PCA), which was compared to that of pristine CsPbBr3 NCs, revealing a blue-shift in the lowest-energy absorption peak due to QCE as L decreased from 11.2 to 3.5 nm.
In the analysis of the CsPbBr3-PCA complex, the absorption spectrum, depicted by the coloured line in Figure 21a, reflects features from both the CsPbBr3 NC and PCA (with PCA absorbing below the 400 nm region). To delve into the mechanism of TET from CsPbBr3 NCs to PCA molecules, steady-state PL for the CsPbBr3-PCA complex was systematically examined. Figure 21b showcases the steady-state PL spectra of CsPbBr3 NCs and the CsPbBr3-PCA complex under selective excitation of NCs at 420 nm.
Notably, the PL quenching efficiency witnessed a remarkable increase from 0.6% for bulk-like NCs with an L of 11.2 nm to an impressive 99% for the more strongly confined 3.5 nm-sized NCs. The energy level diagram presented in the schematic inset of Figure 21b elucidates that this substantial PL quenching can be attributed to the efficient TET process from the CsPbBr3 NCs to the PCA molecule. This compelling finding underscores the efficacy of TET, particularly in the context of strongly confined CsPbBr3 NCs.
Transient absorption spectroscopy (TAS) was employed to monitor TET dynamics. Figure 21c show the kinetics of exciton bleaching (XB) in free NCs and NC–PCA complexes. This indicated a significant shortening of the XB lifetime in NC–PCA complexes, accompanied by the appearance of an absorption feature at ~ 423 nm corresponding to the T1 → Tn transition of PCA triplets. This provided clear evidence of ultrafast TET from NCs to PCAs.
The ultrafast quenching of NC excitons is conclusively attributed to TET from NCs to PCAs. Intriguingly, this ultrafast TET stands in stark contrast to the negligible TET observed in 9.4 nm bulklike NCs, as depicted in Figure 21d. The XB lifetimes in both free NCs and NC–PCA complexes remained similar, and no discernible triplet absorption feature was identified (Figure 21d). This striking disparity in behaviour between small- and large-sized NCs underscores the pivotal role played by the QCE in facilitating efficient TET.
Figure 21e presents the normalized kinetics of the exciton bleaching feature (XB) for various CsPbBr3-PCA complexes utilized in the study. Notably, the TET rate demonstrates a clear trend increasing with the decreasing size of NCs. Particularly, the strongly confined 3.5 nm NCs exhibit a notably higher TET rate. The QCE primarily influences the energetics (driving force and spectral overlap) and electronic coupling involved in the TET process, particularly if TET follows the well-established Dexter mechanism. To unravel the effect of the driving force and spectral overlap, the TET was examined using similar-sized NCs with varying emission peaks achieved by altering x in the CsPbBr3−xClx NCs. Intriguingly, NC–PCA complexes with different x values displayed remarkably similar XB recovery kinetics, suggesting that their TET kinetics should also align. This implies that spectral overlap and driving force play negligible roles in the TET for this system. Consequently, the inference is that quantum confinement primarily influences TET through electronic coupling.
Smaller NCs with stronger QCE are known to possess higher carrier probability densities at their surfaces, leading to stronger electronic coupling with attached acceptors. To understand the size-dependent TET rate (kTET), the carrier probability density at the CsPbBr3 NC surface denoted as |Ψs|2 was plotted against kTET, revealing an increase from nearly 0 for 11.2 nm-sized NCs to 9.1 ± 0.1 ns−1 for 3.5 nm-sized NCs. The linear fitting of this data suggests that the wave function exchange between CsPbBr3 NCs and PCA molecules is the dominant factor governing this Dexter-type mechanism. Notably, the absence of a hole transfer pathway from NCs to PCA in this study further underscores the impact of QCE on efficiently mediating TET to PCA molecules with decreasing NC size.

9.2. Isomerisation of Surface-Anchored Stilbene

Wu et al. demonstrated visible light-driven molecular isomerization and cycloaddition reactions via the TET approach from CsPbBr3 and CsPbBr3−xIx PNCs to the substrates [117]. To showcase photoisomerization, the authors selected trans-4-stilbenecarboxylic acid (trans-SCA), whose carboxylic group anchors to the of PNC surface. The stilbene triplet energy (ET) is ~2.13 eV for the trans isomer and ~2.47 eV for the cis isomer. Consequently, 3.5 nm-sized CsPbBr3 NCs with an excitonic peak at approximately 455 nm (2.73 eV) were chosen to drive the reaction via TET. Figure 22a illustrates the isomerization of trans SCA to cis SCA in hexane under 5 mW 450 nm illumination. The conversion yield (CY) was calculated based on the absorption coefficients of trans- and cis-stilbene. An illumination power of 2.7 W enabled the reaction to reach the steady state within 15 s. The light-modulated CY is depicted in Figure 22d, showing an increase in CY under illumination, ultimately reaching 90.1%. Furthermore, when photoisomerization was conducted with CsPbBr3−xIx NCs, whose bandgap slightly exceeded the ET of trans-SCA (approximately 2.25 eV), it was intriguingly transformed to cis-SCA at a slower rate than that of CsPbBr3 NCs.

9.3. Isomerization of Stilbene in Solution Using Relay

Isomerization of organic molecules in a free form, i.e., non-anchored onto the NC surface, was also possible using surface-anchored triplet transmitters, which act as a relay to accept energy from photoexcited NCs and pass it to the annihilator in solution. Wu et al. demonstrated that 9-phenanthrene carboxylic acid (PTA) with ET of ~2.64 eV as the transmitter ligands showed the TET from 3.5 nm-sized CsPbBr3 NCs to trans-stilbene as shown in Figure 22b. Figure 22e shows the absorption and PL spectra of CsPbBr3 NCs as well as those of PTA-anchored CsPbBr3 NCs. A significant PL quenching (69.2%) was observed in the presence of PTA ligands. In Figure 22f, TA kinetics show that the average lifetime (~42 µs) of CsPbBr3 NC sensitized 3PTA* shortened to ~5.8 µs for the complex of NC-PTA and trans-stilbene. This indicates an efficient TET-2 (86.1%) from 3PTA* to stilbene. This reaction was performed under 2.7 W 450 nm illumination; the CY was of 82.7% in 30 s. Thus, TET and photoisomerization can be comparable between surface-tethered and relayed systems.
Additionally, authors used the relay method to demonstrate the ring-closing isomerization of diarylethene (DAE), as shown in Figure 22c. The DAE open form can be converted into the DAE closed form via efficient TET-2 from the 3PTA* to DAE open form. This could be clearly seen since the colour changed from colourless to purple, corresponding to their absorption at of 270 nm and 565 nm of DAE open and closed forms, respectively. However, control experiments failed to convert the open-to-closed form with direct 450 nm laser illumination. Notably, the selective excitation of the DAE closed form with a 589 nm laser induced the conversion to the open form. This reversible photochromic reaction finds utility in applications such as light-induced information coding and patterning.
The relay method can also be extended to intermolecular [2 + 2] cycloaddition reactions of acenaphthylene (ACN) to dimerization. The ET of ACN was in the range of 1.87–1.95 eV, which is lower than that of PTA ligands.

9.4. Photocatalytic 2 + 2 Cycloadditions

The direct utilization of long-lived triplet states using molecular photocatalysts for organic molecule activation has been extensively explored. This often entails the TET from these molecular photocatalysts to organic substrates, inducing non-selective 2 + 2 cycloaddition reactions, as depicted in Figure 23a. In contrast, using PNC surfaces as templates, dynamically anchored substrates facilitated efficient TET to the substrate and achieved regioselective head-to-head (HH) 2 + 2 cycloaddition reactions in a syn-mode. This syn-binding mode promotes the formation of thermodynamically unfavourable 2 + 2 syn-cyclobutane products, a feat beyond the realm of molecular photocatalysis. Yan et al. demonstrated the photocatalytic synthesis of highly diastereomeric syn-selective 2 + 2 olefin cycloadditions using bulk-like PNCs as photosensitizers via TET [118]. Figure 23b exemplifies the photocatalytic 2 + 2 syn-cycloaddition reaction for substrate 2a (in Figure 23b) under optimized conditions, illustrated in the Figure 23b.
The anchoring of the substrate to the PNC surface through a carboxylic group along with its lower triplet energy than that of bulk-like PNCs mediated the TET. Comparative studies revealed that the same reaction with an Ir(ppy)3 molecular photocatalyst yields a mixture of products, with HH-anti and HH-syn as major and minor products, respectively, with trace amounts of HT-anti and HT-syn products, where the triplet energy acceptor remains in a free (unbound) state. Moreover, the applicability of photocatalytic 2 + 2 cycloadditions can be broadened to encompass both homo- and hetero-variations involving other substrates possessing triplet energies lower than those of PNCs.

10. Benzaldehyde Formation from Oxidation of Benzyl Alcohol and Toluene

The efficient and selective oxidation of saturated C–H bonds in alkanes or aromatics, such as toluene, leading to the generation of aromatic aldehydes and ketones, including benzyl alcohol, benzaldehyde, benzoic acid, and benzyl benzoate, serve as versatile intermediates for the production of pharmaceuticals, dyes, solvents, perfumes, plasticizers, dyestuffs, preservatives, and flame retardants [119]. The inherent thermodynamic strength and dynamic inertness of saturated C–H bonds necessitate harsh reaction conditions, involving high temperatures, high pressures, and the use of extra initiators or toxic reagents, in traditional methods [120,121,122]. Recognizing the environmental impact of such conventional processes, there is a growing emphasis on the development of green chemistry approaches, such as heterogeneous photocatalysis, that provide environment-benign routes for the conversion of alkanes or aromatics into value-added products.

Benzaldehyde Formation Using Pb-Based Perovskite Heterostructures as Photocatalyst

The success of LHP nanoparticles (NPs) in various optoelectronic applications has led to their utilization in photocatalytic applications. Initial studies focused on simple organic molecule transformations, such as the formation of benzaldehyde through the oxidation of benzyl alcohol and toluene, owing to the favourable band alignment of LHPs. First, we would like to discuss the formation of benzaldehyde through the oxidation of benzyl alcohol, followed by that of toluene. The energy band structure of LHPs is well suited for the oxidation of benzyl alcohol. However, studies involving LHPs, such as FAPbBr3 and CsPbBr3, for benzyl alcohol oxidation, have shown a conversion rate of ~15% with a selectivity >99% under visible light illumination. The conversion rate and selectivity can be calculated using the following formula:
Conversion   ( % ) = ( c B c A ) / c B × 100
Selectivity   ( % ) = c p / ( c B c A ) × 100 ,
where c B , c A , and c p are the concentration of reactant before and after the reaction, and the amount of target product, respectively.
Despite of having a suitable band structure, the relatively poor performance of LHPs might be attributed to the strong recombination of photogenerated carriers within the LHPs. Consequently, heterojunctions, such as type-II, direct-Z-scheme, and Schottky, were developed to significantly enhance the photocatalytic activity of LHPs. For example, Tüysüz and co-workers prepared CsPbBr3/TiO2 heterostructures by growing smaller-sized CsPbBr3 NCs (2–4 nm) onto the TiO2 NPs (P25). This resulted in a type-II band alignment at the interface of CsPbBr3 and TiO2, thus enhancing the conversion rate of benzyl alcohol up to 50% with a selectivity >99% [61]. Moreover, Roeffaers and co-workers improved the conversion rate of benzyl alcohol up to 63% (Figure 24a) without compromising the selectivity [92]. They achieved this by depositing FAPbBr3 particles onto TiO2 nanosheets, which help in the effective separation of charges due to type-II band alignment of FAPbBr3/TiO2. Additionally, TiO2 nanosheet morphology provides a large number of active catalytic sites. Figure 24b shows the PL lifetime decay profile, showing a faster decay for 15% FAPbBr3/TiO2 compared to FAPbBr3, thereby suggesting efficient charge carrier dissociation. The fitting parameters (shown in the inset of Figure 24b) indicated a decrease in the exciton PL lifetime (~43 to 24 ns) and an increase in free carrier lifetime contribution (from 266 to 459 ns) for the 15% FAPbBr3/TiO2 composite, indicating an efficient exciton dissociation at the interface of FAPbBr3/TiO2, as shown in the Figure 24c. As a result, FAPbBr3/TiO2 heterostructures exhibited a four-fold enhancement of photocatalytic conversion (63%) of benzyl alcohol with high selectivity (99%) compared to pristine FAPbBr3 and TiO2. In another study, Luo et al. used g-C3N4 as an electron extraction layer to establish a type-II band alignment with FAPbBr3 NPs. The resulting FAPbBr3/C3N4 heterostructures enhanced the conversion of benzaldehyde from 15% to ~46% with 99% selectivity [123].
Figure 24d shows the mechanism of the formation of benzaldehyde. The excitation of FAPbBr3/TiO2 heterostructures with visible light generates a photoexcited electron in the CB, leaving behind holes in the VB. The electron can then transfer to the low-lying CB of TiO2 due to favourable energy level alignment, either diffusing to the surface or being directly injected into O2, eventually leading to the formation of superoxide radicals (O2•−) in both scenarios. However, it should be noted that O2•− formation cannot be ruled out when O2 directly interacts with the surface of FAPbBr3. Simultaneously, the photogenerated hole in the VB of CsPbBr3 is refilled by an electron from the alcohol, resulting in the creation of the alcohol radical cation, which subsequently reacts with the superoxide radical to generate the corresponding aldehyde, as shown in Figure 24d.
The type-II band alignment strategy has also been studied for the toluene oxidation. Li et al. demonstrated that the introduction of Cl into CsPbBr3/TiO2 heterostructures through anion exchange resulted in the formation of asymmetric Cl distribution in CsPbBrxCl3−x [124]. As a consequence, a funnel-like band structure with a Br-rich core and Cl-rich surface was formed, enhancing the photocatalytic activity of toluene oxidation compared to that of CsPbBr3/TiO2. Remarkably, the optimized CsPbBrxCl3−x/TiO2 composite exhibited a benzaldehyde formation rate of 1874 µmol·g−1·h−1, which is approximately 1.5 times higher in performance compared to that of CsPbBr3/TiO2. In contrast, the CsPbBrxCl3−x/TiO2 prepared with direct synthesis, with uniformly distributed Cl, showed a maximum benzaldehyde conversion of 1625 µmol·g−1·h−1, which is 1.15 times lower than the anion exchange samples. The enhanced activity has been attributed to the shift in the VB with increasing Cl content in CsPbBr3/TiO2 and the unique graded band structure.
Zhu et al. used CsPbBr3/TiO2 heterostructures for toluene oxidation. Interestingly, a temperature increase enhanced the activity of toluene and benzyl alcohol oxidization. CsPbBr3/TiO2 composite showed the production rate of benzaldehyde gradually increased from 1089 µmol·g−1·h−1 at 45 °C with a selectivity of 84.2% to 2356 µmol·g−1·h−1 at 75 °C and with a selectivity of 80.3%; this value was four times higher than that of pristine CsPbBr3 NCs and three times higher than that of TiO2 NPs [125]. Moreover, the production rate of benzyl alcohol increased from 205 to 578 µmol·g−1·h−1 at 45 °C and 75 °C, respectively.
Qian et al. demonstrated a Schottky junction by decorating CsPbBr3 with graphene oxide and platinum (GO-Pt) catalyst for preventing the recombination of excitons, as depicted in Figure 25a, to facilitate benzyl alcohol conversion and H2 generation [126]. In this setup, GO plays a crucial role in structurally directing the CsPbBr3 particles, leading to a reduction in their size compared with pristine CsPbBr3 NCs. Moreover, GO-Pt acts as an electron reservoir, effectively extracting the photoinduced electrons, which can be used for H2 formation. Additionally, photoinduced holes are involved in the oxidation of benzyl alcohol to benzaldehyde, as illustrated in Figure 25b. The optimal amount of CsPbBr3/GO-Pt exhibits 1050 µmol·g−1·h−1 of benzaldehyde formation with >99% selectivity and 1060 µmol·g−1·h−1 of H2 generation.
To enhance the spatial separation of electron and hole pairs, Roeffaers et al. designed a solar photocatalyst cell with the NiOx/FAPbBr3/TiO2 configuration, as shown in Figure 26a, for photocatalytic toluene oxidation via Csp3-H bond activation [97]. In this sandwich structure, NiOx and TiO2 serve as hole and electron extracting layers, respectively. NiOx/FAPbBr3/TiO2 improved the photoinduced charge extraction compared to FAPbBr3/TiO2, as shown by the reduction in PL intensity in Figure 26c. Moreover, 2D NiOx have great charge transmission capacities, thus improving the lifetime of the free charge carrier (Figure 26d,e). Figure 26f shows the photocatalytic oxidation of the C (sp3)-H of toluene to benzaldehyde and benzyl alcohol with a conversion rate of 320 and 54 µmol·g−1·h−1 for FAPbBr3 NPs and enhanced by eight-fold upon loading 15 wt% TiO2 under saturated O2 and simulated solar light (AM1.5 G). In addition to this, the loading of 5 wt% NiOx on 15% FAPbBr3/TiO2 led to further enhancing the production rate by twelve-fold up to 3800 µmol·g−1·h−1 benzaldehyde with an 86% selectivity. A ten-fold decrease in the photocatalytic activity was observed when visible light was used for photoirradiation. This suggests that UV photons can be harvested by TiO2 to generate an exciton. NiOx/FAPbBr3/TiO2 photocatalyst also worked well for the oxidation of substituted toluene. It is noteworthy that cycloalkanes can be oxidised to their corresponding ketones with high selectivity (>99%).
Overall, the three component photocatalyst highlights the importance of designing a solar photocatalyst cell for strong absorption, charge separation, and transport. The mechanism of the formation of benzaldehyde is as follows: charge carriers are generated upon the photoexcitation of FAPbBr3, and electrons and holes are efficiently separated by TiO2 and NiOx, respectively. Moreover, molecular oxygen is reduced to a superoxide radical by interacting with electrons on TiO2, and the holes in the NiOx reacted with toluene to form a toluene radical; this is the key step for the oxidation of C(sp3)-H bonds. Subsequently, toluene radicals would react with superoxide radicals to form the respective aldehyde. The formation of benzyl alcohol can be explained by the direct reaction between toluene radicals and dissolved free molecular oxygen in toluene. Furthermore, benzyl alcohol can be oxidised to benzaldehyde.
Roeffaers and co-workers also employed the type-II heterojunction of NiOx/CsPbBr3/TiO2 in the form of planar architecture by making a thin film consisting of consecutive layers of NiOx, CsPbBr3, and TiO2 on top of tin-doped indium oxide (ITO) glass, as shown in the cross-section SEM image in Figure 27a. The main advantage of this planar architecture is that it improves the charge separation by supressing the recombination of excitons and protects the perovskite from the environment. In this type of catalyst, unlike solar cells, discontinuities in the thin film act as active sites. Thus, it is vital to establish a balance between a sufficient adsorption area and efficient charge separation. Intriguingly, the optimal NiOx/CsPbBr3/TiO2 photocatalyst exhibited 12.1 µmol·cm−2·h−1 of benzaldehyde formation from benzyl alcohol compared to CsPbBr3 films (1.1 µmol·cm−2·h−1), as shown in Figure 27b, and 8.7 µmol·cm−2·h−1 of benzaldehyde formation from toluene compared to CsPbBr3 films [127]. Moreover, the stability of the solar photocatalyst cell was evaluated using a recycling test of the cells for 23 successive cycles, where each cycle took 4 h; the results are shown in Figure 27c. Interestingly, an increase in photoactivity was observed with an increase in the reaction time for both systems. They achieved an average production rate of 16.3 µmol·cm−2·h−1 of benzaldehyde with a 99% selectivity after 90 h of testing with the NiOx/CsPbBr3/TiO2 photocatalyst, whereas CsPbBr3 films exhibit 4.8 µmol·cm−2·h−1 for 11 cycles of 4 h testing. In addition, the authors achieved an impressive TON of 45,300 for the first 23 cycles when using a NiOx/CsPbBr3/TiO2 photocatalyst. In contrast with many reports, where, beyond 2–3 cycles, a gradual decrease in the production rate of benzaldehyde was observed, in this case, the gradual increase in performance with the number of cycles was observed. This was attributed to improving the crystal symmetry and thereby increasing the mobility of charge carriers by trace amounts of water generated during the reaction. However, it has been inferred that this approach may not apply for the photocatalysts exhibiting a high conversion rate of benzaldehyde (ranging from a few hundred to several thousand µmol·g−1·h−1due to increase water generation, which can result in the perovskite deterioration.
Furthermore, this planar junction photocatalyst was also successful in oxidising other alcohols, such as 1-phenylethanol, 2-phenyl ethanol, and 4-fluorobenzylalcohol, thus confirming the general applicability of the solar photocatalyst cell.
Most recently, Yi et al. developed hierarchical CsPbBr3/TiO2 heterostructures by assembling CsPbBr3 NPs onto a nanoflower-shaped TiO2, which allowed for an intimate interfacial contact between CsPbBr3 NPs and TiO2 that greatly enhanced the toluene photocatalytic oxidation conversion [128]. This hierarchical heterojunction not only benefited from efficient charge separation due to its type-II band alignment but also offered a large surface area, allowed for increased absorption of light due to refraction and scattering, and enhanced toluene adsorption. As a result, the optimized CsPbBr3/TiO2 nanoflower-shaped photocatalyst demonstrated high performance, yielding a toluene conversion rate into benzaldehyde and benzyl alcohol of 8670 and 1530 µmol·g−1·h−1, respectively. This level of photoactivity surpasses that of pristine CsPbBr3 by 20-fold. This study highlights the critical role of TiO2 nanoflake morphology in elevating the overall performance of the photocatalyst.

11. Preparation of Singlet Oxygen Driven Oxidized Compounds Using g-C3N4/Perovskite

Singlet oxygen (1O2) is a highly reactive, short-lived species, valued for its role in facilitating [4 + 2] and [2 + 2] cycloaddition reactions as well as ene-reactions and epoxidations [129,130]. The generation of 1O2 through photochemical means is considered an environmentally friendly approach compared to chemical methods [131].
Quadrelli et al. demonstrated the photochemical generation of 1O2 using oxidized g-C3N4 and its application in cycloaddition reactions [132]. Furthermore, recognizing the synergetic potential of heterostructures comprising metal halide perovskites and 2D materials for 1O2 generation, Corti et al. explored heterostructures involving g-C3N4 in combination with PEA2MX4 (PEA = phenylethylammonium; M = Pb and Sn; X = Br and Cl) and g-C3N4 in conjunction with DMASnBr3 (DMA = dimethylammonium). These heterostructures were employed in hetero Diels–Alder (HDA) as a benchmark reaction, employing 1,3-cyclohexadiene and 1O2 under the reaction conditions detailed in Figure 28, thereby leading to the formation of 4-hydroxycyclohex-2-en-1-one (2) (in Figure 28) [133]. Using g-C3N4/PEA2MX4 photocatalysts yielded moderate to good product yields, while DMASnBr3 reached yields of up to 63%. Additionally, various other reactions, including ene reactions and oxidations, have also been investigated. Nevertheless, certain limitations were observed in acyclic alkene oxidation, and poor chemo selectivity was noted in some cases.

12. Opening of Epoxide Reactions Using Lead-Free Cs3Bi2Br9 Perovskites

Dai et al. successfully demonstrated the use of lead-free Cs3Bi2Br9 perovskite in ring-opening reactions of epoxides to produce β-alkoxy alcohols [98]. This approach offers an alternative to the conventional method for the preparation of β-alkoxy alcohols, which often involves harsh conditions with corrosive, strong acids and high energy requirements. The Cs3Bi2Br9 perovskite was prepared in situ by adding a precursor solution of CsBr and BiBr3 in DMSO directly into the reaction medium containing the epoxide and alcohol. The alcohol acted as an antisolvent and nucleophile. Microscopy and a powder X-ray diffraction analysis revealed that larger aggregates of Cs3Bi2Br9 perovskite formed in the antisolvent method. The resulting Cs3Bi2Br9 perovskite was used as a photocatalyst for the synthesis of 2-isopropoxy-2-phenylethanol from styrene oxide and isopropanol (Figure 29). Remarkably, this reaction exhibited an activity of 1333 µmol h−1 g−1, a product yield of up to 86%, and an excellent selectivity (>99%) after 6 h irradiation under visible light. This facile photocatalytic system produced only the β-alkoxy alcohol regioisomer. The specific reaction conditions can be observed in the provided scheme.
In contrast to this, when the same reaction was carried out using CsPbBr3, BiBr3, CsBr, and Bi2O3 photocatalysts, a poor photocatalytic activity was observed, specifically a conversion rate of 1%, 13%, <1%, and 9% for CsPbBr3, BiBr3, CsBr, and Bi2O3, respectively. This is indicative of active sites related to Bi atoms for epoxide alcoholysis. Other type of ring opening of epoxides showed that the conversion rates were dependent on the electron density, steric hindrance of epoxide, and polarity of the solvent. An investigation on the epoxide opening mechanism suggests that weak Lewis-acid sites existing on the Cs3Bi2Br9 surface helps the product formation, which is in line with strong acidic conditions used for the ring opening of epoxides in the case of a thermal catalytic approach. The photocatalytic mechanism is shown in the Figure 29. The photoexcitation of the perovskite generates electrons and holes in the CB and VB, respectively, and molecular oxygen forms an oxygen radical anion by interacting with the excited electrons. The photogenerated holes interacting with alcohols result in the formation of an alcohol radical, which acts as an active nucleophile, while the Bi-based Lewis sites on the photocatalyst activate the epoxide by coordinating with the oxygen atom of the three membered heterocyclic ring. Finally, the attack of the nucleophile on the activated epoxide from the rear side results in the formation of the desired product.

13. Benzaldehyde Formation by Toluene Oxidation

13.1. Lead-Free Cs3Bi2Br9 Perovskites

Dail et al. used lead-free Cs3Bi2Br9 platelets for toluene oxidation [134]. They prepared Cs3Bi2Br9 platelets via the rapid cooling of the precursor solution in dilute H2SO4 in the presence of ethyl acetoacetate (EA) under liquid nitrogen conditions (Figure 30a) [134]. The absence of EA resulted in irregularly shaped platelets with a wide size range (Figure 30a). However, the inclusion of EA at varying concentrations of halide perovskite in the dilute H2SO4 solution (ranging from 10 to 4 g/L) led to the formation of platelets with controlled dimensions. The basal plane size of the platelets decreased from 40 µm to 2–10 µm, while the thickness ranged from 500 to 100 nm. The presence of EA was found to induce the preferential growth of Cs3Bi2Br9 platelets along the (003) planes, as confirmed by powder X-ray diffraction (XRD) (Figure 30b) and a scanning electron microscopy (SEM) analysis (Figure 30c–g). This control over crystal growth was attributed to the strong interaction between the carbonyl groups of EA and Bi atoms.
Thinner Cs3Bi2Br9 platelets were employed as a photocatalyst for the C–H bond activation of toluene under open-air conditions and converted 232 µmol of toluene in 36 h of light irradiation. The photocatalytic oxidation of toluene resulted in the formation of benzaldehyde with a yield of 2.2% and a selectivity of at least 88%. As byproducts, benzaldehyde and benzoic acid were also observed. The interaction of photogenerated holes with toluene would result in the formation of benzyl radicals, followed by a reaction with superoxide radicals, which is formed via the interaction of molecular oxygen with excited electrons, such as the oxidation of toluene by the TiO2/FAPbBr3/NiOx composite.

13.2. Cs3Bi2Br9/Pd Heterostructures

Roeffaers and co-workers reported the development of Schottky junction-type heterostructures involving lead-free Cs3Bi2Br9 and Pd for benzyl alcohol oxidation and H2 generation [135,136]. They achieved this by using crystalline Cs3Bi2Br9 as the core material, complemented by amorphous Pd nanocubes forming the outer shell. Cs3Bi2Br9/Pd Schottky junction exhibited a four-fold enhancement in performance compared to pure Cs3Bi2Br9, boosting benzaldehyde formation with >99% selectivity and H2 production rates from 368 ± 30 and 341 µmol·g−1·h−1 to 1457 ± 78 and 1421 ± 91 µmol·g−1·h−1, respectively.

13.3. Cs3Bi2Br9/TiO2 Heterostructures

The lower conversion efficiency of toluene oxidation using Cs3Bi2Br9 platelets would be attributed to an electron and hole recombination before reaching the surface of the platelets, owing to their long distances. Lu and co-workers employed type-II heterojunctions of lead-free layered perovskite, i.e., Cs3Bi2Br9 and TiO2 for oxidation of benzyl alcohol to benzaldehyde [137]. Figure 31a shows the HRTEM image of the Cs3Bi2Br9/TiO2 heterostructure. The band alignment in Figure 31b shows the type-II heterojunction enables the effective separation of photogenerated electrons and supresses the electron–hole recombination. Figure 31c shows the conversion rate of benzyl alcohol to benzaldehyde, where Cs3Bi2Br9 showed a 33.2% yield, while an optimal Cs3Bi2Br9/TiO2 heterojunction shows a 73% conversion with near 100% selectivity and 1465 µmol·g−1·h−1 photocatalytic activity. Interestingly, the recycle test shows a similar conversion rate at up to five cycles. It is noted that the conversion rate of pristine lead-free Cs3Bi2Br9 perovskite shows ~5 times higher performance than LHPs. This could be attributed to the more positive VBM of lead-free Cs3Bi2Br9 perovskite than LHPs, which provides a strong driving force for the hole transfer and activates the oxidation of benzyl alcohol.

13.4. Cs3Bi2Br9/MXene Heterostructures

MXenes are derived from MAX phases with a general formula Mn+1AXn, where n = 1–3, M is a transition metal, A is mostly a group 13 or 14 element, and X is carbon and/or nitrogen; Tx represents surface terminations [138]. These two-dimensional (2D) nanomaterials have garnered significant attention due to their excellent properties, such as high metallic electrical conductivity, mobility, and excellent hydrophobicity [139]. Moreover, its layered nature and abundant surface terminations offer excellent compatibility to construct heterojunctions [140].
Yang and co-workers reported the type-II heterojunction of between Cs3Bi2Br9 NCs and Ti3C2Tx MXene for a toluene oxidation reaction [141]. Figure 32a shows the HRTEM image of Cs3Bi2Br9/Ti3C2Tx composite showing the highly crystalline nature of MXene and Cs3Bi2Br9. The bandgaps are calculated from the Tauc plots. The narrowing of the bandgap from 2.75 eV for pristine Cs3Bi2Br9 NCs to 2.47 eV for Cs3Bi2Br9/MXene-10 suggests the absorption of more visible light due to MXene’s smaller bandgap than that of Cs3Bi2Br9 (Figure 32b). Furthermore, steady-state PL shows a significant decrease in PL intensity upon increasing in the loading of MXene on Cs3Bi2Br9, indicative of the suppression of electron–hole recombination (Figure 32c). In addition, transient photocurrent spectra show the production of a high photocurrent (Figure 32d), and EIS Nyquist plots show less interfacial charge resistance (Figure 32e) for Cs3Bi2Br9/MXene-7.5 and suggests the appropriate composition for the photocatalyst. Cs3Bi2Br9/MXene-7.5 exhibits a higher photocatalytic toluene conversion than pristine Cs3Bi2Br9 NCs (2121 µmol·g−1·h−1 vs. 850 µmol·g−1·h−1) and with 100% selectivity to benzaldehyde under blue LED (5 W) illumination.
Interestingly, this performance was enhanced by almost two times under Xe lamp (300 W) without changing the selectivity. Furthermore, these catalysts can also be used for oxidizing electron-donating and -withdrawing group substituents. The impressive performance of the Cs3Bi2Br9/MXene-7.5 composite is due to the efficient separation of photoinduced electrons, allowed by the excellent electrical conductivity of the monolayer Ti3C2Tx MXene and the intimate Cs3Bi2Br9/MXene contact. Furthermore, the ultra-thin nature of 2D MXenes reduces the migration distance of charge carriers to the MXene surface, thus contributing to its enhanced capabilities.

13.5. Cs3Bi2Br9/g-C3N4 Heterostructures

Graphite-like carbon nitride, g-C3N4, is a metal-free n-type semiconductor and boasts distinct optical, electrical, structural, and physiochemical features [142]. These attributes establish g-C3N4 as an innovative material with a wide range of applications in multifunctional nanoplatforms, particularly in electronics, catalysis, and energy [143,144,145]. Its versatility is evident in its ability to create tailored hybrid photocatalysts with diverse compositions and structures, capturing interest for their effectiveness in solar radiation absorption and charge carrier separation, making g-C3N4 a key player in advancing photocatalysis.
Bai et al. designed a type-II heterojunction employing nanosheets of Cs3Bi2Br9 and g-C3N4. This heterojunction between nanosheets of amorphous g-C3N4 and crystalline Cs3Bi2Br9 allows for a large area interface sharing between Cs3Bi2Br9 and g-C3N4. Consequently, the lifetime of photogenerated charge carriers in Cs3Bi2Br9 perovskite will be enhanced by suppressing the electron–hole recombination [146]. The band offset between Cs3Bi2Br9 and g-C3N4 facilitates the transfer of photoinduced holes from Cs3Bi2Br9 to g-C3N4 and electrons from g-C3N4 to Cs3Bi2Br9 in the type-II heterojunction of Cs3Bi2Br9/g-C3N4.
The photocatalytic properties of the prepared catalysts were assessed for the selective oxidation of toluene under visible light illumination. The optimized photocatalyst 10% Cs3Bi2Br9/g-C3N4 composite demonstrated the highest benzaldehyde formation rate of 4.53 mmol·h−1·g−1, which is 41.8 and 2.3 times higher than g-C3N4 and Cs3Bi2Br9 nanosheets, respectively. Moreover, the selectivity of this composite was approximately 90%.

13.6. Cs3Bi2Br9/CdS Heterostructures

Zou et al. prepared type-II heterojunction of Cs3Bi2Br9/CdS by the in situ growth of Cs3Bi2Br9 NPs on the preformed CdS nanorods as shown in Figure 33a [147]. The electronic band alignment of the Cs3Bi2Br9/CdS composite shows a type-II band structure (Figure 33b), allowing for photoinduced holes to accumulate in the VB of Cs3Bi2Br9 and electrons in the CB of CdS lead to the efficient photooxidation of toluene. Figure 33c shows the conversion rate of toluene for the pristine Cs3Bi2Br9 and Cs3Bi2Br9/CdS composites. As expected, the composite exhibits an excellent performance of 6.79 mmol·g−1·h−1, which is higher than 37.6- and 7.1-fold CdS and Cs3Bi2Br9, respectively. Furthermore, DFT studies suggests that the interaction between C/O atoms of intermediates (benzyl radical) and Cs+ of Cs3Bi2Br9 lead to C–C or O–C bonds. This indicates that Cs+ also play a key role in the surface activation of aromatic C–H bonds.

13.7. Cs3Sb2Br9/g-C3N4 Heterostructures

Kuang et al. demonstrated the type-II heterojunction of Cs3Sb2Br9/C3N4 by in situ growth of crystalline Cs3Sb2Br9 NPs (2–6 nm) on the g-C3N4 amorphous nanosheets, as shown in Figure 34a [148]. Figure 34b shows the absorption spectrum of Cs3Sb2Br9/C3N4 composite and their individual components, suggesting more visible light absorption of Cs3Sb2Br9/C3N4 composite due to Cs3Sb2Br9. Electronic band structure calculation shows that the interface of Cs3Sb2Br9/g-C3N4 composite is a type-II heterojunction as shown Figure 34c. Type-II band structure allows for the accumulation of photoinduced holes in the CB band of g-C3N4, which activate toluene to generate benzyl radical intermediates, while photoinduced electrons move to the CB of Cs3Sb2Br9 and reduce O2 to O2. This efficient charge transfer due to the strong absorption of light results in high photocatalytic activity, specifically of 8346.8 µmol·g−1·h−1 for toluene oxidation, i.e., 26.6- and 6.0-fold higher than g-C3N4 and Cs3Bi2Br9, respectively (Figure 34d).

13.8. Z-Scheme Heterojunction Formation by Cosharing Atoms

Teng et al. demonstrated the 2D/2D heterojunction of CoxBi2−xO2CO3/Cs3Bi2Br9 (x = ~1) in the form of nanosheets [149]. The heterojunction was prepared by in situ acid etching strategy by taking CoxBi2−xO2CO3 as the precursor and fundamental constituent of Bi3+. The Bi atoms of CoxBi2−xO2CO3 nanosheet serve as the nucleation sites to grow the Cs3Bi2Br9 perovskite, as shown in Figure 35a. The TEM image in Figure 35b shows the ridge-like tiny nanosheets with a thickness of ~2 nm that emerged on the surface of Co0.97Bi1.03O2CO3; labels 1 and 2 in Figure 35c correspond to the Cs3Bi2Br9 and Co0.97Bi1.03O2CO3, respectively. It was found that the (112) plane of Cs3Bi2Br9 with a Bi-Bi distance of 14.2 Å is similar to the distance between five Bi atoms in (002) face of Co0.97Bi1.03O2CO3 with a small lattice mismatch (<0.75%). This led to the epitaxial growth of Cs3Bi2Br9 in the (001) face direction, forming an angle of 51.06° with the layer of Co0.97Bi1.03O2CO3, as shown in Figure 35d.
Experimental investigations and DFT analyses indicate the establishment of a staggered band structure, where the Fermi level of Co0.97Bi1.03O2CO3 was observed to be higher than that of Cs3Bi2Br9. Fermi level alignment occurs when these two materials come into close contact, thus leading to a redistribution of charges. This phenomenon induces band bending and an interfacial built-in electric field (IEF) in the direction of Co0.97Bi1.03O2CO3 to Cs3Bi2Br9 (Figure 36a). Furthermore, a Bader charge as high as 0.977 e was observed and corresponded to the electron transfer from Co0.97Bi1.03O2CO3 to Cs3Bi2Br9, indicative of a strong interfacial electronic coupling interaction, which will contribute to a large IEF. This kind of heterointerface with a large IEF can guide the photoinduced electrons and holes transporting via the Z-scheme route rather than type-II heterojunction. To investigate the type of heterojunction of between CoxBi2−xO2CO3 and Cs3Bi2Br9, control studies were conducted with DMPO to capture ·OH species. The CBM of both Co0.97Bi1.03O2CO3 and Cs3Bi2Br9 can drive the production of O2. In addition, the VBM of Co0.97Bi1.03O2CO3 is too negative to form OH radicals. Interestingly, the Co0.97Bi1.03O2CO3/Cs3Bi2Br9 composite showed strong peaks of OH radicals, whereas a physical mixture of Co0.97Bi1.03O2CO3 and Cs3Bi2Br9 failed to show OH radical formation. This suggests the Co0.97Bi1.03O2CO3/Cs3Bi2Br9 composite exhibits a Z-scheme-type heterojunction (Figure 36a), and their physical mixture shows type-II band alignment (Figure 36b).
To check the charge separation efficiency, the photocatalytic activity of toluene oxidation under visible light illumination was chosen because the VBM of Co0.97Bi1.03O2CO3 is not positive enough to form the benzyl radical from toluene. Notably, Co0.97Bi1.03O2CO3 and the physical mixture of Co0.97Bi1.03O2CO3 and Cs3Bi2Br9 show no activity. The Co0.97Bi1.03O2CO3/Cs3Bi2Br9 composite and Cs3Bi2Br9 showed an activity of benzaldehyde formation of 1837 and ~883 µmol·g−1·h−1. A physical mixture shows no activity due to photoinduced holes of Cs3Bi2Br9 that accumulate in the VB of Co0.97Bi1.03O2CO3, which cannot drive the toluene oxidation due to an unfavourable band alignment. These results further highlight the pivotal role of high-quality interfaces with an atomic-level close contact of Co0.97Bi1.03O2CO3/Cs3Bi2Br9 heterojunction.

13.9. Z-Scheme Heterojunction of Cs3Bi2Br9/d-BiOBr

Bai et al. reported the construction of a Z-scheme heterojunction of Cs3Bi2Br9/d-BiOBr (d-defective) for the photocatalytic toluene oxidation [150]. Bi atoms in d-BiOBr serve as a source of Bi and nucleation sites for Cs3Bi2Br9 perovskite growth. Figure 37a,b show the SEM, TEM images of d-BiOBr nanosheets with a thickness of ~4.5 nm, and Figure 37c shows the HRTEM image of the Cs3Bi2Br9/d-BiOBr composite, indicating the epitaxial growth of Cs3Bi2Br9 nanodots with a particle size of ~12.1 nm on d-BiOBr ultrathin nanosheets. To know the interaction between Cs3Bi2Br9 and d-BiOBr, Raman measurements were conducted on the Cs3Bi2Br9, d-BiOBr, and Cs3Bi2Br9/d-BiOBr composite. The vibrational modes of BiBr2 (186.52 cm−1) and BiBr3 (161.62 cm−1) for the Cs3Bi2Br9/d-BiOBr composite are shifted towards a lower wave number, indicating the interaction between Br- of Cs3Bi2Br9 and [Bi2O2]2+ of d-BiOBr (Figure 37d). As a result, the Bi–Br bond stretched the [Bi2Br9]3− cluster, which was evident from the increase of Bi–Br bond distance from 0.276 nm to 0.313 nm in the DFT calculations. Thereby, Bi–Br bond could function as an interfacial bridge between Cs3Bi2Br9 and d-BiOBr.
Band structure of d-BiOBr, Cs3Bi2Br9, Cs3Bi2Br9/d-BiOBr are shown in Figure 37e. The Fermi level (Ef) of Cs3Bi2Br9, which is higher than that of d-BiOBr, leads to the diffusion of electrons from Cs3Bi2Br9 to d-BiOBr to establish a new equilibrium state. This results in generating a built-in electric field in the direction of Cs3Bi2Br9 to d-BiOBr at the interface. Further, steady-state surface photovoltage measurements showed a higher photovoltage for the Cs3Bi2Br9/d-BiOBr composite than the other two compositions at 300–520 nm. This implies that there is a migration of photoinduced holes in d-BiOBr to the illuminated side and migration of photoexcited electrons of Cs3Bi2Br9 to the non-illuminated side (Figure 37f). The remaining photogenerated electrons of d-BiOBr, recombining with the photoinduced holes of Cs3Bi2Br9 via Bi–Br interfacial bridge, led to efficient electron hole separation via Z-scheme route.
The photocatalytic performance of the catalysts was evaluated on the selective toluene oxidation under visible light illumination. The Cs3Bi2Br9(1.7)/d-BiOBr composite exhibited a conversion rate of 7.24 mmol·g−1·h−1, which is 24.1 and 6.8 times higher than that of d-BiOBr and Cs3Bi2Br9, respectively, and with a selectivity of 80% for benzaldehyde (Figure 37g). Furthermore, the Cs3Bi2Br9(1.7)/d-BiOBr composite showed high stability and repeatability of performance up to 5 cycles. In support with photocatalytic activity, the DFT calculation also suggests a lower Ead and activation energy (−1.03 and 0.85 eV, respectively) for the Cs3Bi2Br9/d-BiOBr composite with two holes than d-BiOBr with one hole (−0.62 and 1.79). These results suggest that photogenerated holes transferred to the surface are higher in the Cs3Bi2Br9(1.7)/d-BiOBr composite than in d-BiOBr, thus demonstrating the pivotal role of a Z-scheme heterojunction in the photocatalytic activity.
Similarly, Roeffaers et al. constructed Z-scheme heterojunctions using a FAPbBr3/Bi2WO6 composite [151]. The VB position of Bi2WO6 is suitable for benzyl alcohol oxidation, whereas the CB position of FAPbBr3 is for CO2 reduction. FAPbBr3/Bi2WO6, 12.5%, exhibited high benzaldehyde formation (250 µmol·g−1·h−1), more than 16 and 4 times that of FAPbBr3 and BiWO6 under AM 1.5 G simulate solar light irradiation. In addition, the Z-scheme heterojunction approach was also used for the photooxidation of benzyl alcohol [152] and CO2 reduction [153,154].

13.10. Sb-Doped Cs3Bi2Br9 Perovskite

To enhance the absorption of visible light and activate catalytic sites of the Cs3Bi2Br9 nanosheets, Bai et al. doped the nanosheets with Sb for toluene oxidation [155]. They prepared various compositions of Cs3Bi2−xSbxBr9 (x = 0, 0.2, 0.5, 1.3, and 2) nanosheets and observed that the bandgaps of x = 0.2, 0.5, 1.3 in Cs3Bi2−xSbxBr9 were 2.40, 2.32, and 2.24 eV, respectively, which are smaller than those of Cs3Bi2Br9 (2.63 eV) and Cs3Sb2Br9 (2.45 eV). This phenomenon was attributed to the defect-related absorption resulting from the energy mismatch between Bi and Sb atomic orbitals in the Cs3Bi2−xSbxBr9 perovskite [156].
SEM images for different compositions of Cs3Bi2−xSbxBr9 revealed that Cs3Bi2Br9 exists in the form of nanosheets, and the morphology changed from nanosheets to NPs by increasing the amount of Sb (Figure 38a–e). Specifically for x = 0.2 and 0.5, a sheet-like structure was maintained, while for x = 1.3, NP formation was detected, resembling Cs3Sb2Br9.
Transient photocurrent measurements indicated that Cs3Bi1.8Sb0.2Br9 exhibited the highest photocurrent intensity, surpassing Cs3Bi1.5Sb0.5Br9, Cs3Bi2Br9, Cs3Bi0.7Sb1.3Br9, and Cs3Sb2Br9. This finding suggested that Cs3Bi1.8Sb0.2Br9 was the best composition in terms of charge carrier generation (Figure 38f). Furthermore, this behaviour was reflected in the photocatalytic activity of toluene conversion, with Cs3Bi1.8Sb0.2Br9 nanosheets displaying the highest toluene conversion rate of 5830 µmol h−1 g−1, which is 3.3 times higher than that of the NPs (Figure 38g) indicates the vital role of the high surface area with many active sites.
DFT studies were conducted to determine the effects of Sb doping on hole generation and, subsequently, the photocatalytic activity; DFT studies were conducted. A charge difference analysis for different compositions showed that Cs12Bi7Sb1Br36 had electrons of the layer 0.2 e lower than that of the other layer, indicating hole accumulation on the Sb side (Figure 38h). This suggested that when Sb doping was small, most of the Sb was located on one side, promoting hole generation. Consequently, Cs3Bi1.8Sb0.2Br9 nanosheets exhibited high activity due to extended absorption with efficient charge generation and separation [157].
Similar to this study, Cui et al. conducted a study on the impact of Sb-doping in Cs3BiBr9 for the photocatalytic oxidation of thioanisole, revealing that the strategic introduction of Sb in place of Bi resulted in a reduced bandgap and decreased Bi-related defects near the CBM, leading to enhanced carrier mobility [158]. Specifically, the composition Cs3Bi2−xSbxBr9 (CBSB; x = 0.3) displayed outstanding performance in the photocatalytic oxidation of thioanisole to methyl phenyl sulfoxide in hexane. Notably, CBSB-0.3 demonstrated exceptional efficiency with a 95% conversion rate and 99% selectivity and maintained stability up to five cycles, outperforming all other compositions. In contrast, analogous studies using CsPbBr3 perovskite failed to demonstrate any noticeable activity, highlighting the promising potential of CBSB-0.3 perovskites in facilitating this oxidation reaction. Furthermore, control experiments established that the primary active species responsible for the formation of methyl phenyl sulfoxide were photogenerated holes and O2.

13.11. Effect of A-Site Cation in A3Sb2Br9

To study the influence of A-site cation in A3Sb2Br9 on the photocatalytic organic transformation, Zhang et al. used A3Sb2Br9 (A = MA, Cs) NPs for the visible light-driven photocatalysis of toluene oxidation via C–H bond activation [159]. Interestingly, in the CsxMA3−xSb2Br9 structure, Cs3Sb2Br9 NPs exhibited 5.3 times higher activity than MA3Sb2Br9 despite a slight narrowing of the bandgap (approximately 0.13 eV). The maximum photocatalytic activity was observed for x = 2.3 (Figure 39a). However, this significant enhancement in activity cannot be solely explained by the slight narrowing of the bandgap (Figure 39b). Instead, it is attributed to the raising of the VB, which reduces the oxidation ability of toluene.
XPS analysis was performed to reveal the mechanism of A-site on the photocatalytic activity. It was found that, with increasing Cs in CsxMA3−xSb2Br9, the binding energy of Sb 4d peaks shift to lower energy (0.25 to 0.31 eV), indicating that Sb became electron richer and, subsequently, Br became slightly electron deficient. The octahedron distortion values based on the c/a values suggests that the octahedron distortion increases with MA content in CsxMA3−xSb2Br9 (Figure 39c). Consequently, there was an increase in the electron cloud density of Br sites on the surface increases, which weakened the C–H activation process (Figure 39c). This octahedron distortion is caused by the Jahn–Teller effect via an A-site cation, which influences the crystal field splitting energy of [SbBr6] octahedra. In addition, Cs3Sb2Br9 NPs showed good stability for up to four cycles and a slight decrease in the photocatalytic activity from 2 to 1.62 mmol·h−1·g−1. The findings clearly indicated that the photocatalytic activity strongly correlated with the octahedral distortion of [SbBr6] affected by A-site cations.

13.12. Cs3Bi2Br9/SiO2 Heterostructures

Dai et al. introduced a novel approach to enhance charge separation efficiency and active catalytic centres by integrating small Cs3Bi2Br9 nanoparticles (2–5 nm) into ordered mesoporous SBA-15 silica for photocatalytic C–H bond activation [160]. These nanoparticles are confined within the pore channels of SBA-15, as evidenced by the high-angle annular dark-field mode-scanning electron microscopy (HAADF-STEM) image (in Figure 40a), and they provide a significant number of catalytically active centres. The PL spectrum reveals a decrease in PL intensity for Cs3Bi2Br9-loaded SBA-15 compared to bulk Cs3Bi2Br9, indicating the suppression of radiative recombination of electron–hole pairs (Figure 40b). Notably, the compositions with 5 and 10 wt% Cs3Bi2Br9-loaded SBM-15 exhibited lower PL intensity, indicating an enhanced charge separation within these configurations. The Cs3Bi2Br9/SBA-15 composite was employed for the C–H bond activation of toluene in the presence of air and under visible light irradiation. Comparing different compositions, the 10 wt%-loaded SBA-15 demonstrated the highest conversion rate of 12,600 µmol g−1·h−1, with a high selectivity of 90%, as shown in the Figure 40c. Moreover, the extension of the photocatalytic study to other aliphatic and aromatic hydrocarbons results in an excellent performance. Notably, the highest conversion rate of 32,900 µmol g−1·h−1 was reported for the transformation of ethyl benzene to acetophenone.
DFT studies conducted to gain more insight into the interactions between the catalyst and toluene revealed that the interaction of benzene rings with Bi or Cs occurs through dispersion and charge transfer (π system to Bi σ* orbitals). Additionally, the hydrogen atoms of the -CH3 groups in toluene are located near the surface Br atoms of the catalyst, as shown in the inset of Figure 40c. Furthermore, the VBM of Cs3Bi2Br9 is mainly dominated by Br 4p orbitals, while Bi p-orbitals contribute to the CBM. As a result, the C–H bond activation induced by photogenerated holes on the bromine atoms results in efficient toluene oxidation.

13.13. DP NCs

Ag incorporation into 2D-layered Cs3Bi2Br9 (CBB) perovskite led to the formation of 3D Cs2AgBiBr6 (CABB) double perovskite, thus eliminating the strong localisation of electron–hole pairs [161]. Consequently, CABB exhibited better properties than CBB, such as a low exciton binding energy, effective carrier mass, higher carrier mobility, long carrier lifetimes. In addition, the bandgap becomes narrow (from 2.7 eV of CBB to 2.1 eV of CABB) due to the participation of Ag 4d orbitals in the VB. To further narrow down the bandgap of CABB, Bi3+ was replaced with Tl3+, Sb3+, In3+, and Fe3+ in Cs2AgBiBr6 [162,163,164]. However, Sb-substitution is more suitable for photocatalytic applications; the introduction of 37.5% Sb in CABB narrows the bandgap up to 1.86 eV owing to the rise in VB by participating Sb 5s orbitals above the Bi 6s orbitals [163]. Moreover, introducing Sb3+ in Bi-based perovskites prolongs the charge carrier lifetimes. To have strong light absorption, along with excellent carrier separation, Li et al. used Sb-substituted Cs2AgBiBr6/g-C3N4 (CASBB/CN) composite for photocatalytic C–H bond activation [165]. The CASBB/CN composite was made by mechanical grinding powders of ASBB and using different ratios of g-C3N4. Figure 41a shows the HRTEM image of the 20% CASBB/CN composite, where a clear boundary can be seen between crystalline CASBB perovskite and amorphous g-C3N4. The origin of the chemical connectivity has been attributed to a strong interaction between NHx groups of g-C3N4 and bromide ions.
Figure 41b shows the absorption spectrum, where the g-C3N4 shows an absorption edge of 450 nm (2.7 eV). Notably, the absorption of visible light enhanced by increasing the loading amount of Sb in the CASBB/CN composite was compared with CABB. Furthermore, the band alignment of the CASBB/CN composite shows a type-II heterojunction (in Figure 41c), which facilitates excellent charge separation. The photocatalytic activity of toluene oxidation is shown in Figure 41d. Among all the compositions, CASBB/CN composites showed the highest activity of toluene conversion, followed by CASBB, CABB, and g-C3N4 with high selectivity 96%. The 20% CASBB/CN composite showed the highest activity of benzaldehyde, with a production rate of 1088 µmol·g−1·h−1. The high activity is due to extended wavelengths of light absorption and the improved charge separation. Figure 41e shows the schematic mechanism of benzaldehyde formation.
In a similar strategy, Song et al. also demonstrated the formation of heterostructures of CABB with an ultra-thin carbon nitride (UCNT) sheet for photocatalytic oxidation of toluene [166]. The in-situ growth of Cs2AgBiBr6 on a UCNT sheet led to strong chemical interaction between perovskite and UCNT. Moreover, the type-II band alignment of CABB/UCNT facilitates the efficient charge separation. In the case of the CABB/UCNT composite, the photogenerated electrons transfer to the CB of CABB, whereas photogenerated holes transfer to the VB of the UCNT, as shown in Figure 41f; this behaviour is opposite to the transfer of photogenerated carriers in the CASBB/CN composite. As shown in the scheme, the reduction of O2 takes place on the surface of perovskite, and toluene becomes oxidized on the surface of the UCNT. Consequently, the reaction between the benzyl radical and superoxide radical leads to the formation of benzaldehyde and benzyl alcohol. Thus, the best conversion rate of toluene, 2630 µmol·g−1·h−1, has been achieved using the CABB-80/UCNT composite.

13.14. A4MIIM2IIIX12 LDP NCs

Solis-Ibarra and co-workers conducted a study revealing that Cs4CuSb2Cl12 LDP exhibits promising optical properties for solar and photocatalytic applications. Notably, Cs4CuSb2Cl12 LDPs displayed a significantly higher electrical conductivity, approximately one order of magnitude greater than that of MAPbI3 perovskite. This suggests that these lead-free materials hold great potential for future photocatalytic applications.
In a more recent study, Caruso and co-workers explored the tunability of size and morphology in Cs4ZnSb2Cl12 LDP NCs and their impact on toluene photooxidation reactions. Figure 42 illustrates the conversion rate of toluene into benzaldehyde, with benzyl alcohol and benzoic acid as the byproducts, using various Cs4CuSb2Cl12 LDP NC morphologies. Notably, 9.7 nm nanoplates exhibited a remarkable performance compared with other photocatalysts given in the Table 2, achieving a high benzaldehyde formation rate of up to 1893 µmol·g−1·h−1 with an impressive selectivity of 95%. This achievement is particularly remarkable given that the NCs were capped with long chain ligands, specifically oleylamine/oleic acid ligands. Furthermore, the concept of creating heterostructures by combining A4MIIM2IIIX12 LDPs with 2D materials holds the promise of significantly enhancing photoactivity beyond what A4MIIM2IIIX12-type NCs can achieve on their own. This opens exciting possibilities for advancing photocatalytic applications.

14. Oxidation of 5-Hydroxymethylfurfural to 2,5-Diformylfuran

Biomass has garnered significant attention as a vital renewable carbon feedstock, serving as a promising alternative to petroleum resources to produce sustainable fuels and high-value fine chemicals. Within the realm of biomass-derived compounds, 5-hydroxymethylfurfural (HMF) possesses the remarkable ability to undergo selective transformations, leading to the creation of a variety of multifunctional furanic products. One of the most noteworthy outcomes is 2,5-diformylfuran (DFF), which has found widespread application in both the chemical and pharmaceutical industries.
The traditional method of oxidation for this transformation relies on copious stoichiometric oxidants, resulting in substantial waste generation and environmental pollution. However, for economic and sustainable development considerations, there is a strong preference for a greener approach. The use of green photocatalytic oxidation, employing molecular oxygen to convert HMF into DFF, has emerged as an attractive synthetic pathway.
Lv and co-workers employed bulk MAPbBr3 perovskite-based photocatalysts for the highly selective oxidation of HMF to DFF [170].
This photocatalytic process was conducted in an acetonitrile solvent using an O2 atmosphere and 450 nm LED light irradiation. This resulted in the complete conversion of HMF into DFF, with an impressive selectivity of over 90% and a yield exceeding 96%. In Figure 43a, the mechanism behind the formation of DFF from HMF is depicted, wherein photogenerated electrons are transferred to O2, forming ·O2 and singlet oxygen (1O2). These intermediates then react with HMF to yield the final product, DFF, along with the byproduct, H2O2. It is worth noting that this reaction proved unsuccessful when using MAPbCl3 and MAPbI3 photocatalysts. In the case of MAPbCl3, the self-decomposition of HMF occurred under 365 nm LED light illumination, rendering the reaction unfeasible. Similarly, MAPbI3 decomposed under 450 nm LED illumination.
Wang and co-workers carried out a similar reaction by employing a Schottky heterojunction consisting of lead-free Cs2SnBr6 (bandgap 2.86 eV) and reduced graphene oxide (rGO), as shown in Figure 43b [171], using a 300 W Xe lamp with a 400 nm cut-off filter. The outcome was a selective conversion of HMF to DFF (>99.5%), while a selectivity of 88% for DFF was obtained when using Cs2SnBr6/rGO heterostructures (Figure 43c). Additionally, the Cs2SnBr6/rGO combination produced an overoxidation product of DFF, namely 5-formyl-2-furancarboxylic acid (FFCA). It is important to note that Cs2SnBr6 differs from typical perovskite AMIIX3 materials, forming scattered octahedra [SnI6]2− in a vacancy-ordered double perovskite structure. While this structure has inherent limitations in charge transport, the creation of a Schottky heterojunction significantly enhances the charge extraction efficiency, leading to a notable improvement in the performance of Cs2SnBr6/rGO heterojunctions compared to pristine Cs2SnBr6.

15. Csp3-Csp3 Coupling Reaction Using Lead-Free Cs3Sb2Br9 NCs

Perez-Prieto et al. reported a significant enhancement in the performance of the C–C coupling reaction when using a Cs3Sb2Br9 NCs photocatalyst for benzyl bromide and its substituted aryl bromides [172]. The TON reached approximately 105,000. This high TON was attributed to the favourable alignment of CBM of the Cs3Sb2Br9 NCs relative to the unoccupied molecular orbital (LUMO) of benzyl bromide, which eased the photogenerated electron transfer. The C–C coupling reaction was carried out under a N2 atmosphere, using methanol as a hole scavenger and toluene as the solvent and a 405 nm light excitation. The choice of solvent was found to be crucial in the photoreduction of 1a (in Figure 44a). In the presence of toluene, 1a forms 1b (in Figure 44a) with a yield of 77%, whereas this yield decreased for other solvents, such as hexane (~34), chlorobenzene (~24), and ethyl acetate (~14).
When the same reaction was performed for p-substituted benzyl bromides in the presence of toluene, three types of products were formed: homocoupling product 1b from substrate 1a, 2b from substrate 2a (in Figure 44a), heterocoupling product 2c from substrate 2a and toluene solvent, and the dehalogenated product 2d. On the other hand, when the same reaction was performed under a hexane solvent, only b and d products were formed. The number and yield of the product can be modulated by changing the solvent and electron-donating/electron-withdrawing nature of the substituent. The following trend for the C–C homocoupling has been observed: OCH3 < H < tBu < Br ≈ Cl in hexane solvent. An electron-donating substituent favours the dehalogenated product formation, while p-nitrobenzyl bromide did not undergo photoreduction in any of the solvents. These results suggest the different interaction of the substrate with the NC surface.
Furthermore, investigations into lead-free perovskite NCs have extended to their application in Csp3-Csp3 coupling reactions, such as α-alkylation of aldehydes. In a study conducted by Zhao and collaborators, oleic acid-capped Cs2AgSbCl6 double perovskite NCs, with a bandgap of 2.65 eV, were employed for the C–C coupling reaction involving 2-bromoacetophenone and n-octanal as a model reaction, mirroring the depicted reaction in Figure 7a [173]. The band alignment of Cs2AgSbCl6 NCs was found to resemble that of octylamine-capped CsPbBr3 NCs utilized for α-alkylation, as illustrated in Figure 7a. The authors achieved important reaction yields with a TON of ~14,800, particularly for substrates featuring both electron-donating and electron-withdrawing substituents on the phenyl ring of 2-bromoacetophenone. The meticulous washing of the NCs was imperative to facilitate efficient charge transfer and attain high product yields. The reaction mechanism involves a radical-mediated pathway akin to path III, illustrated in Figure 7b.

16. Conclusions

In this comprehensive review, we have introduced the recent advances in the field of metal halide perovskite-based organic molecular transformations, which have emerged as promising candidates for photocatalysis applications owing to their favourable optical and electronic properties as well as redox characteristics. The ease of preparation and recyclability further positions perovskite-based heterogeneous photocatalysis as superior to metal oxide/sulfide-based heterogeneous photocatalysts and molecular-based homogeneous photocatalysis.
Motivated by the success of Pb-based perovskite materials in solar cell applications, researchers have shifted their focus towards investigating organic molecular transformations. Pb-based perovskites have demonstrated remarkable efficacy in visible light-driven chemical reactions via charge and energy transfer processes, encompassing various organic reactions, such as C–H bond activation, C–X (X = C, N, O) bond formation, cycloaddition reactions, asymmetric synthesis, and chiral molecule synthesis. Leveraging the defect-tolerant nature of Pb-based PNCs with high PL QY, energy transfer-based photocatalytic reactions have yielded moderate-to-high reaction yields. It is noteworthy that smaller NCs exhibit superior photocatalytic activity, albeit with perovskite degradation becoming apparent after a few catalytic cycles.
In response to the instability observed in 3D Pb-based perovskites, researchers have explored low-dimensional and layered perovskites by incorporating larger-sized organic ligands. This approach not only enhances stability, particularly in polar solvents, such as water, but also yields promising results in decarboxylation and dehydrogenation reactions. Furthermore, efforts to address Pb-toxicity concerns have led to the proposal of Pb-free perovskite materials, including lower-dimensional structures, such as DPs and layered DPs. Surprisingly, lower-dimensional perovskite materials have exhibited superior photocatalytic activity compared to Pb-based perovskites, showcasing promise in toluene oxidation and C–C coupling reactions. Lead-free counterparts have demonstrated the ability to facilitate novel chemical reactions, e.g., by the ring-opening reactions of epoxides via Cs3Bi2Br9 due to their Lewis acidic character, along with high selectivity and turnover numbers for benzyl bromide coupling reactions. However, challenges persist due to the poor defect-related luminescence of lead-free counterparts, limiting their application in energy transfer-based photocatalytic reactions.
To optimize visible light absorption, electron–hole separation, and band alignment for accessing redox energy levels, researchers have successfully developed heterostructures, including type-II, Z-scheme, and Schottky heterojunctions. These heterostructures not only safeguard perovskite NCs but also enhance photocatalytic activity and stability. Nevertheless, challenges arise in directly comparing photocatalytic activity due to variations in material bandgap, morphology, and experimental conditions. Thorough characterization of heterostructures is crucial, with the potential existence of p-n junction characteristics in reported type-II heterojunctions requiring confirmation through comprehensive Fermi level characterization. Alternatively, tuning type-II heterojunctions into p-n heterojunctions through doping and changing the stoichiometry of perovskite [174] and other semiconductors (g-C3N4, Mxene) [175,176] could enhance the spatial separation of electrons and holes, thereby promoting efficient photocatalysis.
Furthermore, the engineering of materials and optimization of reaction conditions are vital for surpassing the current state-of-the-art methods in photocatalytic organic chemical transformations. Layered perovskites, exemplified by Cs4ZnSb2Cl12, exhibit promising bandgaps, excellent transport properties, and quantum confinement effects, surpassing the performance of both Pb-based and other Pb-free perovskites. The creation of heterostructures holds immense potential for expanding the scope of photochemical reactions [177].
Despite achieving high performance, the recycling stability of the state-of-the-art photocatalysts is currently limited to three to five cycles. To address this challenge, a promising avenue involves exploring core/shell heterostructure nanocrystals (NCs), with a specific focus on enhancing stability through optimized shell thickness and encapsulation strategies. Examples include CsPbBr3@CdS [178], CsPbBr3@Cs4PbBr6, CsPbBr3@CsPb2Br5 [179], CsPbBr3@Rb4PbBr6 [180], CsPbBr3@PbBr2 [181]. Additionally, encapsulating the active material with inorganic oxides such as SiO2 [182], TiO2 [183], AlOx [184], ZrO2 [185], silsesquioxane [186], and ZrNx [187] as well as hydrophobic polymer matrices, e.g., polystyrene [188], poly(maleic anhydride-alt-1-octadecene) [189], and PMMA [190].
While, porous metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) are currently the primary choices for encapsulation [191,192], there is a need for the development of novel, easily synthesizable encapsulation materials to usher in a new era of stability. Semiconducting polymers and monodispersed capping agents, boasting suitable porous structures, hydrophobicity, chemical interactions, and band energies, present themselves as potential game-changers in the stabilization paradigm [193]. Future investigations should pivot towards combined approaches that simultaneously elevate both activity and stability.
Incorporating surface functional groups and capping layers onto quantum dots (QDs) and NCs allows for harnessing the high activity of small-sized halide perovskite photocatalysts without compromising stability [194]. Moreover, the strategic construction of heterojunctions, specifically type-II and Z-scheme, at the nano and sub-nanoscale within porous semiconductor networks holds significant promise for revitalizing perovskite-based photocatalysis, ensuring heightened performance on both the activity and stability fronts [179,195,196,197,198].
Recent reports on perovskite solar cells suggest that orienting perovskite precursors, such as PbI2, in certain directions on the film can modulate the final perovskite crystal orientations (topochemical assembly) and minimize the defect state density of the perovskite crystal lattice, ultimately realizing high solar cell efficiencies with long stability [199]. This strategy can be applied in solar cell-type-based photocatalysts. Similar strategies can be extended to colloidal heterostructure NCs, such as using PbS QDs as seeds and allowing for the crystal growth of perovskite on the surface of PbS, enabling the formation of heterojunctions with smaller lattice mismatch and showing more stable perovskite crystal structures, especially in polar solvent mediums [195,200,201]. This methodology allows for stable iodide-based perovskites with narrow bandgaps for solar cell and photocatalytic applications [202].
Additionally, the investigation of lead-free perovskite materials, such as copper halides and chalcogenides, with favorable opto-electronic properties stands as a promising direction for advancing the field of photocatalysis [203,204].
In conclusion, this review not only highlights current achievements but also points towards exciting avenues for future research in the dynamic field of perovskite-based photocatalysis. The continuous exploration of materials, optimization strategies, ligand chemistry, and innovative heterostructure designs promises to propel the field towards more sustainable and efficient photocatalytic organic transformations.

Author Contributions

M.J. wrote and formatted the review; Supervision, R.E.G. and J.P.-P.; review editing and project administration, R.E.G. and J.P.-P.; funding acquisition, R.E.G. and J.P.-P. All authors have read and agreed to the published version of the manuscript.

Funding

We thanks to Agencia Estatal de Investigación-AEI (grant number MMCIU Unit of Excellence “Maria de Maeztu” CEX2019-000919-M), and Generalitat Valenciana (grant numbers CIPROM/2022/57), and MICIN and Generalitat Valenciana (MFA/2022/051) for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, W.; Tadé, M.O.; Shao, Z. Research progress of perovskite materials in photocatalysis- and photovoltaics-related energy conversion and environmental treatment. Chem. Soc. Rev. 2015, 44, 5371–5408. [Google Scholar] [CrossRef] [PubMed]
  2. Xuan, J.; Xiao, W.-J. Visible-Light Photoredox Catalysis. Angew. Chem. Int. Ed. 2012, 51, 6828–6838. [Google Scholar] [CrossRef] [PubMed]
  3. Prier, C.K.; Rankic, D.A.; MacMillan, D.W.C. Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis. Chem. Rev. 2013, 113, 5322–5363. [Google Scholar] [CrossRef] [PubMed]
  4. Ravelli, D.; Dondi, D.; Fagnoni, M.; Albini, A. Photocatalysis. A multi-faceted concept for green chemistry. Chem. Soc. Rev. 2009, 38, 1999–2011. [Google Scholar] [CrossRef] [PubMed]
  5. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  6. Gisbertz, S.; Pieber, B. Heterogeneous Photocatalysis in Organic Synthesis. ChemPhotoChem 2020, 4, 456–475. [Google Scholar] [CrossRef]
  7. Caputo, J.A.; Frenette, L.C.; Zhao, N.; Sowers, K.L.; Krauss, T.D.; Weix, D.J. General and Efficient C–C Bond Forming Photoredox Catalysis with Semiconductor Quantum Dots. J. Am. Chem. Soc. 2017, 139, 4250–4253. [Google Scholar] [CrossRef] [PubMed]
  8. Li, X.-B.; Li, Z.-J.; Gao, Y.-J.; Meng, Q.-Y.; Yu, S.; Weiss, R.G.; Tung, C.-H.; Wu, L.-Z. Mechanistic Insights into the Interface-Directed Transformation of Thiols into Disulfides and Molecular Hydrogen by Visible-Light Irradiation of Quantum Dots. Angew. Chem. Int. Ed. 2014, 53, 2085–2089. [Google Scholar] [CrossRef]
  9. Chai, Z.; Zeng, T.-T.; Li, Q.; Lu, L.-Q.; Xiao, W.-J.; Xu, D. Efficient Visible Light-Driven Splitting of Alcohols into Hydrogen and Corresponding Carbonyl Compounds over a Ni-Modified CdS Photocatalyst. J. Am. Chem. Soc. 2016, 138, 10128–10131. [Google Scholar] [CrossRef]
  10. Yuan, J.; Liu, H.; Wang, S.; Li, X. How to apply metal halide perovskites to photocatalysis: Challenges and development. Nanoscale 2021, 13, 10281–10304. [Google Scholar] [CrossRef]
  11. Jin, N.; Sun, Y.; Shi, W.; Wang, P.; Nagaoka, Y.; Cai, T.; Wu, R.; Dube, L.; Nyiera, H.N.; Liu, Y.; et al. Type-I CdS/ZnS Core/Shell Quantum Dot-Gold Heterostructural Nanocrystals for Enhanced Photocatalytic Hydrogen Generation. J. Am. Chem. Soc. 2023, 145, 21886–21896. [Google Scholar] [CrossRef]
  12. Cai, T.; Shi, W.; Hwang, S.; Kobbekaduwa, K.; Nagaoka, Y.; Yang, H.; Hills-Kimball, K.; Zhu, H.; Wang, J.; Wang, Z.; et al. Lead-Free Cs4CuSb2Cl12 Layered Double Perovskite Nanocrystals. J. Am. Chem. Soc. 2020, 142, 11927–11936. [Google Scholar] [CrossRef] [PubMed]
  13. Huang, H.; Bodnarchuk, M.I.; Kershaw, S.V.; Kovalenko, M.V.; Rogach, A.L. Lead Halide Perovskite Nanocrystals in the Research Spotlight: Stability and Defect Tolerance. ACS Energy Lett. 2017, 2, 2071–2083. [Google Scholar] [CrossRef] [PubMed]
  14. Ravi, V.K.; Markad, G.B.; Nag, A. Band Edge Energies and Excitonic Transition Probabilities of Colloidal CsPbX3 (X = Cl, Br, I) Perovskite Nanocrystals. ACS Energy Lett. 2016, 1, 665–671. [Google Scholar] [CrossRef]
  15. Xing, G.; Mathews, N.; Sun, S.; Lim, S.S.; Lam, Y.M.; Grätzel, M.; Mhaisalkar, S.; Sum, T.C. Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342, 344–347. [Google Scholar] [CrossRef] [PubMed]
  16. Yettapu, G.R.; Talukdar, D.; Sarkar, S.; Swarnkar, A.; Nag, A.; Ghosh, P.; Mandal, P. Terahertz Conductivity within Colloidal CsPbBr3 Perovskite Nanocrystals: Remarkably High Carrier Mobilities and Large Diffusion Lengths. Nano Lett. 2016, 16, 4838–4848. [Google Scholar] [CrossRef] [PubMed]
  17. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [PubMed]
  18. Subramanian Periyal, S.; Jagadeeswararao, M.; Ng, S.E.; John, R.A.; Mathews, N. Halide Perovskite Quantum Dots Photosensitized-Amorphous Oxide Transistors for Multimodal Synapses. Adv. Mater. Technol. 2020, 5, 2000514. [Google Scholar] [CrossRef]
  19. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 19 November 2023).
  20. Filip, M.R.; Verdi, C.; Giustino, F. GW Band Structures and Carrier Effective Masses of CH3NH3PbI3 and Hypothetical Perovskites of the Type APbI3: A = NH4, PH4, AsH4, and SbH4. J. Phys. Chem. C 2015, 119, 25209–25219. [Google Scholar] [CrossRef]
  21. Vashishtha, P.; Bishnoi, S.; Li, C.H.A.; Jagadeeswararao, M.; Hooper, T.J.N.; Lohia, N.; Shivarudraiah, S.B.; Ansari, M.S.; Sharma, S.N.; Halpert, J.E. Recent Advancements in Near-Infrared Perovskite Light-Emitting Diodes. ACS Appl. Electron. Mater. 2020, 2, 3470–3490. [Google Scholar] [CrossRef]
  22. John, R.A.; Shah, N.; Vishwanath, S.K.; Ng, S.E.; Febriansyah, B.; Jagadeeswararao, M.; Chang, C.-H.; Basu, A.; Mathews, N. Halide perovskite memristors as flexible and reconfigurable physical unclonable functions. Nat. Commun. 2021, 12, 3681. [Google Scholar] [CrossRef] [PubMed]
  23. Park, S.; Chang, W.J.; Lee, C.W.; Park, S.; Ahn, H.-Y.; Nam, K.T. Photocatalytic hydrogen generation from hydriodic acid using methylammonium lead iodide in dynamic equilibrium with aqueous solution. Nat. Energy 2016, 2, 16185. [Google Scholar] [CrossRef]
  24. Temerov, F.; Baghdadi, Y.; Rattner, E.; Eslava, S. A Review on Halide Perovskite-Based Photocatalysts: Key Factors and Challenges. ACS Appl. Energy Mater. 2022, 5, 14605–14637. [Google Scholar] [CrossRef] [PubMed]
  25. Ketavath, R.; Mohan, L.; Sumukam, R.R.; Alsulami, Q.A.; Premalatha, A.; Murali, B. Can perovskites be efficient photocatalysts in organic transformations? J. Mater. Chem. A 2022, 10, 12317–12333. [Google Scholar] [CrossRef]
  26. Zhu, X.; Lin, Y.; San Martin, J.; Sun, Y.; Zhu, D.; Yan, Y. Lead halide perovskites for photocatalytic organic synthesis. Nat. Commun. 2019, 10, 2843. [Google Scholar] [CrossRef] [PubMed]
  27. Vargas, B.; Ramos, E.; Pérez-Gutiérrez, E.; Alonso, J.C.; Solis-Ibarra, D. A Direct Bandgap Copper–Antimony Halide Perovskite. J. Am. Chem. Soc. 2017, 139, 9116–9119. [Google Scholar] [CrossRef] [PubMed]
  28. Singhal, N.; Chakraborty, R.; Ghosh, P.; Nag, A. Low-Bandgap Cs4CuSb2Cl12 Layered Double Perovskite: Synthesis, Reversible Thermal Changes, and Magnetic Interaction. Chem.–Asian J. 2018, 13, 2085–2092. [Google Scholar] [CrossRef] [PubMed]
  29. Mai, H.; Li, X.; Lu, J.; Wen, X.; Le, T.C.; Russo, S.P.; Chen, D.; Caruso, R.A. Synthesis of Layered Lead-Free Perovskite Nanocrystals with Precise Size and Shape Control and Their Photocatalytic Activity. J. Am. Chem. Soc. 2023, 145, 17337–17350. [Google Scholar] [CrossRef]
  30. Huang, H.; Pradhan, B.; Hofkens, J.; Roeffaers, M.B.J.; Steele, J.A. Solar-Driven Metal Halide Perovskite Photocatalysis: Design, Stability, and Performance. ACS Energy Lett. 2020, 5, 1107–1123. [Google Scholar] [CrossRef]
  31. Zhang, Z.; Zhou, R.; Li, D.; Jiang, Y.; Wang, X.; Tang, H.; Xu, J. Recent Progress in Halide Perovskite Nanocrystals for Photocatalytic Hydrogen Evolution. Nanomaterials 2023, 13, 106. [Google Scholar] [CrossRef]
  32. Dandia, A.; Saini, P.; Sharma, R.; Parewa, V. Visible light driven perovskite-based photocatalysts: A new candidate for green organic synthesis by photochemical protocol. Curr. Res. Green Sustain. Chem. 2020, 3, 100031. [Google Scholar] [CrossRef]
  33. Bresolin, B.-M.; Park, Y.; Bahnemann, D.W. Recent Progresses on Metal Halide Perovskite-Based Material as Potential Photocatalyst. Catalysts 2020, 10, 709. [Google Scholar] [CrossRef]
  34. Ren, K.; Yue, S.; Li, C.; Fang, Z.; Gasem, K.A.M.; Leszczynski, J.; Qu, S.; Wang, Z.; Fan, M. Metal halide perovskites for photocatalysis applications. J. Mater. Chem. A 2022, 10, 407–429. [Google Scholar] [CrossRef]
  35. Wang, X.; Peng, Y.; Yang, S.; Yang, H.G.; Hou, Y. Recent progress in metal halide perovskite photocatalysts for hydrogen evolution. Mater. Chem. Front. 2023, 7, 4635–4657. [Google Scholar] [CrossRef]
  36. Wu, W.-B.; Wong, Y.-C.; Tan, Z.-K.; Wu, J. Photo-induced thiol coupling and C–H activation using nanocrystalline lead-halide perovskite catalysts. Catal. Sci. Technol. 2018, 8, 4257–4263. [Google Scholar] [CrossRef]
  37. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with Wide Color Gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed]
  38. Dey, A.; Ye, J.; De, A.; Debroye, E.; Ha, S.K.; Bladt, E.; Kshirsagar, A.S.; Wang, Z.; Yin, J.; Wang, Y.; et al. State of the Art and Prospects for Halide Perovskite Nanocrystals. ACS Nano 2021, 15, 10775–10981. [Google Scholar] [CrossRef] [PubMed]
  39. Schmidt, L.C.; Pertegás, A.; González-Carrero, S.; Malinkiewicz, O.; Agouram, S.; Mínguez Espallargas, G.; Bolink, H.J.; Galian, R.E.; Pérez-Prieto, J. Nontemplate Synthesis of CH3NH3PbBr3 Perovskite Nanoparticles. J. Am. Chem. Soc. 2014, 136, 850–853. [Google Scholar] [CrossRef] [PubMed]
  40. Butkus, J.; Vashishtha, P.; Chen, K.; Gallaher, J.K.; Prasad, S.K.K.; Metin, D.Z.; Laufersky, G.; Gaston, N.; Halpert, J.E.; Hodgkiss, J.M. The Evolution of Quantum Confinement in CsPbBr3 Perovskite Nanocrystals. Chem. Mater. 2017, 29, 3644–3652. [Google Scholar] [CrossRef]
  41. Yang, B.; Chen, J.; Yang, S.; Hong, F.; Sun, L.; Han, P.; Pullerits, T.; Deng, W.; Han, K. Lead-Free Silver-Bismuth Halide Double Perovskite Nanocrystals. Angew. Chem. Int. Ed. 2018, 57, 5359–5363. [Google Scholar] [CrossRef]
  42. Huang, S.; Li, Z.; Kong, L.; Zhu, N.; Shan, A.; Li, L. Enhancing the Stability of CH3NH3PbBr3 Quantum Dots by Embedding in Silica Spheres Derived from Tetramethyl Orthosilicate in “Waterless” Toluene. J. Am. Chem. Soc. 2016, 138, 5749–5752. [Google Scholar] [CrossRef]
  43. Zhong, Q.; Cao, M.; Hu, H.; Yang, D.; Chen, M.; Li, P.; Wu, L.; Zhang, Q. One-Pot Synthesis of Highly Stable CsPbBr3@SiO2 Core–Shell Nanoparticles. ACS Nano 2018, 12, 8579–8587. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, Q.; Lan, X.; Ren, C.; Qi, B.; Shi, J. Multi-dimensional collaborations boost lead halide perovskite driven superior and long-period CO2 photoreduction under liquid-phase H2O environment. J. Alloys Compd. 2022, 899, 163316. [Google Scholar] [CrossRef]
  45. van der Stam, W.; Geuchies, J.J.; Altantzis, T.; van den Bos, K.H.W.; Meeldijk, J.D.; Van Aert, S.; Bals, S.; Vanmaekelbergh, D.; de Mello Donega, C. Highly Emissive Divalent-Ion-Doped Colloidal CsPb1–xMxBr3 Perovskite Nanocrystals through Cation Exchange. J. Am. Chem. Soc. 2017, 139, 4087–4097. [Google Scholar] [CrossRef] [PubMed]
  46. Hu, Q.; Li, Z.; Tan, Z.; Song, H.; Ge, C.; Niu, G.; Han, J.; Tang, J. Rare Earth Ion-Doped CsPbBr3 Nanocrystals. Adv. Opt. Mater. 2018, 6, 1700864. [Google Scholar] [CrossRef]
  47. Pan, G.; Bai, X.; Yang, D.; Chen, X.; Jing, P.; Qu, S.; Zhang, L.; Zhou, D.; Zhu, J.; Xu, W.; et al. Doping Lanthanide into Perovskite Nanocrystals: Highly Improved and Expanded Optical Properties. Nano Lett. 2017, 17, 8005–8011. [Google Scholar] [CrossRef] [PubMed]
  48. Mir, W.J.; Mahor, Y.; Lohar, A.; Jagadeeswararao, M.; Das, S.; Mahamuni, S.; Nag, A. Postsynthesis Doping of Mn and Yb into CsPbX3 (X = Cl, Br, or I) Perovskite Nanocrystals for Downconversion Emission. Chem. Mater. 2018, 30, 8170–8178. [Google Scholar] [CrossRef]
  49. Mir, W.J.; Jagadeeswararao, M.; Das, S.; Nag, A. Colloidal Mn-Doped Cesium Lead Halide Perovskite Nanoplatelets. ACS Energy Lett. 2017, 2, 537–543. [Google Scholar] [CrossRef]
  50. Jellicoe, T.C.; Richter, J.M.; Glass, H.F.J.; Tabachnyk, M.; Brady, R.; Dutton, S.E.; Rao, A.; Friend, R.H.; Credgington, D.; Greenham, N.C.; et al. Synthesis and Optical Properties of Lead-Free Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2016, 138, 2941–2944. [Google Scholar] [CrossRef]
  51. Zhang, J.; Yang, Y.; Deng, H.; Farooq, U.; Yang, X.; Khan, J.; Tang, J.; Song, H. High Quantum Yield Blue Emission from Lead-Free Inorganic Antimony Halide Perovskite Colloidal Quantum Dots. ACS Nano 2017, 11, 9294–9302. [Google Scholar] [CrossRef]
  52. Creutz, S.E.; Crites, E.N.; De Siena, M.C.; Gamelin, D.R. Colloidal Nanocrystals of Lead-Free Double-Perovskite (Elpasolite) Semiconductors: Synthesis and Anion Exchange To Access New Materials. Nano Lett. 2018, 18, 1118–1123. [Google Scholar] [CrossRef] [PubMed]
  53. Prasanna, R.; Gold-Parker, A.; Leijtens, T.; Conings, B.; Babayigit, A.; Boyen, H.-G.; Toney, M.F.; McGehee, M.D. Band Gap Tuning via Lattice Contraction and Octahedral Tilting in Perovskite Materials for Photovoltaics. J. Am. Chem. Soc. 2017, 139, 11117–11124. [Google Scholar] [CrossRef] [PubMed]
  54. Byranvand, M.M.; Otero-Martínez, C.; Ye, J.; Zuo, W.; Manna, L.; Saliba, M.; Hoye, R.L.Z.; Polavarapu, L. Recent Progress in Mixed A-Site Cation Halide Perovskite Thin-Films and Nanocrystals for Solar Cells and Light-Emitting Diodes. Adv. Opt. Mater. 2022, 10, 2200423. [Google Scholar] [CrossRef]
  55. Hazarika, A.; Zhao, Q.; Gaulding, E.A.; Christians, J.A.; Dou, B.; Marshall, A.R.; Moot, T.; Berry, J.J.; Johnson, J.C.; Luther, J.M. Perovskite Quantum Dot Photovoltaic Materials beyond the Reach of Thin Films: Full-Range Tuning of A-Site Cation Composition. ACS Nano 2018, 12, 10327–10337. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, Z.; Yates, J.T., Jr. Band Bending in Semiconductors: Chemical and Physical Consequences at Surfaces and Interfaces. Chem. Rev. 2012, 112, 5520–5551. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, Y.-F.; Yang, M.-Z.; Chen, B.-X.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 5660–5663. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, H.; Verhaeghe, D.; Weng, B.; Ghosh, B.; Zhang, H.; Hofkens, J.; Steele, J.A.; Roeffaers, M.B.J. Metal Halide Perovskite Based Heterojunction Photocatalysts. Angew. Chem. Int. Ed. 2022, 61, e202203261. [Google Scholar] [CrossRef]
  59. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  60. Bard, A.J. Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors. J. Photochem. 1979, 10, 59–75. [Google Scholar] [CrossRef]
  61. Schünemann, S.; van Gastel, M.; Tüysüz, H. A CsPbBr3/TiO2 Composite for Visible-Light-Driven Photocatalytic Benzyl Alcohol Oxidation. ChemSusChem 2018, 11, 2057–2061. [Google Scholar] [CrossRef]
  62. Xu, F.; Meng, K.; Cheng, B.; Wang, S.; Xu, J.; Yu, J. Unique S-scheme heterojunctions in self-assembled TiO2/CsPbBr3 hybrids for CO2 photoreduction. Nat. Commun. 2020, 11, 4613. [Google Scholar] [CrossRef] [PubMed]
  63. Ravelli, D.; Protti, S.; Fagnoni, M. Carbon–Carbon Bond Forming Reactions via Photogenerated Intermediates. Chem. Rev. 2016, 116, 9850–9913. [Google Scholar] [CrossRef] [PubMed]
  64. Romero, N.A.; Nicewicz, D.A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef] [PubMed]
  65. Zhu, X.; Lin, Y.; Sun, Y.; Beard, M.C.; Yan, Y. Lead-Halide Perovskites for Photocatalytic α-Alkylation of Aldehydes. J. Am. Chem. Soc. 2019, 141, 733–738. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, K.; Lu, H.; Zhu, X.; Lin, Y.; Beard, M.C.; Yan, Y.; Chen, X. Ultrafast Reaction Mechanisms in Perovskite Based Photocatalytic C–C Coupling. ACS Energy Lett. 2020, 5, 566–571. [Google Scholar] [CrossRef]
  67. Rosa-Pardo, I.; Casadevall, C.; Schmidt, L.; Claros, M.; Galian, R.E.; Lloret-Fillol, J.; Pérez-Prieto, J. The synergy between the CsPbBr3 nanoparticle surface and the organic ligand becomes manifest in a demanding carbon–carbon coupling reaction. Chem. Commun. 2020, 56, 5026–5029. [Google Scholar] [CrossRef]
  68. 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. Int. Ed. 2020, 59, 22563–22569. [Google Scholar] [CrossRef] [PubMed]
  69. Fiuza-Maneiro, N.; Sun, K.; López-Fernández, I.; Gómez-Graña, S.; Müller-Buschbaum, P.; Polavarapu, L. Ligand Chemistry of Inorganic Lead Halide Perovskite Nanocrystals. ACS Energy Lett. 2023, 8, 1152–1191. [Google Scholar] [CrossRef]
  70. Mir, W.J.; Alamoudi, A.; Yin, J.; Yorov, K.E.; Maity, P.; Naphade, R.; Shao, B.; Wang, J.; Lintangpradipto, M.N.; Nematulloev, S.; et al. Lecithin Capping Ligands Enable Ultrastable Perovskite-Phase CsPbI3 Quantum Dots for Rec. 2020 Bright-Red Light-Emitting Diodes. J. Am. Chem. Soc. 2022, 144, 13302–13310. [Google Scholar] [CrossRef]
  71. Manna, A.; Dinda, T.K.; Ghosh, S.; Mal, P. CsPbBr3 in the Activation of the C–Br Bond of CBrX3 (X = Cl, Br) under Sunlight. Chem. Mater. 2023, 35, 628–637. [Google Scholar] [CrossRef]
  72. Shi, T.; Sun, K.; Chen, X.-L.; Zhang, Z.-X.; Huang, X.-Q.; Peng, Y.-Y.; Qu, L.-B.; Yu, B. Recyclable Perovskite as Heterogeneous Photocatalyst for Aminomethylation of Imidazo-Fused Heterocycles. Adv. Synth. Catal. 2020, 362, 2143–2149. [Google Scholar] [CrossRef]
  73. Rohani, S.; Ziarati, A.; Ziarani, G.M.; Badiei, A.; Burgi, T. Engineering of highly active Au/Pd supported on hydrogenated urchin-like yolk@shell TiO2 for visible light photocatalytic Suzuki coupling. Catal. Sci. Technol. 2019, 9, 3820–3827. [Google Scholar] [CrossRef]
  74. Jiao, Z.; Zhai, Z.; Guo, X.; Guo, X.-Y. Visible-Light-Driven Photocatalytic Suzuki–Miyaura Coupling Reaction on Mott–Schottky-type Pd/SiC Catalyst. J. Phys. Chem. C 2015, 119, 3238–3243. [Google Scholar] [CrossRef]
  75. Raza, F.; Yim, D.; Park, J.H.; Kim, H.-I.; Jeon, S.-J.; Kim, J.-H. Structuring Pd Nanoparticles on 2H-WS2 Nanosheets Induces Excellent Photocatalytic Activity for Cross-Coupling Reactions under Visible Light. J. Am. Chem. Soc. 2017, 139, 14767–14774. [Google Scholar] [CrossRef] [PubMed]
  76. Meng, F.; Liu, Y.; Wang, J.; Tan, X.; Sun, H.; Liu, S.; Wang, S. Temperature dependent photocatalysis of g-C3N4, TiO2 and ZnO: Differences in photoactive mechanism. J. Colloid Interface Sci. 2018, 532, 321–330. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Y.; Ren, P.; Zhang, D.; Qiao, W.; Wang, D.; Yang, X.; Wen, X.; Rummeli, M.H.; Niemantsverdriet, H.; Lewis, J.P.; et al. Rationally Designed Metal Cocatalyst for Selective Photosynthesis of Bibenzyls via Dehalogenative C–C Homocoupling. ACS Catal. 2021, 11, 4338–4348. [Google Scholar] [CrossRef]
  78. Wang, C.; Weng, B.; Keshavarz, M.; Yang, M.-Q.; Huang, H.; Ding, Y.; Lai, F.; Aslam, I.; Jin, H.; Romolini, G.; et al. Photothermal Suzuki Coupling Over a Metal Halide Perovskite/Pd Nanocube Composite Catalyst. ACS Appl. Mater. Interfaces 2022, 14, 17185–17194. [Google Scholar] [CrossRef]
  79. Li, X.-H.; Baar, M.; Blechert, S.; Antonietti, M. Facilitating room-temperature Suzuki coupling reaction with light: Mott-Schottky photocatalyst for C-C-coupling. Sci. Rep. 2013, 3, 1743. [Google Scholar] [CrossRef]
  80. Hong, Z.; Chong, W.K.; Ng, A.Y.R.; Li, M.; Ganguly, R.; Sum, T.C.; Soo, H.S. Hydrophobic Metal Halide Perovskites for Visible-Light Photoredox C–C Bond Cleavage and Dehydrogenation Catalysis. Angew. Chem. Int. Ed. 2019, 58, 3456–3460. [Google Scholar] [CrossRef]
  81. Zou, S.; Liu, Y.; Li, J.; Liu, C.; Feng, R.; Jiang, F.; Li, Y.; Song, J.; Zeng, H.; Hong, M.; et al. Stabilizing Cesium Lead Halide Perovskite Lattice through Mn(II) Substitution for Air-Stable Light-Emitting Diodes. J. Am. Chem. Soc. 2017, 139, 11443–11450. [Google Scholar] [CrossRef]
  82. Mu, Y.-F.; Zhang, W.; Guo, X.-X.; Dong, G.-X.; Zhang, M.; Lu, T.-B. Water-Tolerant Lead Halide Perovskite Nanocrystals as Efficient Photocatalysts for Visible-Light-Driven CO2 Reduction in Pure Water. ChemSusChem 2019, 12, 4769–4774. [Google Scholar] [CrossRef]
  83. Tang, C.; Chen, C.; Xu, W.; Xu, L. Design of doped cesium lead halide perovskite as a photo-catalytic CO2 reduction catalyst. J. Mater. Chem. A 2019, 7, 6911–6919. [Google Scholar] [CrossRef]
  84. Shyamal, S.; Dutta, S.K.; Pradhan, N. Doping Iron in CsPbBr3 Perovskite Nanocrystals for Efficient and Product Selective CO2 Reduction. J. Phys. Chem. Lett. 2019, 10, 7965–7969. [Google Scholar] [CrossRef]
  85. Martin, J.S.; Zeng, X.; Chen, X.; Miller, C.; Han, C.; Lin, Y.; Yamamoto, N.; Wang, X.; Yazdi, S.; Yan, Y.; et al. A Nanocrystal Catalyst Incorporating a Surface Bound Transition Metal to Induce Photocatalytic Sequential Electron Transfer Events. J. Am. Chem. Soc. 2021, 143, 11361–11369. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, K.; Deng, X.; Dodekatos, G.; Tüysüz, H. Photocatalytic Polymerization of 3,4-Ethylenedioxythiophene over Cesium Lead Iodide Perovskite Quantum Dots. J. Am. Chem. Soc. 2017, 139, 12267–12273. [Google Scholar] [CrossRef] [PubMed]
  87. Li, Y.; Shu, Q.; Du, Q.; Dai, Y.; Zhao, S.; Zhang, J.; Li, L.; Chen, K. Surface Modification for Improving the Photocatalytic Polymerization of 3,4-Ethylenedioxythiophene over Inorganic Lead Halide Perovskite Quantum Dots. ACS Appl. Mater. Interfaces 2020, 12, 451–460. [Google Scholar] [CrossRef]
  88. Trivedi, M.V.; Laurence, J.S.; Siahaan, T.J. The Role of Thiols and Disulfides on Protein Stability. Curr. Protein Pept. Sci. 2009, 10, 614–625. [Google Scholar] [CrossRef] [PubMed]
  89. Kirkpatrick, D.L.; Powis, G.; Hornsveld, M.; Dansen, T.B.; Maulucci, G.; Bačić, G.; Bridal, L.; Schmidt, H.H.; Tavitian, B.; Viel, T.; et al. Clinically Evaluated Cancer Drugs Inhibiting Redox Signaling. Antioxidants Redox Signal. 2017, 26, 262–273. [Google Scholar] [CrossRef]
  90. Berkmen, M. Production of disulfide-bonded proteins in Escherichia coli. Protein Expr. Purif. 2012, 82, 240–251. [Google Scholar] [CrossRef]
  91. Ramanathan, R.K.; Abbruzzese, J.; Dragovich, T.; Kirkpatrick, L.; Guillen, J.M.; Baker, A.F.; Pestano, L.A.; Green, S.; Von Hoff, D.D. A randomized phase II study of PX-12, an inhibitor of thioredoxin in patients with advanced cancer of the pancreas following progression after a gemcitabine-containing combination. Cancer Chemother. Pharmacol. 2011, 67, 503–509. [Google Scholar] [CrossRef]
  92. Huang, H.; Yuan, H.; Janssen, K.P.F.; Solís-Fernández, G.; Wang, Y.; Tan, C.Y.X.; Jonckheere, D.; Debroye, E.; Long, J.; Hendrix, J.; et al. Efficient and Selective Photocatalytic Oxidation of Benzylic Alcohols with Hybrid Organic–Inorganic Perovskite Materials. ACS Energy Lett. 2018, 3, 755–759. [Google Scholar] [CrossRef]
  93. McClelland, K.P.; Weiss, E.A. Selective Photocatalytic Oxidation of Benzyl Alcohol to Benzaldehyde or C–C Coupled Products by Visible-Light-Absorbing Quantum Dots. ACS Appl. Energy Mater. 2019, 2, 92–96. [Google Scholar] [CrossRef]
  94. Zheng, Z.; Wang, T.; Han, F.; Yang, Q.; Li, B. Synthesis of Ni modified Au@CdS core–shell nanostructures for enhancing photocatalytic coproduction of hydrogen and benzaldehyde under visible light. J. Colloid Interface Sci. 2022, 606, 47–56. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, S.G.; Kang, M.J.; Park, M.; Kim, K.-j.; Lee, H.; Kim, H.S. Selective photocatalytic conversion of benzyl alcohol to benzaldehyde or deoxybenzoin over ion-exchanged CdS. Appl. Catal. B Environ. 2022, 304, 120967. [Google Scholar] [CrossRef]
  96. Ramesh, A.; Da, C.T.; Manigandan, R.; Bhargav, P.B.; Nguyen-Le, M.-T. Selectivity oxidation of benzyl alcohol using mesoporous g-C3N4 catalysts prepared by hard template method. Colloid Interface Sci. Commun. 2022, 48, 100608. [Google Scholar] [CrossRef]
  97. Huang, H.; Yuan, H.; Zhao, J.; Solís-Fernández, G.; Zhou, C.; Seo, J.W.; Hendrix, J.; Debroye, E.; Steele, J.A.; Hofkens, J.; et al. C(sp3)–H Bond Activation by Perovskite Solar Photocatalyst Cell. ACS Energy Lett. 2019, 4, 203–208. [Google Scholar] [CrossRef]
  98. Dai, Y.; Tüysüz, H. Lead-Free Cs3Bi2Br9 Perovskite as Photocatalyst for Ring-Opening Reactions of Epoxides. ChemSusChem 2019, 12, 2587–2592. [Google Scholar] [CrossRef]
  99. Wong, Y.-C.; De Andrew Ng, J.; Tan, Z.-K. Perovskite-Initiated Photopolymerization for Singly Dispersed Luminescent Nanocomposites. Adv. Mater. 2018, 30, 1800774. [Google Scholar] [CrossRef]
  100. Long, G.; Sabatini, R.; Saidaminov, M.I.; Lakhwani, G.; Rasmita, A.; Liu, X.; Sargent, E.H.; Gao, W. Chiral-perovskite optoelectronics. Nat. Rev. Mater. 2020, 5, 423–439. [Google Scholar] [CrossRef]
  101. Mishra, K.; Guyon, D.; San Martin, J.; Yan, Y. Chiral Perovskite Nanocrystals for Asymmetric Reactions: A Highly Enantioselective Strategy for Photocatalytic Synthesis of N–C Axially Chiral Heterocycles. J. Am. Chem. Soc. 2023, 145, 17242–17252. [Google Scholar] [CrossRef]
  102. Ghosh, I.; Shaikh, R.S.; König, B. Sensitization-Initiated Electron Transfer for Photoredox Catalysis. Angew. Chem. Int. Ed. 2017, 56, 8544–8549. [Google Scholar] [CrossRef]
  103. Ravetz, B.D.; Pun, A.B.; Churchill, E.M.; Congreve, D.N.; Rovis, T.; Campos, L.M. Photoredox catalysis using infrared light via triplet fusion upconversion. Nature 2019, 565, 343–346. [Google Scholar] [CrossRef]
  104. Schulze, T.F.; Schmidt, T.W. Photochemical upconversion: Present status and prospects for its application to solar energy conversion. Energy Environ. Sci. 2015, 8, 103–125. [Google Scholar] [CrossRef]
  105. Schmidt, T.W.; Castellano, F.N. Photochemical Upconversion: The Primacy of Kinetics. J. Phys. Chem. Lett. 2014, 5, 4062–4072. [Google Scholar] [CrossRef] [PubMed]
  106. Hu, H.; Meier, F.; Zhao, D.; Abe, Y.; Gao, Y.; Chen, B.; Salim, T.; Chia, E.E.M.; Qiao, X.; Deibel, C.; et al. Efficient Room-Temperature Phosphorescence from Organic–Inorganic Hybrid Perovskites by Molecular Engineering. Adv. Mater. 2018, 30, 1707621. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, C.; Zhao, J.; Wu, S.; Wang, Z.; Wu, W.; Ma, J.; Guo, S.; Huang, L. Intramolecular RET Enhanced Visible Light-Absorbing Bodipy Organic Triplet Photosensitizers and Application in Photooxidation and Triplet–Triplet Annihilation Upconversion. J. Am. Chem. Soc. 2013, 135, 10566–10578. [Google Scholar] [CrossRef] [PubMed]
  108. Moor, K.; Kim, J.-H.; Snow, S.; Kim, J.-H. [C70] Fullerene-sensitized triplet–triplet annihilation upconversion. Chem. Commun. 2013, 49, 10829–10831. [Google Scholar] [CrossRef] [PubMed]
  109. Wu, W.; Guo, H.; Wu, W.; Ji, S.; Zhao, J. Organic Triplet Sensitizer Library Derived from a Single Chromophore (BODIPY) with Long-Lived Triplet Excited State for Triplet–Triplet Annihilation Based Upconversion. J. Org. Chem. 2011, 76, 7056–7064. [Google Scholar] [CrossRef] [PubMed]
  110. Yanai, N.; Kimizuka, N. New Triplet Sensitization Routes for Photon Upconversion: Thermally Activated Delayed Fluorescence Molecules, Inorganic Nanocrystals, and Singlet-to-Triplet Absorption. Acc. Chem. Res. 2017, 50, 2487–2495. [Google Scholar] [CrossRef] [PubMed]
  111. Mongin, C.; Garakyaraghi, S.; Razgoniaeva, N.; Zamkov, M.; Castellano, F.N. Direct observation of triplet energy transfer from semiconductor nanocrystals. Science 2016, 351, 369–372. [Google Scholar] [CrossRef]
  112. Wu, M.; Congreve, D.N.; Wilson, M.W.B.; Jean, J.; Geva, N.; Welborn, M.; Van Voorhis, T.; Bulović, V.; Bawendi, M.G.; Baldo, M.A. Solid-state infrared-to-visible upconversion sensitized by colloidal nanocrystals. Nat. Photonics 2016, 10, 31–34. [Google Scholar] [CrossRef]
  113. Mase, K.; Okumura, K.; Yanai, N.; Kimizuka, N. Triplet sensitization by perovskite nanocrystals for photon upconversion. Chem. Commun. 2017, 53, 8261–8264. [Google Scholar] [CrossRef]
  114. Han, Y.; Luo, X.; Lai, R.; Li, Y.; Liang, G.; Wu, K. Visible-Light-Driven Sensitization of Naphthalene Triplets Using Quantum-Confined CsPbBr3 Nanocrystals. J. Phys. Chem. Lett. 2019, 10, 1457–1463. [Google Scholar] [CrossRef]
  115. DuBose, J.T.; Kamat, P.V. Energy Versus Electron Transfer: Managing Excited-State Interactions in Perovskite Nanocrystal–Molecular Hybrids. Chem. Rev. 2022, 122, 12475–12494. [Google Scholar] [CrossRef]
  116. Luo, X.; Lai, R.; Li, Y.; Han, Y.; Liang, G.; Liu, X.; Ding, T.; Wang, J.; Wu, K. Triplet Energy Transfer from CsPbBr3 Nanocrystals Enabled by Quantum Confinement. J. Am. Chem. Soc. 2019, 141, 4186–4190. [Google Scholar] [CrossRef]
  117. Liu, M.; Xia, P.; Zhao, G.; Nie, C.; Gao, K.; He, S.; Wang, L.; Wu, K. Energy-Transfer Photocatalysis Using Lead Halide Perovskite Nanocrystals: Sensitizing Molecular Isomerization and Cycloaddition. Angew. Chem. Int. Ed. 2022, 61, e202208241. [Google Scholar] [CrossRef]
  118. Lin, Y.; Avvacumova, M.; Zhao, R.; Chen, X.; Beard, M.C.; Yan, Y. Triplet Energy Transfer from Lead Halide Perovskite for Highly Selective Photocatalytic 2 + 2 Cycloaddition. ACS Appl. Mater. Interfaces 2022, 14, 25357–25365. [Google Scholar] [CrossRef] [PubMed]
  119. Coperet, C. C–H Bond Activation and Organometallic Intermediates on Isolated Metal Centers on Oxide Surfaces. Chem. Rev. 2010, 110, 656–680. [Google Scholar] [CrossRef] [PubMed]
  120. Li, X.-H.; Chen, J.-S.; Wang, X.; Sun, J.; Antonietti, M. Metal-Free Activation of Dioxygen by Graphene/g-C3N4 Nanocomposites: Functional Dyads for Selective Oxidation of Saturated Hydrocarbons. J. Am. Chem. Soc. 2011, 133, 8074–8077. [Google Scholar] [CrossRef] [PubMed]
  121. Kesavan, L.; Tiruvalam, R.; Rahim, M.H.A.; bin Saiman, M.I.; Enache, D.I.; Jenkins, R.L.; Dimitratos, N.; Lopez-Sanchez, J.A.; Taylor, S.H.; Knight, D.W.; et al. Solvent-Free Oxidation of Primary Carbon-Hydrogen Bonds in Toluene Using Au-Pd Alloy Nanoparticles. Science 2011, 331, 195–199. [Google Scholar] [CrossRef] [PubMed]
  122. Partenheimer, W. Methodology and scope of metal/bromide autoxidation of hydrocarbons. Catal. Today 1995, 23, 69–158. [Google Scholar] [CrossRef]
  123. Zhang, M.; Wang, W.; Gao, F.; Luo, D. g-C3N4-Stabilised Organic–Inorganic Halide Perovskites for Efficient Photocatalytic Selective Oxidation of Benzyl Alcohol. Catalysts 2021, 11, 505. [Google Scholar] [CrossRef]
  124. Zhao, Y.; Dai, Y.; Wang, Q.; Dong, Y.; Song, T.; Mudryi, A.; Chen, Q.; Li, Y. Anions-Exchange-Induced Efficient Carrier Transport at CsPbBrxCl3-x/TiO2 Interface for Photocatalytic Activation of C(sp3)–H bond in Toluene Oxidation. ChemCatChem 2021, 13, 2592–2598. [Google Scholar] [CrossRef]
  125. Zhu, E.; Zhao, Y.; Dai, Y.; Wang, Q.; Dong, Y.; Chen, Q.; Li, Y. Heterojunction-Type Photocatalytic System Based on Inorganic Halide Perovskite CsPbBr3 Chin. J. Chem. 2020, 38, 1718–1722. [Google Scholar] [CrossRef]
  126. Chen, T.; Li, M.; Shen, L.; Roeffaers, M.B.J.; Weng, B.; Zhu, H.; Chen, Z.; Yu, D.; Pan, X.; Yang, M.-Q.; et al. Photocatalytic Anaerobic Oxidation of Aromatic Alcohols Coupled with H2 Production Over CsPbBr3/GO-Pt Catalysts. Front. Chem. 2022, 10, 833784. [Google Scholar] [CrossRef] [PubMed]
  127. Wang, C.; Huang, H.; Weng, B.; Verhaeghe, D.; Keshavarz, M.; Jin, H.; Liu, B.; Xie, H.; Ding, Y.; Gao, Y.; et al. Planar heterojunction boosts solar-driven photocatalytic performance and stability of halide perovskite solar photocatalyst cell. Appl. Catal. B Environ. 2022, 301, 120760. [Google Scholar] [CrossRef]
  128. Yi, J.; Ke, S.; Lu, S.; Weng, B.; Shen, L.; Yang, X.; Xue, H.; Yang, M.-Q.; Qian, Q. High-efficiency visible-light-driven oxidation of primary C–H bonds in toluene over a CsPbBr3 perovskite supported by hierarchical TiO2 nanoflakes. Nanoscale 2023, 15, 14584–14594. [Google Scholar] [CrossRef]
  129. Clennan, E.L.; Pace, A. Advances in singlet oxygen chemistry. Tetrahedron 2005, 61, 6665–6691. [Google Scholar] [CrossRef]
  130. Alberti, M.N.; Orfanopoulos, M. Unraveling the Mechanism of the Singlet Oxygen Ene Reaction: Recent Computational and Experimental Approaches. Chem.–Eur. J. 2010, 16, 9414–9421. [Google Scholar] [CrossRef]
  131. Ogilby, P.R. Singlet oxygen: There is indeed something new under the sun. Chem. Soc. Rev. 2010, 39, 3181–3209. [Google Scholar] [CrossRef]
  132. Camussi, I.; Mannucci, B.; Speltini, A.; Profumo, A.; Milanese, C.; Malavasi, L.; Quadrelli, P. g-C3N4—Singlet Oxygen Made Easy for Organic Synthesis: Scope and Limitations. ACS Sustain. Chem. Eng. 2019, 7, 8176–8182. [Google Scholar] [CrossRef]
  133. Corti, M.; Chiara, R.; Romani, L.; Mannucci, B.; Malavasi, L.; Quadrelli, P. g-C3N4/metal halide perovskite composites as photocatalysts for singlet oxygen generation processes for the preparation of various oxidized synthons. Catal. Sci. Technol. 2021, 11, 2292–2298. [Google Scholar] [CrossRef]
  134. Dai, Y.; Tüysüz, H. Rapid Acidic Media Growth of Cs3Bi2Br9 Halide Perovskite Platelets for Photocatalytic Toluene Oxidation. Sol. RRL 2021, 5, 2100265. [Google Scholar] [CrossRef]
  135. Wang, C.; Weng, B.; Liao, Y.; Liu, B.; Keshavarz, M.; Ding, Y.; Huang, H.; Verhaeghe, D.; Steele, J.A.; Feng, W.; et al. Simultaneous photocatalytic H2 generation and organic synthesis over crystalline–amorphous Pd nanocube decorated Cs3Bi2Br9. Chem. Commun. 2022, 58, 10691–10694. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, C.; Ding, Y.; Liu, B.; Weng, B.; Hofkens, J.; Roeffaers, M.B.J. Crystal structure engineering of metal halide perovskites for photocatalytic organic synthesis. Chem. Commun. 2023, 59, 3122–3125. [Google Scholar] [CrossRef]
  137. Sun, Q.; Ye, W.; Wei, J.; Li, L.; Wang, J.; He, J.-H.; Lu, J.-M. Lead-free perovskite Cs3Bi2Br9 heterojunctions for highly efficient and selective photocatalysis under mild conditions. J. Alloys Compd. 2022, 893, 162326. [Google Scholar] [CrossRef]
  138. Eklund, P.; Beckers, M.; Jansson, U.; Högberg, H.; Hultman, L. The Mn+1AXn phases: Materials science and thin-film processing. Thin Solid Films 2010, 518, 1851–1878. [Google Scholar] [CrossRef]
  139. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Transition Metal Carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef]
  140. Pandey, P.; Sengupta, A.; Parmar, S.; Bansode, U.; Gosavi, S.; Swarnkar, A.; Muduli, S.; Mohite, A.D.; Ogale, S. CsPbBr3–Ti3C2Tx MXene QD/QD Heterojunction: Photoluminescence Quenching, Charge Transfer, and Cd Ion Sensing Application. ACS Appl. Nano Mater. 2020, 3, 3305–3314. [Google Scholar] [CrossRef]
  141. Li, Q.; Song, T.; Zhang, Y.; Wang, Q.; Yang, Y. Boosting Photocatalytic Activity and Stability of Lead-Free Cs3Bi2Br9 Perovskite Nanocrystals via In Situ Growth on Monolayer 2D Ti3C2Tx MXene for C–H Bond Oxidation. ACS Appl. Mater. Interfaces 2021, 13, 27323–27333. [Google Scholar] [CrossRef]
  142. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  143. Xu, J.; Brenner, T.J.K.; Chabanne, L.; Neher, D.; Antonietti, M.; Shalom, M. Liquid-Based Growth of Polymeric Carbon Nitride Layers and Their Use in a Mesostructured Polymer Solar Cell with Voc Exceeding 1 V. J. Am. Chem. Soc. 2014, 136, 13486–13489. [Google Scholar] [CrossRef]
  144. Xu, J.; Wang, G.; Fan, J.; Liu, B.; Cao, S.; Yu, J. g-C3N4 modified TiO2 nanosheets with enhanced photoelectric conversion efficiency in dye-sensitized solar cells. J. Power Sources 2015, 274, 77–84. [Google Scholar] [CrossRef]
  145. Shanker, G.S.; Markad, G.B.; Jagadeeswararao, M.; Bansode, U.; Nag, A. Colloidal Nanocomposite of TiN and N-Doped Few-Layer Graphene for Plasmonics and Electrocatalysis. ACS Energy Lett. 2017, 2, 2251–2256. [Google Scholar] [CrossRef]
  146. Bai, Z.-J.; Mao, Y.; Wang, B.-H.; Chen, L.; Tian, S.; Hu, B.; Li, Y.-J.; Au, C.-T.; Yin, S.-F. Tuning photocatalytic performance of Cs3Bi2Br9 perovskite by g-C3N4 for C(sp3)—H bond activation. Nano Res. 2023, 16, 6104–6112. [Google Scholar] [CrossRef]
  147. Yang, Y.; Chen, Z.; Huang, H.; Liu, Y.; Zou, J.; Shen, S.; Yan, J.; Zhang, J.; Zhuang, Z.; Luo, Z.; et al. Synergistic surface activation during photocatalysis on perovskite derivative sites in heterojunction. Appl. Catal. B Environ. 2023, 323, 122146. [Google Scholar] [CrossRef]
  148. Zhang, H.-H.; Zhou, Z.-C.; Dong, Y.-J.; Zhang, L.; Chen, H.-Y.; Kuang, D.-B. Constructing a Cs3Sb2Br9/g-C3N4 Hybrid for Photocatalytic Aromatic C(sp3)-H Bond Activation. Sol. RRL 2021, 5, 2100559. [Google Scholar] [CrossRef]
  149. Teng, Y.; Chen, J.-H.; Huang, Y.-H.; Zhou, Z.-C.; Wang, X.-D.; Kuang, D.-B.; Chen, H.-Y. Atom-triggered epitaxial growth of Bi-based perovskite heterojunctions for promoting interfacial charge transfer. Appl. Catal. B Environ. 2023, 335, 122889. [Google Scholar] [CrossRef]
  150. Bai, Z.-J.; Tian, S.; Zeng, T.-Q.; Chen, L.; Wang, B.-H.; Hu, B.; Wang, X.; Zhou, W.; Pan, J.-B.; Shen, S.; et al. Cs3Bi2Br9 Nanodots Stabilized on Defective BiOBr Nanosheets by Interfacial Chemical Bonding: Modulated Charge Transfer for Photocatalytic C(sp3)–H Bond Activation. ACS Catal. 2022, 12, 15157–15167. [Google Scholar] [CrossRef]
  151. Huang, H.; Zhao, J.; Du, Y.; Zhou, C.; Zhang, M.; Wang, Z.; Weng, Y.; Long, J.; Hofkens, J.; Steele, J.A.; et al. Direct Z-Scheme Heterojunction of Semicoherent FAPbBr3/Bi2WO6 Interface for Photoredox Reaction with Large Driving Force. ACS Nano 2020, 14, 16689–16697. [Google Scholar] [CrossRef]
  152. Wang, W.; Huang, H.; Ke, X.; Liu, X.; Yang, S.; Wang, K.; Huang, L.; Tu, C.; Zheng, Z.; Luo, D.; et al. Morphology- and size-dependent FAPbBr3/WO3 Z-scheme photocatalysts for the controllable photo-oxidation of benzyl alcohol. Mater. Des. 2022, 215, 110502. [Google Scholar] [CrossRef]
  153. Mu, Y.-F.; Zhang, C.; Zhang, M.-R.; Zhang, W.; Zhang, M.; Lu, T.-B. Direct Z-Scheme Heterojunction of Ligand-Free FAPbBr3/α-Fe2O3 for Boosting Photocatalysis of CO2 Reduction Coupled with Water Oxidation. ACS Appl. Mater. Interfaces 2021, 13, 22314–22322. [Google Scholar] [CrossRef] [PubMed]
  154. Jiang, Y.; Chen, H.-Y.; Li, J.-Y.; Liao, J.-F.; Zhang, H.-H.; Wang, X.-D.; Kuang, D.-B. Z-Scheme 2D/2D Heterojunction of CsPbBr3/Bi2WO6 for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2020, 30, 2004293. [Google Scholar] [CrossRef]
  155. Bai, Z.-J.; Tan, X.-P.; Chen, L.; Hu, B.; Tan, Y.-X.; Mao, Y.; Shen, S.; Guo, J.-K.; Au, C.-T.; Liang, Z.-W.; et al. Efficient photocatalytic toluene selective oxidation over Cs3Bi1.8Sb0.2Br9 Nanosheets: Enhanced charge carriers generation and C–H bond dissociation. Chem. Eng. Sci. 2022, 247, 116983. [Google Scholar] [CrossRef]
  156. Goyal, A.; McKechnie, S.; Pashov, D.; Tumas, W.; van Schilfgaarde, M.; Stevanović, V. Origin of Pronounced Nonlinear Band Gap Behavior in Lead–Tin Hybrid Perovskite Alloys. Chem. Mater. 2018, 30, 3920–3928. [Google Scholar] [CrossRef]
  157. Shi, M.; Zhou, H.; Tian, W.; Yang, B.; Yang, S.; Han, K.; Li, R.; Li, C. Lead-free B-site bimetallic perovskite photocatalyst for efficient benzylic C–H bond activation. Cell Rep. Phys. Sci. 2021, 2, 100656. [Google Scholar] [CrossRef]
  158. Cui, Z.; Wu, Y.; Zhang, S.; Fu, H.; Chen, G.; Lou, Z.; Liu, X.; Zhang, Q.; Wang, Z.; Zheng, Z.; et al. Insight into a strategy to improve charge carrier migration in lead-free bismuth-based halide perovskite for efficient selective oxidation of thioanisole under visible light. Chem. Eng. J. 2023, 451, 138927. [Google Scholar] [CrossRef]
  159. Zhang, Z.; Yang, Y.; Wang, Y.; Yang, L.; Li, Q.; Chen, L.; Xu, D. Revealing the A-Site Effect of Lead-Free A3Sb2Br9 Perovskite in Photocatalytic C(sp3)–H Bond Activation. Angew. Chem. Int. Ed. 2020, 59, 18136–18139. [Google Scholar] [CrossRef] [PubMed]
  160. Dai, Y.; Poidevin, C.; Ochoa-Hernández, C.; Auer, A.A.; Tüysüz, H. A Supported Bismuth Halide Perovskite Photocatalyst for Selective Aliphatic and Aromatic C–H Bond Activation. Angew. Chem. Int. Ed. 2020, 59, 5788–5796. [Google Scholar] [CrossRef]
  161. Shi, M.; Li, G.; Tian, W.; Jin, S.; Tao, X.; Jiang, Y.; Pidko, E.A.; Li, R.; Li, C. Understanding the Effect of Crystalline Structural Transformation for Lead-Free Inorganic Halide Perovskites. Adv. Mater. 2020, 32, 2002137. [Google Scholar] [CrossRef]
  162. Slavney, A.H.; Leppert, L.; Bartesaghi, D.; Gold-Parker, A.; Toney, M.F.; Savenije, T.J.; Neaton, J.B.; Karunadasa, H.I. Defect-Induced Band-Edge Reconstruction of a Bismuth-Halide Double Perovskite for Visible-Light Absorption. J. Am. Chem. Soc. 2017, 139, 5015–5018. [Google Scholar] [CrossRef] [PubMed]
  163. Du, K.-z.; Meng, W.; Wang, X.; Yan, Y.; Mitzi, D.B. Bandgap Engineering of Lead-Free Double Perovskite Cs2AgBiBr6 through Trivalent Metal Alloying. Angew. Chem. Int. Ed. 2017, 56, 8158–8162. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, G.; Zhang, Z.; Wu, C.; Zhang, Y.; Li, X.; Yu, W.; Yao, G.; Liu, S.; Shi, J.-j.; Liu, K.; et al. Extending Absorption of Cs2AgBiBr6 to Near-Infrared Region (≈1350 nm) with Intermediate Band. Adv. Funct. Mater. 2022, 32, 2109891. [Google Scholar] [CrossRef]
  165. Li, X.; Mai, H.; Cox, N.; Lu, J.; Wen, X.; Chen, D.; Caruso, R.A. Sb-Substituted Cs2AgBiBr6/g-C3N4 Composite for Photocatalytic C(sp3)–H Bond Activation in Toluene. Chem. Mater. 2023, 35, 3105–3114. [Google Scholar] [CrossRef]
  166. Song, J.; Zhang, C.; Zhang, H.; Dai, D.; Zhang, Q.; Wang, Z.; Zheng, Z.; Liu, Y.; Cheng, H.; Dai, Y.; et al. In situ growth of lead-free perovskite Cs2AgBiBr6 on a flexible ultrathin carbon nitride sheet for highly efficient photocatalytic benzylic C(sp3)–H bond activation. Chem. Eng. J. 2023, 453, 139748. [Google Scholar] [CrossRef]
  167. Zhang, Y.; Zhang, N.; Tang, Z.-R.; Xu, Y.-J. Transforming CdS into an efficient visible light photocatalyst for selective oxidation of saturated primary C–H bonds under ambient conditions. Chem. Sci. 2012, 3, 2812–2822. [Google Scholar] [CrossRef]
  168. Cao, F.-L.; Wang, J.-G.; Lv, F.-J.; Zhang, D.-Q.; Huo, Y.-N.; Li, G.-S.; Li, H.-X.; Zhu, J. Photocatalytic oxidation of toluene to benzaldehyde over anatase TiO2 hollow spheres with exposed {001} facets. Catal. Commun. 2011, 12, 946–950. [Google Scholar] [CrossRef]
  169. Liu, Y.; Chen, L.; Yuan, Q.; He, J.; Au, C.-T.; Yin, S.-F. A green and efficient photocatalytic route for the highly-selective oxidation of saturated alpha-carbon C–H bonds in aromatic alkanes over flower-like Bi2WO6. Chem. Commun. 2016, 52, 1274–1277. [Google Scholar] [CrossRef]
  170. Zhang, M.; Li, Z.; Xin, X.; Zhang, J.; Feng, Y.; Lv, H. Selective Valorization of 5-Hydroxymethylfurfural to 2,5-Diformylfuran Using Atmospheric O2 and MAPbBr3 Perovskite under Visible Light. ACS Catal. 2020, 10, 14793–14800. [Google Scholar] [CrossRef]
  171. Yin, L.; Wu, Y.; Bao, X.; Liu, X.; Dai, D.; Zhang, M.; Wang, Z.; Zheng, Z.; Liu, Y.; Cheng, H.; et al. Lead-Free Perovskite Cs2SnBr6/rGO Composite for Photocatalytic Selective Oxidation of 5-Hydroxymethylfurfural to 2,5-Diformylfuran. Chem.–Eur. J. 2023, 29, e202300999. [Google Scholar] [CrossRef]
  172. Rosa-Pardo, I.; Zhu, D.; Cortés-Villena, A.; Prato, M.; De Trizio, L.; Manna, L.; Galian, R.E.; Pérez-Prieto, J. The Dark Side of Lead-Free Metal Halide Nanocrystals: Substituent-Modulated Photocatalytic Activity in Benzyl Bromide Reduction. ACS Energy Lett. 2023, 8, 2789–2798. [Google Scholar] [CrossRef]
  173. Guo, Y.; Lou, Y.; Chen, J.; Zhao, Y. Lead-Free Cs2AgSbCl6 Double Perovskite Nanocrystals for Effective Visible-Light Photocatalytic C–C Coupling Reactions. ChemSusChem 2022, 15, e202102334. [Google Scholar] [CrossRef] [PubMed]
  174. Jagadeeswararao, M.; Sim, K.M.; Lee, S.; Kang, M.; An, S.; Nam, G.-H.; Sim, H.R.; Oleiki, E.; Lee, G.; Chung, D.S. Stoichiometric Engineering of Cs2AgBiBr6 for Photomultiplication-Type Photodetectors. Chem. Mater. 2023, 35, 3095–3104. [Google Scholar] [CrossRef]
  175. Jing, J.; Chen, Z.; Feng, C.; Sun, M.; Hou, J. Transforming g-C3N4 from amphoteric to n-type semiconductor: The important role of p/n type on photoelectrochemical cathodic protection. J. Alloys Compd. 2021, 851, 156820. [Google Scholar] [CrossRef]
  176. Luo, J.; Dong, G.; Zhu, Y.; Yang, Z.; Wang, C. Switching of semiconducting behavior from n-type to p-type induced high photocatalytic NO removal activity in g-C3N4. Appl. Catal. B Environ. 2017, 214, 46–56. [Google Scholar] [CrossRef]
  177. Bera, S.; Pradhan, N. Perovskite Nanocrystal Heterostructures: Synthesis, Optical Properties, and Applications. ACS Energy Lett. 2020, 5, 2858–2872. [Google Scholar] [CrossRef]
  178. Kipkorir, A.; DuBose, J.; Cho, J.; Kamat, P.V. CsPbBr3–CdS heterostructure: Stabilizing perovskite nanocrystals for photocatalysis. Chem. Sci. 2021, 12, 14815–14825. [Google Scholar] [CrossRef] [PubMed]
  179. Rosa-Pardo, I.; Ciccone, A.; Arenal, R.; Galian, R.E.; Pérez-Prieto, J. One-Pot Synthesis of Stable CsPbBr3@CsPb2Br5 Core–Shell Heteronanocrystals with Controlled Permeability to Halide Ions. Chem. Mater. 2023, 35, 7011–7019. [Google Scholar] [CrossRef]
  180. Wang, B.; Zhang, C.; Huang, S.; Li, Z.; Kong, L.; Jin, L.; Wang, J.; Wu, K.; Li, L. Postsynthesis Phase Transformation for CsPbBr3/Rb4PbBr6 Core/Shell Nanocrystals with Exceptional Photostability. ACS Appl. Mater. Interfaces 2018, 10, 23303–23310. [Google Scholar] [CrossRef]
  181. Ruan, L.J.; Tang, B.; Ma, Y. Improving the Stability of CsPbBr3 Nanocrystals in Ethanol by Capping with PbBr2-Adlayers. J. Phys. Chem. C 2019, 123, 11959–11967. [Google Scholar] [CrossRef]
  182. Cai, J.; Gu, K.; Zhu, Y.; Zhu, J.; Wang, Y.; Shen, J.; Trinchi, A.; Li, C.; Wei, G. Highly stable CsPbBr3@SiO2 nanocomposites prepared via confined condensation for use as a luminescent ink. Chem. Commun. 2018, 54, 8064–8067. [Google Scholar] [CrossRef]
  183. Li, Z.-J.; Hofman, E.; Li, J.; Davis, A.H.; Tung, C.-H.; Wu, L.-Z.; Zheng, W. Photoelectrochemically Active and Environmentally Stable CsPbBr3/TiO2 Core/Shell Nanocrystals. Adv. Funct. Mater. 2018, 28, 1704288. [Google Scholar] [CrossRef]
  184. Li, Z.; Kong, L.; Huang, S.; Li, L. Highly Luminescent and Ultrastable CsPbBr3 Perovskite Quantum Dots Incorporated into a Silica/Alumina Monolith. Angew. Chem. Int. Ed. 2017, 56, 8134–8138. [Google Scholar] [CrossRef] [PubMed]
  185. Liu, H.; Tan, Y.; Cao, M.; Hu, H.; Wu, L.; Yu, X.; Wang, L.; Sun, B.; Zhang, Q. Fabricating CsPbX3-Based Type I and Type II Heterostructures by Tuning the Halide Composition of Janus CsPbX3/ZrO2 Nanocrystals. ACS Nano 2019, 13, 5366–5374. [Google Scholar] [CrossRef] [PubMed]
  186. Huang, H.; Chen, B.; Wang, Z.; Hung, T.F.; Susha, A.S.; Zhong, H.; Rogach, A.L. Water resistant CsPbX3 nanocrystals coated with polyhedral oligomeric silsesquioxane and their use as solid state luminophores in all-perovskite white light-emitting devices. Chem. Sci. 2016, 7, 5699–5703. [Google Scholar] [CrossRef] [PubMed]
  187. Xiao, M.; Yuan, G.; Lu, Z.; Xia, J.; Li, D.; Chen, Y.; Zhang, Y.; Pei, F.; Chen, C.; Bai, Y.; et al. Engineering Amorphous–Crystallized Interface of ZrNx Barriers for Stable Inverted Perovskite Solar Cells. Adv. Mater. 2023, 35, 2301684. [Google Scholar] [CrossRef] [PubMed]
  188. Raja, S.N.; Bekenstein, Y.; Koc, M.A.; Fischer, S.; Zhang, D.; Lin, L.; Ritchie, R.O.; Yang, P.; Alivisatos, A.P. Encapsulation of Perovskite Nanocrystals into Macroscale Polymer Matrices: Enhanced Stability and Polarization. ACS Appl. Mater. Interfaces 2016, 8, 35523–35533. [Google Scholar] [CrossRef] [PubMed]
  189. Wu, H.; Wang, S.; Cao, F.; Zhou, J.; Wu, Q.; Wang, H.; Li, X.; Yin, L.; Yang, X. Ultrastable Inorganic Perovskite Nanocrystals Coated with a Thick Long-Chain Polymer for Efficient White Light-Emitting Diodes. Chem. Mater. 2019, 31, 1936–1940. [Google Scholar] [CrossRef]
  190. Pan, A.; Wang, J.; Jurow, M.J.; Jia, M.; Liu, Y.; Wu, Y.; Zhang, Y.; He, L.; Liu, Y. General Strategy for the Preparation of Stable Luminous Nanocomposite Inks Using Chemically Addressable CsPbX3 Peroskite Nanocrystals. Chem. Mater. 2018, 30, 2771–2780. [Google Scholar] [CrossRef]
  191. Wu, L.-Y.; Mu, Y.-F.; Guo, X.-X.; Zhang, W.; Zhang, Z.-M.; Zhang, M.; Lu, T.-B. Encapsulating Perovskite Quantum Dots in Iron-Based Metal–Organic Frameworks (MOFs) for Efficient Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 9491–9495. [Google Scholar] [CrossRef]
  192. Chen, S.; Yin, H.; Liu, P.; Wang, Y.; Zhao, H. Stabilization and Performance Enhancement Strategies for Halide Perovskite Photocatalysts. Adv. Mater. 2023, 35, 2203836. [Google Scholar] [CrossRef] [PubMed]
  193. Jiang, G.; Guhrenz, C.; Kirch, A.; Sonntag, L.; Bauer, C.; Fan, X.; Wang, J.; Reineke, S.; Gaponik, N.; Eychmüller, A. Highly Luminescent and Water-Resistant CsPbBr3–CsPb2Br5 Perovskite Nanocrystals Coordinated with Partially Hydrolyzed Poly(methyl methacrylate) and Polyethylenimine. ACS Nano 2019, 13, 10386–10396. [Google Scholar] [CrossRef] [PubMed]
  194. Song, S.; Lv, Y.; Cao, B.; Wang, W. Surface Modification Strategy Synthesized CsPbX3 Perovskite Quantum Dots with Excellent Stability and Optical Properties in Water. Adv. Funct. Mater. 2023, 33, 2300493. [Google Scholar] [CrossRef]
  195. Jagadeeswararao, M.; Vashishtha, P.; Hooper, T.J.N.; Kanwat, A.; Lim, J.W.M.; Vishwanath, S.K.; Yantara, N.; Park, T.; Sum, T.C.; Chung, D.S.; et al. One-Pot Synthesis and Structural Evolution of Colloidal Cesium Lead Halide–Lead Sulfide Heterostructure Nanocrystals for Optoelectronic Applications. J. Phys. Chem. Lett. 2021, 12, 9569–9578. [Google Scholar] [CrossRef] [PubMed]
  196. Das, R.; Patra, A.; Dutta, S.K.; Shyamal, S.; Pradhan, N. Facets-Directed Epitaxially Grown Lead Halide Perovskite-Sulfobromide Nanocrystal Heterostructures and Their Improved Photocatalytic Activity. J. Am. Chem. Soc. 2022, 144, 18629–18641. [Google Scholar] [CrossRef] [PubMed]
  197. Bera, S.; Patra, A.; Shyamal, S.; Nasipuri, D.; Pradhan, N. Lead Halide Perovskite Cube Coupled Star Nanocrystal Photocatalysts. ACS Energy Lett. 2022, 7, 3015–3023. [Google Scholar] [CrossRef]
  198. Rusch, P.; Toso, S.; Ivanov, Y.P.; Marras, S.; Divitini, G.; Manna, L. Nanocrystal Heterostructures Based On Halide Perovskites and Lead–Bismuth Chalcogenides. Chem. Mater. 2023, 35, 10684–10693. [Google Scholar] [CrossRef]
  199. Zhu, C.; Wang, C.; Zhang, P.; Ma, S.; Chen, Y.; Zhang, Y.; Yang, N.; Xiao, M.; Cheng, X.; Gao, Z.; et al. Topochemical assembly minimizes lattice heterogeneity in polycrystalline halide perovskites. Joule 2023, 7, 2361–2375. [Google Scholar] [CrossRef]
  200. Yang, Y.; Wang, D.; Li, Y.; Xia, J.; Wei, H.; Ding, C.; Hu, Y.; Wei, Y.; Li, H.; Liu, D.; et al. In Situ Room-Temperature Synthesis of All-Colloidal Quantum Dot CsPbBr3–PbS Heterostructures. ACS Photonics 2023, 10, 4305–4314. [Google Scholar] [CrossRef]
  201. Gaulding, E.A.; Chen, X.; Yang, Y.; Harvey, S.P.; To, B.; Kim, Y.-H.; Beard, M.C.; Sercel, P.C.; Luther, J.M. Embedding PbS Quantum Dots (QDs) in Pb-Halide Perovskite Matrices: QD Surface Chemistry and Antisolvent Effects on QD Dispersion and Confinement Properties. ACS Mater. Lett. 2020, 2, 1464–1472. [Google Scholar] [CrossRef]
  202. Yan, D.; Liu, M.; Li, Z.; Hou, B. Colloidal quantum dots and metal halide perovskite hybridization for solar cell stability and performance enhancement. J. Mater. Chem. A 2021, 9, 15522–15541. [Google Scholar] [CrossRef]
  203. Vashishtha, P.; Nutan, G.V.; Griffith, B.E.; Fang, Y.; Giovanni, D.; Jagadeeswararao, M.; Sum, T.C.; Mathews, N.; Mhaisalkar, S.G.; Hanna, J.V.; et al. Cesium Copper Iodide Tailored Nanoplates and Nanorods for Blue, Yellow, and White Emission. Chem. Mater. 2019, 31, 9003–9011. [Google Scholar] [CrossRef]
  204. Swarnkar, A.; Mir, W.J.; Chakraborty, R.; Jagadeeswararao, M.; Sheikh, T.; Nag, A. Are Chalcogenide Perovskites an Emerging Class of Semiconductors for Optoelectronic Properties and Solar Cell? Chem. Mater. 2019, 31, 565–575. [Google Scholar] [CrossRef]
Figure 2. (a) UV–visible absorption and PL spectra (dashed lines) of different-sized CsPbBr3 NCs, P2-P5; (b) formation of Csp3-Csp3 and Csp3-Nsp3 bonds shown in red colour and their respective yields in reaction 1 and 2.
Figure 2. (a) UV–visible absorption and PL spectra (dashed lines) of different-sized CsPbBr3 NCs, P2-P5; (b) formation of Csp3-Csp3 and Csp3-Nsp3 bonds shown in red colour and their respective yields in reaction 1 and 2.
Nanomaterials 14 00094 g002
Figure 3. (a) UV–vis absorption spectrum of various-sized Cs4ZnSb2Cl12 layered double perovskite (LDP) NCs; (b) scheme of the energy levels of the conduction and valence band edges for bulk and different-sized Cs4ZnSb2Cl12 NCs and mechanism of photocatalytic toluene oxidation to benzaldehyde formation; (c) time-dependent photo catalytic activity of benzaldehyde formation using bulk and NCs of Cs4ZnSb2Cl12. Reproduced with permission. Copyright: 2023, American Chemical Society.
Figure 3. (a) UV–vis absorption spectrum of various-sized Cs4ZnSb2Cl12 layered double perovskite (LDP) NCs; (b) scheme of the energy levels of the conduction and valence band edges for bulk and different-sized Cs4ZnSb2Cl12 NCs and mechanism of photocatalytic toluene oxidation to benzaldehyde formation; (c) time-dependent photo catalytic activity of benzaldehyde formation using bulk and NCs of Cs4ZnSb2Cl12. Reproduced with permission. Copyright: 2023, American Chemical Society.
Nanomaterials 14 00094 g003
Figure 4. C–C bond product formation (shown in red colour bonds in the products 3c and 3d) and respective chemical yields.
Figure 4. C–C bond product formation (shown in red colour bonds in the products 3c and 3d) and respective chemical yields.
Nanomaterials 14 00094 g004
Figure 5. (a) Tunability of photoluminescence (PL) spectra of CsPbBr3 NCs using anion exchange with trimethylsilyl chloride or iodide and their photographs under UV illumination; (b) band edges of APbClxBryI3−x−y; (c) excited state potential (E*) range of APbClxBryI3−x−y comparing with noble-transition metal photocatalysts; (d) absorption (dashed lines) and PL spectra of Zn- and Cd-doped CsPbBr3 NCs. Reproduced with permission [45]. Copyright: 2017, American Chemical Society; (e) absorption and PL spectra (dashed lines) of CsSnX3 NCs. Reproduced with permission [26]. Copyright: 2016, American Chemical Society; (f) C–O bond formation with perovskite band-tuning.
Figure 5. (a) Tunability of photoluminescence (PL) spectra of CsPbBr3 NCs using anion exchange with trimethylsilyl chloride or iodide and their photographs under UV illumination; (b) band edges of APbClxBryI3−x−y; (c) excited state potential (E*) range of APbClxBryI3−x−y comparing with noble-transition metal photocatalysts; (d) absorption (dashed lines) and PL spectra of Zn- and Cd-doped CsPbBr3 NCs. Reproduced with permission [45]. Copyright: 2017, American Chemical Society; (e) absorption and PL spectra (dashed lines) of CsSnX3 NCs. Reproduced with permission [26]. Copyright: 2016, American Chemical Society; (f) C–O bond formation with perovskite band-tuning.
Nanomaterials 14 00094 g005
Figure 6. Schematic illustration of the band alignments of several possible heterojunctions. (a) Schottky heterojunction; (b) type-II heterojunction; (c) all-solid-z-scheme hetero junction; (d). direct Z-scheme heterojunction; (e) the relative band positions and Fermi level of semiconductor (SC) I and SC II before and after contact; (f,g) After illumination with light, the charge migration follows the Z-scheme and type-II pathways. VB: valence band; CB: conduction band; Ef: Fermi level (dashed line); IEF: internal electric field. Reproduced with permission [58]. Copyright: John Wiley and Sons.
Figure 6. Schematic illustration of the band alignments of several possible heterojunctions. (a) Schottky heterojunction; (b) type-II heterojunction; (c) all-solid-z-scheme hetero junction; (d). direct Z-scheme heterojunction; (e) the relative band positions and Fermi level of semiconductor (SC) I and SC II before and after contact; (f,g) After illumination with light, the charge migration follows the Z-scheme and type-II pathways. VB: valence band; CB: conduction band; Ef: Fermi level (dashed line); IEF: internal electric field. Reproduced with permission [58]. Copyright: John Wiley and Sons.
Nanomaterials 14 00094 g006
Figure 7. (a) Photocatalytic reductive dehalogenation, sp3-C coupling, and α-alkylation of aldehydes; (b) proposed mechanism for perovskite catalyzed dehalogenation, sp3 carbon coupling, and α-alkylation; (c) TEMPO trapped experiment for radical intermediate validation. Reproduced with permission [65]. Copyright: 2019, American Chemical Society.
Figure 7. (a) Photocatalytic reductive dehalogenation, sp3-C coupling, and α-alkylation of aldehydes; (b) proposed mechanism for perovskite catalyzed dehalogenation, sp3 carbon coupling, and α-alkylation; (c) TEMPO trapped experiment for radical intermediate validation. Reproduced with permission [65]. Copyright: 2019, American Chemical Society.
Nanomaterials 14 00094 g007
Figure 8. (a) Normalized transient absorption (TA) kinetics probed at the centre of the NCs exciton bleach spectrum for A–D, red-dashed traces are fits to kinetics; (b) normalized TA kinetics and fits for A and E where A, B, C, D, E represent pure CsPbBr3 NCs, CsPbBr3 NCs + 2-bromoacetophenone, CsPbBr3 NCs + octanal, CsPbBr3 NCs + octanal + dicyclohexylamine, CsPbBr3 NCs + dicyclohexylamine, respectively; (c) electrochemical potential and energy level for reactants and CsPbBr3 NCs. Reproduced with permission. Copyright: 2020, American Chemical Society.
Figure 8. (a) Normalized transient absorption (TA) kinetics probed at the centre of the NCs exciton bleach spectrum for A–D, red-dashed traces are fits to kinetics; (b) normalized TA kinetics and fits for A and E where A, B, C, D, E represent pure CsPbBr3 NCs, CsPbBr3 NCs + 2-bromoacetophenone, CsPbBr3 NCs + octanal, CsPbBr3 NCs + octanal + dicyclohexylamine, CsPbBr3 NCs + dicyclohexylamine, respectively; (c) electrochemical potential and energy level for reactants and CsPbBr3 NCs. Reproduced with permission. Copyright: 2020, American Chemical Society.
Nanomaterials 14 00094 g008
Figure 9. (a) Photocatalytic aryl bromide coupling reaction; (b) schematic representation of the cooperative action between the NC surface and the capping for the catalysed coupling reaction; (c) the process occurring after excitation of the CsPbBr3 NCs by vis-lamp in the presence of benzyl halide as electron acceptor and DIPEA as electron donor within the capping.
Figure 9. (a) Photocatalytic aryl bromide coupling reaction; (b) schematic representation of the cooperative action between the NC surface and the capping for the catalysed coupling reaction; (c) the process occurring after excitation of the CsPbBr3 NCs by vis-lamp in the presence of benzyl halide as electron acceptor and DIPEA as electron donor within the capping.
Nanomaterials 14 00094 g009
Figure 10. (a) Photocatalytic stereo-selective C–C oxidative coupling reaction of 2-(3,4-dimethoxyphenyl)-3-oxobutanenitrile; (b) proposed reaction scheme for the stereoselective dimerization of α-keto nitriles photocatalyzed by zwitterion capped CsPbBr3 NCs. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Figure 10. (a) Photocatalytic stereo-selective C–C oxidative coupling reaction of 2-(3,4-dimethoxyphenyl)-3-oxobutanenitrile; (b) proposed reaction scheme for the stereoselective dimerization of α-keto nitriles photocatalyzed by zwitterion capped CsPbBr3 NCs. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Nanomaterials 14 00094 g010
Figure 11. (a) Aliphatic C–Br bond activation of CBrX3 (X = Br, Cl) using the CsPbBr3 NCs. Reproduced with permission. Copyright: 2023, American Chemical Society; (b) chemodivergent synthesis of 3,8-dibromo-1-methyl-4-phenyl-1-azaspiro[4.5]deca-3,6,9-trien-2-on or 3-bromo-1-methyl-4-phenyl-1-azaspiro[4.5] deca-3,6,9-trien-2,8-dione or 3-bromo-6-(tert-butyl)-1-methyl-4-phenylquinolin-2(1H)-one using CsPbBr3 NCs photocatalyst under visible light irradiation. Reproduced with permission. Copyright: 2023, American Chemical Society.
Figure 11. (a) Aliphatic C–Br bond activation of CBrX3 (X = Br, Cl) using the CsPbBr3 NCs. Reproduced with permission. Copyright: 2023, American Chemical Society; (b) chemodivergent synthesis of 3,8-dibromo-1-methyl-4-phenyl-1-azaspiro[4.5]deca-3,6,9-trien-2-on or 3-bromo-1-methyl-4-phenyl-1-azaspiro[4.5] deca-3,6,9-trien-2,8-dione or 3-bromo-6-(tert-butyl)-1-methyl-4-phenylquinolin-2(1H)-one using CsPbBr3 NCs photocatalyst under visible light irradiation. Reproduced with permission. Copyright: 2023, American Chemical Society.
Nanomaterials 14 00094 g011
Figure 12. (a) Photocatalytic aminomethylation of imidazo-fused heterocycles; (b) plausible mechanism of aminomethylation of imidazo-fused heterocycles using CsPbBr3 NCs under the irradiation of visible light. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Figure 12. (a) Photocatalytic aminomethylation of imidazo-fused heterocycles; (b) plausible mechanism of aminomethylation of imidazo-fused heterocycles using CsPbBr3 NCs under the irradiation of visible light. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Nanomaterials 14 00094 g012
Figure 13. Schematic representation of the sp2carbon coupling reaction of iodobenzene with phenylboronic acid using Pd/CsPbBr3 Schottky junction photocatalyst. Reproduced with permission. Copyright: 2022, American Chemical Society.
Figure 13. Schematic representation of the sp2carbon coupling reaction of iodobenzene with phenylboronic acid using Pd/CsPbBr3 Schottky junction photocatalyst. Reproduced with permission. Copyright: 2022, American Chemical Society.
Nanomaterials 14 00094 g013
Figure 14. (a) Crystal structures of (HDA)2PbI4 and (HDA)2SnI4; (b) representative decarboxylation and dehydrogenation reactions catalyzed by the (HDA)2PbI4 and (HDA)2SnI4 perovskites; (c) suggested mechanism for the decarboxylation (left) and dehydrogenation reactions (right). [P] represents the perovskite. Reproduced with permission. Copyright: 2019, John Wiley and Sons.
Figure 14. (a) Crystal structures of (HDA)2PbI4 and (HDA)2SnI4; (b) representative decarboxylation and dehydrogenation reactions catalyzed by the (HDA)2PbI4 and (HDA)2SnI4 perovskites; (c) suggested mechanism for the decarboxylation (left) and dehydrogenation reactions (right). [P] represents the perovskite. Reproduced with permission. Copyright: 2019, John Wiley and Sons.
Nanomaterials 14 00094 g014
Figure 15. (a) Photocatalytic NN heterocyclization reaction; (b) photocatalytic sequential electron transfer approach for di-radical path N-heterocyclization using Cu-doped CsPbBr3 NCs. Red circle represents the Cu(I)-doping in the perovskite crystal structure. Reproduced with permission. Copyright: 2021, American Chemical Society.
Figure 15. (a) Photocatalytic NN heterocyclization reaction; (b) photocatalytic sequential electron transfer approach for di-radical path N-heterocyclization using Cu-doped CsPbBr3 NCs. Red circle represents the Cu(I)-doping in the perovskite crystal structure. Reproduced with permission. Copyright: 2021, American Chemical Society.
Nanomaterials 14 00094 g015
Figure 16. (a,b) Illustration of the proposed mechanism for photocatalytic polymerization of TerEDOT over CsPbI3 NCs under visible light irradiation; (c) TEM image of CsPbI3 NCs encapsulated by PEDOT and CsPbI3 NCs. Reproduced with permission. Copyright: 2017, American Chemical Society.
Figure 16. (a,b) Illustration of the proposed mechanism for photocatalytic polymerization of TerEDOT over CsPbI3 NCs under visible light irradiation; (c) TEM image of CsPbI3 NCs encapsulated by PEDOT and CsPbI3 NCs. Reproduced with permission. Copyright: 2017, American Chemical Society.
Nanomaterials 14 00094 g016
Figure 17. (a) Thiol-coupling reaction; (b) cross-dehydrogenative coupling reaction between tertiary amines and phosphite esters.
Figure 17. (a) Thiol-coupling reaction; (b) cross-dehydrogenative coupling reaction between tertiary amines and phosphite esters.
Nanomaterials 14 00094 g017
Figure 18. Chiral perovskite NCs. NCs formulated with L2(APbBr3)n−1PbBr4. Right: proposed structure of NCs used in this study where n = ∞. Reproduced with permission. Copyright: 2023, American Chemical Society.
Figure 18. Chiral perovskite NCs. NCs formulated with L2(APbBr3)n−1PbBr4. Right: proposed structure of NCs used in this study where n = ∞. Reproduced with permission. Copyright: 2023, American Chemical Society.
Nanomaterials 14 00094 g018
Figure 19. (a) Circular dichroism (CD) spectrum of R-NC-3 and S-NC-3; (b) illustration of N–C axially chiral atroposelective N-heterocyclization using L2(APbBr3)n−1PbBr4 photocatalyst; (c) proposed mechanism for N-arylindole; (d) DFT studies, binding energy discrepancies between the chiral surface and substrate 1a (up), 1a+ (after hole transfer). Reproduced with permission. Copyright: 2023, American Chemical Society.
Figure 19. (a) Circular dichroism (CD) spectrum of R-NC-3 and S-NC-3; (b) illustration of N–C axially chiral atroposelective N-heterocyclization using L2(APbBr3)n−1PbBr4 photocatalyst; (c) proposed mechanism for N-arylindole; (d) DFT studies, binding energy discrepancies between the chiral surface and substrate 1a (up), 1a+ (after hole transfer). Reproduced with permission. Copyright: 2023, American Chemical Society.
Nanomaterials 14 00094 g019
Figure 20. (Left) Sensitization of naphthalene triplets (Nap3*) using dyes: light absorption populates the first singlet excited state (S1), which is followed by intersystem crossing (ISC) to generate the first triplet states (T1). These T1 states can subsequently transfer their energy to Nap via Dexter-type triplet energy transfer (TET) mechanism. Considering that the energy of Nap3* is ∼2.6 eV and that there is an estimated energy loss of ∼0.5 eV during the ISC process, it implies that the energy of S1 of the dye must be at least ∼3.1 eV, indicating that UV photons are required to drive the sensitization process; (Right) Sensitization of naphthalene triplets (Nap3*) using semiconductor CsPbBr3 NCs: light absorption leads to the generation of band edge excitonic states (X) within the NCs. These excitonic states can directly undergo TET to produce Nap3*. This means that visible photons are capable of efficiently driving the sensitization process.
Figure 20. (Left) Sensitization of naphthalene triplets (Nap3*) using dyes: light absorption populates the first singlet excited state (S1), which is followed by intersystem crossing (ISC) to generate the first triplet states (T1). These T1 states can subsequently transfer their energy to Nap via Dexter-type triplet energy transfer (TET) mechanism. Considering that the energy of Nap3* is ∼2.6 eV and that there is an estimated energy loss of ∼0.5 eV during the ISC process, it implies that the energy of S1 of the dye must be at least ∼3.1 eV, indicating that UV photons are required to drive the sensitization process; (Right) Sensitization of naphthalene triplets (Nap3*) using semiconductor CsPbBr3 NCs: light absorption leads to the generation of band edge excitonic states (X) within the NCs. These excitonic states can directly undergo TET to produce Nap3*. This means that visible photons are capable of efficiently driving the sensitization process.
Nanomaterials 14 00094 g020
Figure 21. (a) Absorption spectra of CsPbBr3 NCs (grey solid lines) and NC–PCA complexes (coloured solid lines) with different NC sizes (edge length L indicated) dispersed in hexane. The variation between them is due to PCA absorption. The inset illustrates the schematic structures of NCs and PCAs; (b) PL spectra of CsPbBr3 NCs (grey solid lines) and NC–PCA complexes (coloured solid lines) excited at 420 nm. The inset illustrates the schematic energy level alignment between NCs and PCAs; (c) TA kinetics probed at the exciton bleaching (XB) centre (∼474 nm; solid symbols) and at the PCA triplet absorption (∼423 nm; open symbols) of L = 4.9 nm NCs (squares) and NC–PCA complexes (circles). The grey solid lines represent their fits using stretched-exponential functions; (d) Similar plot as in c for L = 9.4 nm NC–PCA complexes; (e) Normalized TA kinetics probed at the XB feature of NC–PCA complexes with varying NC sizes (coloured symbols). The grey solid lines represent their fits; (f) size-dependent TET rate (kTET) plotted as a function of carrier probability density at the NC surface (|Ψs|2) for different NC sizes (coloured symbols). The grey solid line depicts a linear fit. Reproduced with permission. Copyright: 2019, American Chemical Society.
Figure 21. (a) Absorption spectra of CsPbBr3 NCs (grey solid lines) and NC–PCA complexes (coloured solid lines) with different NC sizes (edge length L indicated) dispersed in hexane. The variation between them is due to PCA absorption. The inset illustrates the schematic structures of NCs and PCAs; (b) PL spectra of CsPbBr3 NCs (grey solid lines) and NC–PCA complexes (coloured solid lines) excited at 420 nm. The inset illustrates the schematic energy level alignment between NCs and PCAs; (c) TA kinetics probed at the exciton bleaching (XB) centre (∼474 nm; solid symbols) and at the PCA triplet absorption (∼423 nm; open symbols) of L = 4.9 nm NCs (squares) and NC–PCA complexes (circles). The grey solid lines represent their fits using stretched-exponential functions; (d) Similar plot as in c for L = 9.4 nm NC–PCA complexes; (e) Normalized TA kinetics probed at the XB feature of NC–PCA complexes with varying NC sizes (coloured symbols). The grey solid lines represent their fits; (f) size-dependent TET rate (kTET) plotted as a function of carrier probability density at the NC surface (|Ψs|2) for different NC sizes (coloured symbols). The grey solid line depicts a linear fit. Reproduced with permission. Copyright: 2019, American Chemical Society.
Nanomaterials 14 00094 g021
Figure 22. (a) Photoisomerization of surface tethered stilbene; (b) the 9-phenanthrene carboxylic acid (PTA) ligands are anchored on CsPbBr3 NCs surfaces, whereas the stilbene molecules are freely diffusing in solution; (c) photoisomerization of diarylethene; (d) light-modulated conversion yields under 2.7 W 450 nm illumination, the blue shaded areas correspond with light-on; (e) absorption (solid lines) and PL (dashed lines) spectra of CsPbBr3 NCs with (green) and without (blue) surface-anchored PTA ligands; (f) nanosecond-TA kinetics probed for NC-sensitized 3PTA* with (green open circles) and without (blue open squares) trans-stilbene in the solution. Reproduced with permission. Copyright: 2022, John Wiley and Sons.
Figure 22. (a) Photoisomerization of surface tethered stilbene; (b) the 9-phenanthrene carboxylic acid (PTA) ligands are anchored on CsPbBr3 NCs surfaces, whereas the stilbene molecules are freely diffusing in solution; (c) photoisomerization of diarylethene; (d) light-modulated conversion yields under 2.7 W 450 nm illumination, the blue shaded areas correspond with light-on; (e) absorption (solid lines) and PL (dashed lines) spectra of CsPbBr3 NCs with (green) and without (blue) surface-anchored PTA ligands; (f) nanosecond-TA kinetics probed for NC-sensitized 3PTA* with (green open circles) and without (blue open squares) trans-stilbene in the solution. Reproduced with permission. Copyright: 2022, John Wiley and Sons.
Nanomaterials 14 00094 g022
Figure 23. (a) Molecule- and perovskite-induced photocatalysis for TET based 2 + 2 cycloaddition. HH: head-to-head, HT: head-to-tail; star represents the excitation of the molecular photocatalyst; (b) proposed mechanism for the intermolecular 2 + 2 cycloaddition facilitated by CsPbBr3 NCs through TET induced by visible light. Reproduced with permission. Copyright: 2022, American Chemical Society.
Figure 23. (a) Molecule- and perovskite-induced photocatalysis for TET based 2 + 2 cycloaddition. HH: head-to-head, HT: head-to-tail; star represents the excitation of the molecular photocatalyst; (b) proposed mechanism for the intermolecular 2 + 2 cycloaddition facilitated by CsPbBr3 NCs through TET induced by visible light. Reproduced with permission. Copyright: 2022, American Chemical Society.
Nanomaterials 14 00094 g023
Figure 24. (a) Investigation of the photocatalytic oxidation of benzyl alcohol using pure FAPbBr3, TiO2, and a series of FAPbBr3/TiO2 hybrids; (b) analysis of PL decay spectra of the prepared FAPbBr3 and 15% FAPbBr3/TiO2 and their multiexponential fitting parameters; (c) experimentally supported mechanistic energy diagram of photoinduced charge transfer in FAPbBr3 and FAPbBr3/TiO2 hybrid; (d) schematic of the proposed selective photocatalytic oxidation of benzyl alcohol to benzaldehyde using the FAPbBr3/TiO2 hybrid. Reproduced with permission. Copyright: 2022, American Chemical Society.
Figure 24. (a) Investigation of the photocatalytic oxidation of benzyl alcohol using pure FAPbBr3, TiO2, and a series of FAPbBr3/TiO2 hybrids; (b) analysis of PL decay spectra of the prepared FAPbBr3 and 15% FAPbBr3/TiO2 and their multiexponential fitting parameters; (c) experimentally supported mechanistic energy diagram of photoinduced charge transfer in FAPbBr3 and FAPbBr3/TiO2 hybrid; (d) schematic of the proposed selective photocatalytic oxidation of benzyl alcohol to benzaldehyde using the FAPbBr3/TiO2 hybrid. Reproduced with permission. Copyright: 2022, American Chemical Society.
Nanomaterials 14 00094 g024
Figure 25. (a) TEM image of CsPbBr3/GO-Pt composite; (b) schematic illustration of photocatalytic H2 production integrated with benzaldehyde synthesis. Reproduced with permission. Copyright: 2022, Frontiers Media S.A.
Figure 25. (a) TEM image of CsPbBr3/GO-Pt composite; (b) schematic illustration of photocatalytic H2 production integrated with benzaldehyde synthesis. Reproduced with permission. Copyright: 2022, Frontiers Media S.A.
Nanomaterials 14 00094 g025
Figure 26. (a) HRTEM images of NiOx/FAPbBr3/TiO2 composite; (b) UV–vis DRS spectra with the corresponding bandgaps shown in the inset; (c) steady-state PL spectra and (d) PL decay spectra of the prepared FAPbBr3/TiO2 and 5% NiOx/FAPbBr3/TiO2; (e) experimentally supported mechanistic energy diagram of photoinduced charge transfer in NiOx/FAPbBr3/TiO2; (f) Investigation of the photocatalytic oxidation of toluene using pure FAPbBr3, FAPbBr3/TiO2, and a series of NiOx/FAPbBr3/TiO2 composites. Reproduced with permission. Copyright: 2019, American Chemical Society.
Figure 26. (a) HRTEM images of NiOx/FAPbBr3/TiO2 composite; (b) UV–vis DRS spectra with the corresponding bandgaps shown in the inset; (c) steady-state PL spectra and (d) PL decay spectra of the prepared FAPbBr3/TiO2 and 5% NiOx/FAPbBr3/TiO2; (e) experimentally supported mechanistic energy diagram of photoinduced charge transfer in NiOx/FAPbBr3/TiO2; (f) Investigation of the photocatalytic oxidation of toluene using pure FAPbBr3, FAPbBr3/TiO2, and a series of NiOx/FAPbBr3/TiO2 composites. Reproduced with permission. Copyright: 2019, American Chemical Society.
Nanomaterials 14 00094 g026
Figure 27. (a) Cross-sectional SEM of CsPbBr3-based solar photocatalyst cell; (b) selective photocatalytic oxidation of benzyl alcohol to benzaldehyde over films of CsPbBr3 and NiOx/CsPbBr3/TiO2 photocatalysts, where NiOx/CsPbBr3/TiO2-n (n = 1, 2, 3, 4, and 5) denotes various concentration of CsPbBr3 precursor solution: 0.05 M, 0.15 M, 0.25 M, 0.35 M, and 0.45 M, respectively, red color stars represents the selectivity of the photocatalysts; (c) assessment of the recycling test of NiOx/CsPbBr3/TiO2 solar photocatalyst cell and pure CsPbBr3 thin film photocatalyst. Reproduced with permission. Copyright: 2022, Elsevier.
Figure 27. (a) Cross-sectional SEM of CsPbBr3-based solar photocatalyst cell; (b) selective photocatalytic oxidation of benzyl alcohol to benzaldehyde over films of CsPbBr3 and NiOx/CsPbBr3/TiO2 photocatalysts, where NiOx/CsPbBr3/TiO2-n (n = 1, 2, 3, 4, and 5) denotes various concentration of CsPbBr3 precursor solution: 0.05 M, 0.15 M, 0.25 M, 0.35 M, and 0.45 M, respectively, red color stars represents the selectivity of the photocatalysts; (c) assessment of the recycling test of NiOx/CsPbBr3/TiO2 solar photocatalyst cell and pure CsPbBr3 thin film photocatalyst. Reproduced with permission. Copyright: 2022, Elsevier.
Nanomaterials 14 00094 g027
Figure 28. Hetero Diels–Alder (HDA) reaction of 1,3-cyclohexadiene and 1O2 using heterostructures g-C3N4 in combination with PEA2MX4.
Figure 28. Hetero Diels–Alder (HDA) reaction of 1,3-cyclohexadiene and 1O2 using heterostructures g-C3N4 in combination with PEA2MX4.
Nanomaterials 14 00094 g028
Figure 29. Proposed photocatalytic cycle for epoxide alcoholysis reaction using Cs3Bi2Br9 with proper Lewis acid sites on surface. Reproduced with permission. Copyright: 2019, John Wiley and Sons.
Figure 29. Proposed photocatalytic cycle for epoxide alcoholysis reaction using Cs3Bi2Br9 with proper Lewis acid sites on surface. Reproduced with permission. Copyright: 2019, John Wiley and Sons.
Nanomaterials 14 00094 g029
Figure 30. (a) The synthesis procedure for Cs3Bi2Br9 platelet crystals; (b) X-ray diffraction patterns of the Cs3Bi2Br9 microcrystals; (c) scanning electron microscopy (SEM) image of the irregular crystals grown without the addition of EA; (cg) SEM images displaying platelet-shaped crystals grown in the presence of EA with different perovskite concentrations of 10, 8, 6, and 4 g L−1, respectively. Reproduced with permission. Copyright: 2021, John Wiley and Sons.
Figure 30. (a) The synthesis procedure for Cs3Bi2Br9 platelet crystals; (b) X-ray diffraction patterns of the Cs3Bi2Br9 microcrystals; (c) scanning electron microscopy (SEM) image of the irregular crystals grown without the addition of EA; (cg) SEM images displaying platelet-shaped crystals grown in the presence of EA with different perovskite concentrations of 10, 8, 6, and 4 g L−1, respectively. Reproduced with permission. Copyright: 2021, John Wiley and Sons.
Nanomaterials 14 00094 g030
Figure 31. (a) HRTEM of 12.4% Cs3Bi2Br9/TiO2; (b) band structures of TiO2, Cs3Bi2Br9 and photoinduced charge carrier transfer under illumination. VB is valence band, and CB is conduction band. Substrate * represents the hole transfer to the substrate; (c) photocatalytic selective oxidation of benzyl alcohol over X% Cs3Bi2Br9/TiO2; (d) recycle test of photocatalytic oxidation of benzyl alcohol over 12.4% Cs3Bi2Br9/TiO2. Reproduced with permission. Copyright: 2022, Elsevier.
Figure 31. (a) HRTEM of 12.4% Cs3Bi2Br9/TiO2; (b) band structures of TiO2, Cs3Bi2Br9 and photoinduced charge carrier transfer under illumination. VB is valence band, and CB is conduction band. Substrate * represents the hole transfer to the substrate; (c) photocatalytic selective oxidation of benzyl alcohol over X% Cs3Bi2Br9/TiO2; (d) recycle test of photocatalytic oxidation of benzyl alcohol over 12.4% Cs3Bi2Br9/TiO2. Reproduced with permission. Copyright: 2022, Elsevier.
Nanomaterials 14 00094 g031
Figure 32. (a) High-resolution-TEM image Cs3Bi2Br9/MXene-7.5; (b) bandgap profile, (c) PL spectra (excitation at 380 nm); (d) transient photocurrent spectra and (e) electrochemical impedance spectroscopy (EIS) Nyquist plots of Cs3Bi2Br9 NCs, Cs3Bi2Br9/MXene-3, Cs3Bi2Br9/MXene-5, Cs3Bi2Br9/MXene-7.5, Cs3Bi2Br9/MXene-10, and Ti3C2Tx; (f) Schematic representation of the separation of photoinduced carriers of the Cs3Bi2Br9/MXene heterostructure and the proposed mechanism for the photocatalytic reaction of alkane. Reproduced with permission. Copyright: 2021, American Chemical Society.
Figure 32. (a) High-resolution-TEM image Cs3Bi2Br9/MXene-7.5; (b) bandgap profile, (c) PL spectra (excitation at 380 nm); (d) transient photocurrent spectra and (e) electrochemical impedance spectroscopy (EIS) Nyquist plots of Cs3Bi2Br9 NCs, Cs3Bi2Br9/MXene-3, Cs3Bi2Br9/MXene-5, Cs3Bi2Br9/MXene-7.5, Cs3Bi2Br9/MXene-10, and Ti3C2Tx; (f) Schematic representation of the separation of photoinduced carriers of the Cs3Bi2Br9/MXene heterostructure and the proposed mechanism for the photocatalytic reaction of alkane. Reproduced with permission. Copyright: 2021, American Chemical Society.
Nanomaterials 14 00094 g032
Figure 33. (a) TEM image of Cs3Bi2Br9/CdS hybrid; (b) band alignment of Cs3Bi2Br9 and CdS, and the proposed transfer pathways for photogenerated carriers and redox reactions under visible light (λ > 420 nm); (c) Photocatalytic activities of toluene oxidation using Cs3Bi2Br9/CdS, Cs3Bi2Br9, and CdS as photocatalysts. Reproduced with permission. Copyright: 2023, Elsevier.
Figure 33. (a) TEM image of Cs3Bi2Br9/CdS hybrid; (b) band alignment of Cs3Bi2Br9 and CdS, and the proposed transfer pathways for photogenerated carriers and redox reactions under visible light (λ > 420 nm); (c) Photocatalytic activities of toluene oxidation using Cs3Bi2Br9/CdS, Cs3Bi2Br9, and CdS as photocatalysts. Reproduced with permission. Copyright: 2023, Elsevier.
Nanomaterials 14 00094 g033
Figure 34. (a) TEM images of Cs3Sb2Br9/g-C3N4, inset shows the high-resolution image of single particle of the NC highlighted in the TEM image; (b) UV-visible absorption spectra of g-C3N4, Cs3Sb2Br9, and Cs3Sb2Br9/g-C3N4; (c) schematic illustration of the band position of Cs3Sb2Br9 and g-C3N4 and the proposed transfer pathways for photogenerated carriers; (d) photocatalytic oxidation of toluene conversion rate over diverse photocatalysts. Reproduced with permission. Copyright: 2021, John Wiley and Sons.
Figure 34. (a) TEM images of Cs3Sb2Br9/g-C3N4, inset shows the high-resolution image of single particle of the NC highlighted in the TEM image; (b) UV-visible absorption spectra of g-C3N4, Cs3Sb2Br9, and Cs3Sb2Br9/g-C3N4; (c) schematic illustration of the band position of Cs3Sb2Br9 and g-C3N4 and the proposed transfer pathways for photogenerated carriers; (d) photocatalytic oxidation of toluene conversion rate over diverse photocatalysts. Reproduced with permission. Copyright: 2021, John Wiley and Sons.
Nanomaterials 14 00094 g034
Figure 35. (a) The formation schematic diagram of Co0.97Bi1.03O2CO3/Cs3Bi2Br9 heterojunction; (b) TEM; (c) HRTEM images, box 1 and 2 represent ridge-like tiny nanosheets of and Cs3Bi2Br9 and Co0.97Bi1.03O2CO3 nanosheets, respectively; (d) the lattice match of Co0.97Bi1.03O2CO3/Cs3Bi2Br9 hybrid; Reproduced with permission. Copyright: 2023, Elsevier.
Figure 35. (a) The formation schematic diagram of Co0.97Bi1.03O2CO3/Cs3Bi2Br9 heterojunction; (b) TEM; (c) HRTEM images, box 1 and 2 represent ridge-like tiny nanosheets of and Cs3Bi2Br9 and Co0.97Bi1.03O2CO3 nanosheets, respectively; (d) the lattice match of Co0.97Bi1.03O2CO3/Cs3Bi2Br9 hybrid; Reproduced with permission. Copyright: 2023, Elsevier.
Nanomaterials 14 00094 g035
Figure 36. The photoinduced carriers transfer process of (a) in situ grown Co0.97Bi1.03O2CO3/Cs3Bi2Br9 heterojunction; (b) physically mixed Co0.97Bi1.03O2CO3/Cs3Bi2Br9 sample. Reproduced with permission. Copyright: 2023, Elsevier.
Figure 36. The photoinduced carriers transfer process of (a) in situ grown Co0.97Bi1.03O2CO3/Cs3Bi2Br9 heterojunction; (b) physically mixed Co0.97Bi1.03O2CO3/Cs3Bi2Br9 sample. Reproduced with permission. Copyright: 2023, Elsevier.
Nanomaterials 14 00094 g036
Figure 37. (a) SEM and (b) TEM images of d-BiOBr. (c) HRTEM images of CBB(1.7)/d-BiOBr. (d) Raman spectra of d-BiOBr, CBB, and CBB(1.7)/d-BiOBr; (e) represents the band alignment of CBB and d-BiOBr before and after contact; (f) SPV spectra; (g) comparison of photocatalytic toluene oxidation performance over d-BiOBr, CBB, and the CBB/d-BiOBr composite. Reproduced with permission. Copyright: 2022, American Chemical Society.
Figure 37. (a) SEM and (b) TEM images of d-BiOBr. (c) HRTEM images of CBB(1.7)/d-BiOBr. (d) Raman spectra of d-BiOBr, CBB, and CBB(1.7)/d-BiOBr; (e) represents the band alignment of CBB and d-BiOBr before and after contact; (f) SPV spectra; (g) comparison of photocatalytic toluene oxidation performance over d-BiOBr, CBB, and the CBB/d-BiOBr composite. Reproduced with permission. Copyright: 2022, American Chemical Society.
Nanomaterials 14 00094 g037
Figure 38. SEM image of (a) Cs3Bi2Br9, (b) Cs3Bi1.8Sb0.2Br9, (c) Cs3Bi1.5Sb0.5Br9, (d) Cs3Bi0.7Sb1.3Br9, (e) Cs3Sb2Br9, inset in (ac) show the magnified view of the SEM images; (f). Transient photocurrent response; (g) Photocatalytic performances of the prepared samples after irradiation of 3 h; (h) Bader electrons of upper layer minus that of lower layer over different models. Cs12Bi8Br36 is used as a model of Cs3Bi2Br9, Cs12Bi8Br36 were substituted by 0, 1, 2, 5, and 8 Sb atoms to simulate the experimental ratio of Sb doping. Reproduced with permission. Copyright: 2022, Elsevier.
Figure 38. SEM image of (a) Cs3Bi2Br9, (b) Cs3Bi1.8Sb0.2Br9, (c) Cs3Bi1.5Sb0.5Br9, (d) Cs3Bi0.7Sb1.3Br9, (e) Cs3Sb2Br9, inset in (ac) show the magnified view of the SEM images; (f). Transient photocurrent response; (g) Photocatalytic performances of the prepared samples after irradiation of 3 h; (h) Bader electrons of upper layer minus that of lower layer over different models. Cs12Bi8Br36 is used as a model of Cs3Bi2Br9, Cs12Bi8Br36 were substituted by 0, 1, 2, 5, and 8 Sb atoms to simulate the experimental ratio of Sb doping. Reproduced with permission. Copyright: 2022, Elsevier.
Nanomaterials 14 00094 g038
Figure 39. (a) Photocatalytic activity of toluene oxidation with CsxMA3−xSb2Br9 NP (x = 0–3) photocatalysts; (b) optical absorption spectra of CsxMA3−xSb2Br9 NPs (x = 0–3). Possible mechanism of A-site effect on C(sp3)–H bond activation; (c) photocatalytic process of toluene oxidation; middle part: octahedron distortion caused by A-site cations. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Figure 39. (a) Photocatalytic activity of toluene oxidation with CsxMA3−xSb2Br9 NP (x = 0–3) photocatalysts; (b) optical absorption spectra of CsxMA3−xSb2Br9 NPs (x = 0–3). Possible mechanism of A-site effect on C(sp3)–H bond activation; (c) photocatalytic process of toluene oxidation; middle part: octahedron distortion caused by A-site cations. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Nanomaterials 14 00094 g039
Figure 40. (a) HAADF-STEM images of 10 wt% loading; (b) PL spectra of supported and bulk Cs3Bi2Br9; (c) Toluene conversion rate over different samples. For all supported samples, SBA-15 was used as support and the effective catalyst mass was the weight of loaded halide perovskite phase. Inset. Electronic structure calculations. Focus on a typical Br-H geometry. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Figure 40. (a) HAADF-STEM images of 10 wt% loading; (b) PL spectra of supported and bulk Cs3Bi2Br9; (c) Toluene conversion rate over different samples. For all supported samples, SBA-15 was used as support and the effective catalyst mass was the weight of loaded halide perovskite phase. Inset. Electronic structure calculations. Focus on a typical Br-H geometry. Reproduced with permission. Copyright: 2020, John Wiley and Sons.
Nanomaterials 14 00094 g040
Figure 41. (a) HRTEM images of the 20% CASBB/CN composite; yellow dotted line represents the interface between g-C3N4 and Cs2AgBiBr6; (b) UV-vis DRS of g-C3N4, Cs2AgBiBr6, CASBB, and CASBB/CN composites; (c) band alignment of g- C3N4 and 20% CASBB/CN; (d) photocatalytic performance of toluene conversion and selectivity of benzaldehyde over the difference photocatalysts; (e) schematic of charge migration for the photocatalytic reactions in 20% CASBB/CN. Reproduced with permission. Copyright: 2023, American Chemical Society; (f) Possible mechanism for the photocatalytic oxidation of toluene by CABB-80@UCNT. Reproduced with permission. Copyright: 2023, Elsevier.
Figure 41. (a) HRTEM images of the 20% CASBB/CN composite; yellow dotted line represents the interface between g-C3N4 and Cs2AgBiBr6; (b) UV-vis DRS of g-C3N4, Cs2AgBiBr6, CASBB, and CASBB/CN composites; (c) band alignment of g- C3N4 and 20% CASBB/CN; (d) photocatalytic performance of toluene conversion and selectivity of benzaldehyde over the difference photocatalysts; (e) schematic of charge migration for the photocatalytic reactions in 20% CASBB/CN. Reproduced with permission. Copyright: 2023, American Chemical Society; (f) Possible mechanism for the photocatalytic oxidation of toluene by CABB-80@UCNT. Reproduced with permission. Copyright: 2023, Elsevier.
Nanomaterials 14 00094 g041
Figure 42. Toluene conversion rate over different samples of Cs4ZnSb2Cl12. Reproduced with permission. Copyright: 2023, American Chemical Society.
Figure 42. Toluene conversion rate over different samples of Cs4ZnSb2Cl12. Reproduced with permission. Copyright: 2023, American Chemical Society.
Nanomaterials 14 00094 g042
Figure 43. (a) Proposed catalytic mechanism on the selective oxidation of HMF utilizing MAPbBr3 Photocatalyst and Atmospheric O2; (b) schematic representation for the photocatalytic selective oxidation of HMF based on the Cs2SnBr6/rGO catalytic system. Reproduced with permission. Copyright: 2020, American Chemical Society; (c) photocatalytic selective oxidation of HMF performance using different compositions of Cs2SnBr6/rGO. Reproduced with permission. Copyright: 2023, John Wiley and Sons.
Figure 43. (a) Proposed catalytic mechanism on the selective oxidation of HMF utilizing MAPbBr3 Photocatalyst and Atmospheric O2; (b) schematic representation for the photocatalytic selective oxidation of HMF based on the Cs2SnBr6/rGO catalytic system. Reproduced with permission. Copyright: 2020, American Chemical Society; (c) photocatalytic selective oxidation of HMF performance using different compositions of Cs2SnBr6/rGO. Reproduced with permission. Copyright: 2023, John Wiley and Sons.
Nanomaterials 14 00094 g043
Figure 44. (a) C-C coupling reactions of benzyl bromide (1a), para-substituted benzyl bromide (2a); (b) proposed mechanism of the photocatalytic benzyl bromide photoreduction; (1). photoinduced electron and hole generation, charge diffusion, and substrate–surface approximation; (2). electron transfer to the substrate to generate the radical anion of the benzyl bromide and hole transfer to the MeOH (formaldehyde and H2 formation); (3) adsorption of the Br anion and benzyl bromide radical to the NC surface; the strength of the binding of the benzyl radical to the NC surface would depend on the electronegativity of the substituent at the p-position; and (4) competitive formation of the reductive dehalogenation product vs. the C–C coupling product.
Figure 44. (a) C-C coupling reactions of benzyl bromide (1a), para-substituted benzyl bromide (2a); (b) proposed mechanism of the photocatalytic benzyl bromide photoreduction; (1). photoinduced electron and hole generation, charge diffusion, and substrate–surface approximation; (2). electron transfer to the substrate to generate the radical anion of the benzyl bromide and hole transfer to the MeOH (formaldehyde and H2 formation); (3) adsorption of the Br anion and benzyl bromide radical to the NC surface; the strength of the binding of the benzyl radical to the NC surface would depend on the electronegativity of the substituent at the p-position; and (4) competitive formation of the reductive dehalogenation product vs. the C–C coupling product.
Nanomaterials 14 00094 g044
Table 2. State-of-the-art of the photocatalytic toluene oxidation of lead and lead-free perovskites. C% and S% represent conversion and selectivity, respectively.
Table 2. State-of-the-art of the photocatalytic toluene oxidation of lead and lead-free perovskites. C% and S% represent conversion and selectivity, respectively.
PhotocatalystLight SourceBand
Gap
Synthesis
Method
C %S %Activity
(µmol·g−1·h−1)
Ref
FAPbBr3150 W Xe lamp, AM 1.5 G simulated light irradiation2.2LARP0.07581320[97]
CsPbBr3300 W Xe lamp with 420 nm filter2.3LARP--710[124]
CdS300 W Xe lamp, visible light (λ > 420 nm)2.25room temp.33100240[167]
TiO2Six 6 W UV lamps (310 nm)~3.7Hydrothermal 21%9094[168]
Flower-like Bi2WO6300 W Xe lamp, visible light (λ > 420 nm)2.96Hydrothermal 1.5%~85464[169]
CsPbBr3xClx/TiO2300 W Xe lamp with 420 nm filter2.62LARP & anion exchange--1874[124]
5% NiOx/FAPbBr3/TiO2150 W Xe lamp, AM 1.5 G-LARP0.85%863800[97]
Cs3Bi2Br9 NCs300 W Xe lamp2.75LARP-100850[141]
Cs3Bi2Br9/Mxene-7.5300 W Xe lamp~2.5In-situ growth--4011[141]
Cs3Bi2Br9/SBA-15(10%)300 W Xe lamp with 420 nm filter2.74wetness impregnation-9012,600[160]
Cs3Sb2Br9150 W Xe lamp, AM 1.5 G with 420 nm filter2.3Anti-solvent precipitation--1195[148]
g-C3N4150 W Xe lamp, AM 1.5 G with 420 nm filter2.67Pyrolysis--298.3[148]
Cs3Sb2Br9/g-C3N4 150 W Xe lamp, AM 1.5 G with 420 nm filter2.3Anti-solvent precipitation--8347[148]
Cs2AgBiBr6500 W Xe lamp, with 420 nm filter2.21hydrothermal-1053[165]
Cs2AgSbxBi1−xBr6500 W Xe lamp, with 420 nm filter2.07hydrothermal-40448[165]
Cs2AgSbxBi1−xBr6/CN (20%)500 W Xe lamp, with 420 nm filter-Mechanical grinding-961088[165]
Cs4ZnSb2Cl12 nanoplates500 W Xe lamp, with 420 nm filter2.36Hot injection-951893[29]
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

Jagadeeswararao, M.; Galian, R.E.; Pérez-Prieto, J. Photocatalysis Based on Metal Halide Perovskites for Organic Chemical Transformations. Nanomaterials 2024, 14, 94. https://doi.org/10.3390/nano14010094

AMA Style

Jagadeeswararao M, Galian RE, Pérez-Prieto J. Photocatalysis Based on Metal Halide Perovskites for Organic Chemical Transformations. Nanomaterials. 2024; 14(1):94. https://doi.org/10.3390/nano14010094

Chicago/Turabian Style

Jagadeeswararao, Metikoti, Raquel E. Galian, and Julia Pérez-Prieto. 2024. "Photocatalysis Based on Metal Halide Perovskites for Organic Chemical Transformations" Nanomaterials 14, no. 1: 94. https://doi.org/10.3390/nano14010094

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop