Next Article in Journal
A Guided Walk through the World of Mesoporous Bioactive Glasses (MBGs): Fundamentals, Processing, and Applications
Next Article in Special Issue
The Photocatalytic Conversion of Carbon Dioxide to Fuels Using Titanium Dioxide Nanosheets/Graphene Oxide Heterostructure as Photocatalyst
Previous Article in Journal
Biodegradable Poly(Lactic Acid) Nanocomposites for Fused Deposition Modeling 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments in Lead and Lead-Free Halide Perovskite Nanostructures towards Photocatalytic CO2 Reduction

Department of Energy Science & Engineering, DGIST, 333 Techno Jungang-daero, Hyeonpung-eup, Dalseong-gun, Daegu 42988, Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(12), 2569; https://doi.org/10.3390/nano10122569
Submission received: 19 November 2020 / Revised: 17 December 2020 / Accepted: 17 December 2020 / Published: 21 December 2020
(This article belongs to the Special Issue Nanomaterials toward CO2 Reduction and Conversion)

Abstract

:
Perovskite materials have been widely considered as emerging photocatalysts for CO2 reduction due to their extraordinary physicochemical and optical properties. Perovskites offer a wide range of benefits compared to conventional semiconductors, including tunable bandgap, high surface energy, high charge carrier lifetime, and flexible crystal structure, making them ideal for high-performance photocatalytic CO2 reduction. Notably, defect-induced perovskites, for example, crystallographic defects in perovskites, have given excellent opportunities to tune perovskites’ catalytic properties. Recently, lead (Pb) halide perovskite and their composites or heterojunction with other semiconductors, metal nanoparticles (NPs), metal complexes, graphene, and metal-organic frameworks (MOFs) have been well established for CO2 conversion. Besides, various halide perovskites have come under focus to avoid the toxicity of lead-based materials. Therefore, we reviewed the recent progress made by Pb and Pb-free halide perovskites in photo-assisted CO2 reduction into useful chemicals. We also discussed the importance of various factors like change in solvent, structure defects, and compositions in the fabrication of halide perovskites to efficiently convert CO2 into value-added products.

1. Introduction

Photocatalytic CO2 reduction, which transforms solar energy into usable chemical fuels, has drawn considerable attention for solving environmental pollution and global energy problems [1,2,3,4,5,6,7,8,9,10,11]. Many attempts have been taken to upgrade the photocatalytic CO2 conversion process in terms of material design. In any event, the discovery of a new photocatalyst for better synergistic performance has never stopped. To date, several types of catalysts have been employed for CO2 conversion, including metal oxides, nitrides, sulfides, selenides, chalcogenides, and perovskite materials [12,13,14,15,16,17,18,19,20,21,22,23]. These materials have made significant progress, but many of them have several drawbacks, such as high-cost synthetic approaches, lengthy/complicated synthesis process, long-term instability, and less catalytic activity. While recently, perovskites have been significantly attracted as a better replacement for traditional semiconductors for the photocatalytic CO2 reduction process due to their extraordinary optoelectronic properties and cost-effectiveness [24,25].
Typically, perovskite materials are indicated with the chemical formula ABX3 [26,27]; here, A site occupies large size cation (e.g., Cs+, Rb+, methylammonium), and B sites are occupied by small size cation (e.g., Pb2+, Sn2+). At the same time, X (e.g., O2−, Br, Cl, I) holds an anion that bonds to both A and B. According to the crystallographic perspective, the ideal perovskite structure is cubic and unbending; however, most perovskites are generally distorted. Based on the perovskite elements, the properties such as chemical stability, bandgap energies, optical stability, and crystal structure of the catalyst can be tunable [28]. In 1978, Hemminger et al. first-time observed photosynthetic reaction using SrTiO2 perovskite materials for the conversion of CO2 into CH4 in the presence of gaseous water and CO2 [29]. The SrTiO2 (111) crystalline phase with Pt foil was active for the CO2 conversion. The Pt foil was responsible for the adsorption of the CO2 on the metal faces, but the limitation was metal poisoning. In the case of SrTiO3, the regeneration of the Ti3+ after the oxidation reaction in water, surface properties, and crystallite size play a key role. Afterward, SrTiO3, Sr2TiO4, NaNbO3, and H2SrTa2O7 have been documented for the CO2 reduction process [30,31,32,33]. Luo et al. observed the relationship between the surface structure of SrTiO3 and photocatalytic activity for CO2 photoreduction [30]. The SrTiO3 was treated with the etching process and functionalized with the OH group. The SrO-terminated surface exhibits nucleophilicity, which allows CO2 adsorption, and TiO2-terminated is electrophilic. The effect of the electronic properties also differs from these two surfaces; the Sr 4d orbital conduction band level is more negative than the Ti 3d orbital. Though in the case of SrO-terminated or Sr(OH)2-decorated allowed the highest CO2 fixation than the Ti-rich surface, but the photoreduction activity was low. Because of the produced surface, molecules were attached to the weakly active Sr ions and exhibited lower reactivity. The Ti-rich catalyst showed the highest activity because the active Ti-edges shifted light absorption in the visible region. Therefore, these studies showed that the properties of perovskites, especially structural flexibility, will be of great interest to study as efficient materials for large-scale photocatalytic applications.
Among the various perovskite materials, halide perovskites have been successfully emerging as an efficient catalyst due to their extraordinary properties like cost-effectiveness, easy synthesis process, visible light absorption, high CO2 adsorption surface area, surface disorders for charge trapping, and tunable structure [34,35,36,37,38,39,40,41,42]. There are two classes of materials in the halide perovskites: (1) Lead-based halide perovskites and (2) lead-free halide perovskites. Most of the lead halide perovskites offer a narrow bandgap as compared to traditional semiconductors. The hybrid composites of lead halide perovskites with other semiconductors/materials can efficiently enhance the yield and stability of the CO2 conversion. Later, however, few more studies have been reported on lead-free halide perovskites to avoid toxicity. Therefore, lead and lead-free halide perovskites’ recent development towards reducing CO2 in different materials has fascinated us. This review would offer a thorough look into the recent success of optimal halide perovskites for the CO2 photoreduction application.

2. Fundamentals of Photocatalytic CO2 Reduction

The process of thermal catalysis is the easiest option for CO2 reduction. Still, it is not environmentally appropriate due to an endothermic reaction so that the input energy must be high [43]. As we already knew, nature is doing the same process that converts CO2 to hydrocarbons by the photosynthesis process. So why not artificial photosynthesis can help to solve the problem of CO2 reduction? The process of CO2 reduction begins after the chemisorption of the CO2 molecules on the catalytic surface. Therefore, the CO2 reduction process’s high efficiency can be achieved by employing high surface area catalysts to maximize CO2 molecules’ adsorption on the catalyst’s surface. As we know, CO2 is a highly stable molecule, therefore to break the C-O bonding, it requires 750 kJ mol−1 of the bond dissociation energy, which is approximately 54% higher than C-H bond dissociation energy (411 kJ mol−1). To reduce CO2 molecules, the conduction and valence band potentials of particular catalysts should be above and below the standard redox potentials of the products, respectively. Thus, to convert the CO2 into any product, it is necessary to cross the high energy of the activation barrier for forwarding reaction [44]. In the conversion of CO2, a single-electron transfer is practically inconceivable to produce CO2●− due to the necessity of high redox potential of −1.90 V vs. NHE (see Table 1). Therefore, as a single electron process is practically challenging, the proton assisted CO2 reduction process is thermodynamically feasible to form various products. Depending upon the concentration of available electrons and protons, the CO2 reduction proceeds into the formation of various products; for instance, CO evolution can be achieved by contributing 2e and 2H+ at the redox potential −0.53 V (pH = 7). Consequently, an increase in the reduction potential concerning the number of electrons, converting CO2 to ethane, desires the 14e and 14H+. In recent days, there have been several excellent reviews based on the fundamentals of photo-reduction of CO2 [45,46,47,48,49,50]; thus, readers may turn to these reviews for more details. Our review article focuses on the fabrication of lead and lead-free halide perovskites’ and their composites, the effect of solvents, and other parameters to improve the catalytic performance towards the CO2 conversion process.

3. Lead Halide Perovskites for Photocatalytic CO2 Reduction

So far, organic–inorganic lead halide perovskites (LHPs) have been well established for photovoltaic applications with high power conversion proficiency [25]. Later, they have been successfully used for various applications, including photodetector, laser, LED, thermoelectric, and piezoelectric [51,52]. Recently, Zhu et al. reported LHPs for organic synthesis that have fundamental significance in drug production [53]. The high efficiency of LHPs was mainly attributed to the properties that are very suitable for photocatalytic applications, such as high absorption coefficient, greater defect tolerance, superior photogenerated charge-carrier lifetime, and high carrier mobility [54]. In recent years, LHPs have also been used as photocatalytic materials for various applications like hydrogen evolution reaction [55,56], organic pollutant degradation [57,58], and alkylation of aldehyde [41,51]. In addition, they have been proved the best materials for oxygen evolution reactions; thus, water can be used as an electron source for the photoreduction of CO2 by preventing the use of sacrificial agents [59]. Thus, the introduction of halide perovskites for CO2 photoreduction has made substantial progress in the field of catalysis. These materials possess poor water stability due to the ionic nature of the LHPs. Numerous attempts have been made to boost the catalytic activity and stability of LHPs towards photocatalytic CO2 reduction. In this contribution, Wu et al. encapsulated methylammonium lead iodide quantum dots (QDs) in iron-based metal–organic frameworks (MOFs) and successfully utilized them for CO2 photoreduction [60]. They have used Fe-porphyrin-based MOFs to increase the water stability of perovskite. Fe act as an active catalytic site for CO2 photoreduction, which also suppresses the charge recombination and effectively enhances the charge transportation. The optimized samples of MAPbI3 with Fe-porphyrin MOF PCN-221(Fex) exhibited a total hydrocarbon yield of 1559 µmol g−1 combined for CO and CH4 production. As the concentration of Fe in PCN-221 (Fex) increases, the improvement in CO and CH4 formation was observed. The catalyst showed excellent stability over 80 h, which was much higher than pristine halide perovskite QDs. Remarkably, water was used as a sacrificial agent with ethyl acetate, which acts as an electron source for the CO2 reduction reaction. The light-harvesting efficiency of MAPbI3 and fast electron transfer from MAPbI3 to Fe were attributed to high catalytic activity.
Apart from this exceptional research, most methylammonium halides have been documented for hydrogen production. Previous studies have shown that MAPbI3 is unstable in humid conditions; thus, other halide perovskites have been intended for CO2 photoreduction [61]. Subsequently, major advances have been made in the field of research with cesium lead halide perovskites as a novel and most common catalyst for CO2 reduction in recent years. Most of the inorganic Cs-based perovskites possess suitable valence and conduction band potentials for the CO2 reduction reaction. To this end, Hou et al. suggested a controlled synthesis of colloidal QDs of CsPbBr3 and investigated the size-dependent CO2 photoreduction into CH4, CO, and H2 [61]. Using less costly inorganic precursors and oleylamine/oleic acid as a surface ligand, the QDs were synthesized using a simple solution process. These colloidal QDs display the tunable particle size due to the quantum confinement effect at different temperatures, as shown in Figure 1a–d. The optical properties of CsPbBr3 showed a significant impact on photocatalytic efficiency (Figure 1e,f). During the photocatalytic reaction, the average electron yield for CO2 reduction was 20.9 μmol g−1. A Time-resolved PL study signifies that an optimized catalyst has the most extended lifetime responsible for enhanced catalytic activity (Figure 1g). The band alignment of the valence and conduction band were well matched for CO2 reduction and water oxidation leading to the formation of CO, CH4, and H2 (Figure 1h). Therefore, such QDs can be used as inexpensive catalysts for catalytic applications.
Mixed halide perovskites, i.e., a mixture of the cation (MA+, Cs+) and anion (Cl, I, Br), can further increase the efficiency of catalytic reaction towards CO2 reduction. Tuning the halide ratio can change the structural combination of perovskites that significantly impact catalytic behavior. Thus, Su and co-workers reported a low-cost, cubic phase CsPb(Br0.5/Cl0.5)3 perovskite with varied ratios of Br and Cl and studied for photocatalytic CO2 reduction [62]. The controlled synthesis of mixed halide perovskite was carried out by the hot injection method. Mixed halide perovskite, i.e., CsPb(Brx/Cl1−x)3, offers excellent absorption in the visible region ranging from 400–700 nm, relatively broader than distinctive semiconductors like TiO2 and ZnO. The calculated band edges were quite suitable for CO2 reduction potentials of CO and CH4, with the bandgap ranging from 2.33 to 2.98 eV for CsPb(Brx/Cl1−x)3 samples. While in the emission spectra, the gradual shifting and quenching of pristine CsPbBr3 from 517 to 413 nm was observed as the amount of Cl increases, which is attributed to the improved electron separation. The authors performed a CO2 reduction test in ethyl acetate. The increased CO and CH4 formation was found with an increase in the Cl concentration, and the optimized sample achieved much higher catalytic activity than the pristine CsPbBr3 and CsPbCl3 perovskites. The time-dependent CO2 reduction rate for the optimized catalyst demonstrated strong stability over the 9 h for CH4 and CO evolution. The increased activity and stability of mixed halide perovskite were attributed to the controlled ratios of Br and Cl.
It is well known that pristine semiconductor/perovskites suffer from poor catalytic activity and stability due to the rapid recombination of photogenerated charges or lack of suitable optical absorption, or ineffective CO2 adsorption. Therefore, several studies have been documented with metal doping for cesium halide perovskites to boost catalytic activity, stability, and selectivity. In order to attain these distinct characteristics, it is crucial to consider the mechanism of product formation and the pathway of CO2 reduction over the catalyst. Tang et al. theoretically studied the effect of metal doping in CsPbBr3 on product selectivity by employing DFT calculations (Figure 2) [63]. The model used for this analysis has considered the chemical potential of an electron and proton is equal to half of the hydrogen gas phase at standard pressure. The first step of CO2 reduction formed the HCOO*, and further adsorption of the excessive proton on the oxygen atom leads to HCO*OH formation. The HCO*OH is a spontaneous reaction; however, dissociation of HCO*OH to CO and H2O is the nonspontaneous reaction in the case of pristine-CsPbBr3. After Co and Fe doping, the catalytic activity was improved and showed a downhill reaction. The numerical values of Gibbs free energy indicated that Co and Fe doping increases the rate of chemical reaction approximately two times for catalytic activity than pristine-CsPbBr3 material. Pristine CsPbBr3 was initially examined to reduce CO2, and the findings of the free barrier energy showed that CsPbBr3 is inactive for the evolution of any hydrocarbons. While Fe and Co-doped CsPbBr3 were investigated for selective formation of CH4, these metals have great potential to break the O-C-O bond. The resultant activity and selectivity towards CH4 formation were ascribed to the successful adsorption and activation of CO2 on the doped CsPbBr3. After this, Pradhan and co-workers practically explored Fe (II)-CsPbBr3 for selective CH4 evolution via CO2 photoreduction [64]. Doping of Fe (II) replaces Pb (II) in the CsPbBr3 lattice, and an increased CH4 selectivity was observed over the increment of Fe (II) concentration; however, pristine CsPbBr3 forms CO. Later, another study by Su et al. showed that Mn2+ substitution to perovskite could significantly improve the optical and thermal properties of halide perovskites [65]. They observed that Mn-doped cesium lead halide (Br/Cl) perovskite showed more than 14 times improvement in catalytic performance than pristine catalysts. The Mn-CsPb(Br/Cl) exhibited the catalytic towards CO and CH4 formation with a yield of 1917 µmol g−1 and 82 µmol g−1, respectively. Other metals such as Fe and Co have been reported as ideal dopants with cesium perovskite, and improved water sustainability was reported [63]. For instance, Fe (II) doped CsPbBr3 and Co-doped CsPbBr3 have been reported for the evolution of CO and CH4 by Prahan et al. and Lu et al., respectively. Particularly, Fe doped CsPbBr3 predominantly forms CH4 while pristine nanocrystals (NCs) are selective towards CO formation [64].
Afterward, more outcomes of metal confined halide perovskites are reported. Lu and co-workers’ study showed that Co-doped CsPbBr3/Cs4PbBr6 combination improved catalytic performance for CH4 and CO evolution [59]. Co doping to perovskite has two benefits; first, it produces surface trap states due to the presence of Co2+ and extends the lifetime of photogenerated charges. Second, it broadens the adsorption of CO2* intermediate to the catalytic surface. Remarkably, they have employed hexafluorobutyl methacrylate to improve NC’s stability and dispersity in an aqueous medium. Then, in this class of materials, Zhang and coworkers reported Co-doped CsPbBr3 embedded in the matrix of Cs4PbBr6 (a surface protector) [66]. The reason behind such protection of Cs4PbBr6 was to improve the stability of ligand-free CsPbBr3 perovskite. As a result, the optimized Co-doped perovskite composite exhibited excellent CO2 reduction towards CO formation with 1835 µmol g−1 in 15 h. In combination with the acetonitrile/water solution, the methanol was used as a hole scavenger to improve the catalytic activity towards CO2 reduction. Interestingly, Cs4PbBr6 does not take part in the redox reaction due to its unsuitable band potentials/alignment; however, CsPbBr3 takes part in the CO2 reduction process and triggers the catalytic activity towards CO formation. After these reports, Pt has also been successfully used as a co-catalyst for CsPbBr3 [67]. This study stated the solvent effect, where it was observed that acetate is the most potent solvent, which provides a stable atmosphere for perovskite to conduct the CO2 reduction process. The optimized Pt loaded catalyst demonstrated an electron consumption rate of 5.6 µmol g−1 h−1 in ethyl acetate. In another study of Zhu and co-workers, the water-stable CsPbCl3 was documented with Mn and Ni doping. They illustrated the significance of the Pb-rich surface, which extends the lifetime of PL to increase the catalytic activity [68]. By interpenetrating solid–liquid, the synthesis of surface Pb enriched CsPbCl3 was accomplished by allowing the water to drain through the CsPbCl3 layer originating with Cs+ and Cl. These Cs+ and Cl inhibit the decomposition of CsPbCl3 and increase the PL lifetime. Therefore, Ni-doped Pb-rich CsPbCl3 QDs displayed superior CO2 reduction behavior against CO evolution with a rate of 169.37 μmol g−1 h−1.
In addition to metal doping/deposition, perovskites are combined with supporting material to increase light absorption and charge separation. Among various materials, graphene can be a good choice as a support material for perovskites due to its well-known surface, optoelectronic, and physicochemical properties [69,70], which ultimately prolongs the electron/hole pair’s lifetime and makes perovskite ideal for photocatalytic CO2 conversion. Hence, Xu et al. fabricated CsPbBr3 QDs/graphene oxide (GO) composite by a simple precipitation process for photoconversion of CO2 in a nonaqueous medium [71]. The CO2 reduction reaction was conducted in a Pyrex bottle using ethyl acetate as a solvent. The use of ethyl acetate has two benefits; (i) it stabilizes CsPbBr3 due to its moderate polarity; and (ii) it increases the solubility of CO2 even more than in water. The presence of GO provides additional electron transfer from CsPbBr3 to GO; therefore, it was found that the CsPbBr3/GO composite exhibited improved catalytic activity compared to pristine CsPbBr3 QDs. The catalyst displayed stability over 12 h, and no phase change of CsPbBr3 was observed during the catalytic reaction. Excited electrons from CsPbBr3 quickly transfer to the GO sheet, suppress the electron-hole pair’s recombination, and increase catalytic activity.
Shortly after, few more studies have been reported in combination with graphene. For instance, Eslava and the group reported a surfactant-free synthesis of CsPbBr3 NCs [72]. They have synthesized CsPbBr3 on a gram scale by employing the simple mechanochemical process to get the different morphology of NCs, including nanorods, nanosheets, and nanospheres. To improve the catalytic efficiency, CsPbBr3 was further combined with Cu-RGO by the mechanochemical process. The CsPbBr3 nanosheets with Cu-RGO achieved 12.7, 0.46, and 0.27 μmol g−1 h−1 of CH4, CO, and H2 evolution rates after CO2 reduction. The catalyst composed of CsPbBr3-Cu-RGO achieved 1.10% apparent quantum efficiency and showed excellent stability over three consecutive runs. Such an alternative method of large-scale synthesis is notable and essential for advancing the photocatalytic technology practically. Later, Wang and co-workers demonstrated the effectiveness of RGO sheets combined with Cs4PbBr6 for CO2 reduction [73]. They revealed that the dual nature of the RGO is responsible for increased catalytic activity and stability, (i) the defect-induced RGO efficiently traps the electron excited by Cs4PbBr6, and (ii) oxygen-deficient RGO adsorbs and stimulates CO2 molecule. In addition to these studies, Mu and co-workers described the ultrathin small-sized graphene oxide (USGO) as an electron mediator between CsPbBr3/α-Fe2O3 Z-scheme photocatalyst [74]. Intense contact between CaPbBr3/USGO through the Br-O-C bond; and USGO/α-Fe2O3 via the C-O-Fe bond accelerates the transfer of electrons through the Z-scheme. As a result, such a Z-scheme combination achieved 9 times greater catalytic efficiency compared to CsPbBr3 NCs. Later on, such a Z-scheme heterojunction of CsPbBr3 QDs was also reported with Bi2WO6 nanosheet [75]. Therefore, the Z-scheme combination then performs efficient charge separation across the closely connected interface and enhances the catalytic activity.
Recently, 2D materials have grabbed huge attention as the best supporting materials in many applications. A halide perovskites/2D composite materials strategy can achieve an efficient charge transfer and reduced electron-hole recombination among the perovskite-based photocatalysts. Xu et al. produced the CsPbBr3/Pd Schottky junction in 2018 and studied the enhanced consumption rate of electrons for photoreduction of CO2 [76]. For this analysis, CsPbBr3 NCs were deposited on Pd nanosheets under atmospheric conditions using a simple technique on a glass substrate, and the photocatalytic reduction of CO2 to CO/CH4 was studied. The charge transport and charge carrier dynamics among the composite were studied by PL and fs-TAS (femtosecond transient absorption spectroscopy). The PL quenching of 0.5–8.6% in the CsPbBr3/Pd composite was observed compared to the pristine sample (Figure 3a). Likewise, the decreased PL decay (Figure 3b) of the composite was observed in TRPL analysis with an average lifetime of 2.71–13.38 ns; however, such a decay lifetime for CsPbBr3 was measured to be 52.03 ns. Furthermore, a similar trend was observed in the fs-TAS analysis (Figure 3c), where a large decrease in peak intensity of composite than the pristine sample was observed. These findings indicate that the recombination rate of the electron-hole in CsPbBr3 was reduced by integrating it over Pd nanosheets, which consciously promotes the process of CO2 reduction to form CH4 and CO (Figure 3d). Remarkably, the Schottky contact between CsPbBr3/Pd composite increases the electron consumption rate up to 33.80 µmol g−1 (for CH4 and CO evolution), 2.43 times higher than the pristine CsPbBr3 with the improved quantum efficiency (Figure 3e,f). Therefore, a combination of halide perovskite with 2D materials proved to be a good strategy for the photocatalytic CO2 reduction reaction.
Later on, Liu et al. introduced a functional CsPbBr3/MXene nanocomposite for CO2 reduction to CO and CH4 under visible light [77]. The consistent growth of CsPbBr3 on exfoliated MXene−n (n = 10, 20, 30, 40, and 50, a different amount of MXene) nanosheet was achieved by an in-situ method, as shown in Figure 4a. The etching of Ti3AlC2 was carried out by using the HCl-HF solution to form a Ti3C2Tx nanosheet. Then, the final nanocomposite was obtained by the exfoliation of multilayered Ti3C2Tx and in-situ growth of CsPbBr3 on Ti3C2Tx nanosheets. The dispersion of cubic CsPbBr3 with an average size of 25 nm on MXene nanosheets was observed in the TEM images (Figure 4b–d), and the presence of Ti, Pb, Br, and Cs confirms the constant growth of perovskite NCs on MXene nanosheet (Figure 4e–i). As expected, the PL and TRPL quenching in composite compared to pristine perovskite confirm the efficient charge transfer among the interface of CsPbBr3/MXene. The photocatalytic CO2 reduction test was carried out under light irradiation using ethyl acetate solvent towards CO and CH4 formation with a rate of 26.32 and 7.25 μmol g−1 h−1, respectively. Therefore, such perovskite/2D composites can be used as an efficient catalyst for photocatalytic applications.
Construction of perovskites/semiconductor heterojunction is advantageous to improve the optoelectronic or photochemical properties of catalytic systems. As well known, when two semiconductors with different band potentials combine, heterojunction forms at the interface of particular semiconductors and facilitate the charge separation process. Among various candidates, TiO2 was proven to be an excellent semiconducting material all over the years for photocatalytic applications. In this aspect, Xu et al. in 2018 reported an amorphous TiO2 encapsulated CsPbBr3 composite with enhanced catalytic efficiency [78]. The improved separation of the charge between CsPbBr3 and TiO2 was shown by the decay of the PL and TRPL, where dramatic quenching of radiative recombination was observed. Moreover, the fs-TAS analysis revealed that the decreased electron-hole recombination improves the charge separation efficiently among the composite. As a result, the CsPbBr3/amorphous TiO2 composite demonstrated 6.5 times better photoelectron intake during the CO2 photoreduction and stability over 30 h. The high selectivity towards CH4 formation was attributed to the perfect combination of CsPbBr3 and amorphous TiO2. The electron generated by CsPbBr3 accumulates on TiO2 to break the dynamic barrier and therefore fast-track the rate of CH4 formation. Thus, such a study helps to understand the surface modification of halide perovskites for various applications.
Afterward, in 2020 Yu and co-workers developed a self-assembled CsPbBr3 QDs/TiO2 nanofibers, an S-scheme heterojunction hybrid for CO2 reduction under the irradiation of UV-Vis light [79]. The electron microscopy results revealed that CsPbBr3 QDs were uniformly distributed on TiO2 nanofibers. Density functional theory (DFT) and experimental studies were combined to comprehend the interfacial charge transfer among the composite. The chemical states of pristine and composite samples were explored by in-situ and ex-situ XPS analysis. As shown in Figure 5a,b, the Ti 2p peak of Ti4+ ions, O 1s of lattice oxygen, and surface -OH group are present in all the samples. However, in the in-situ measurement, the binding energies (BE) of Ti 2p and O 1s peaks shifted towards higher BE than ex-situ spectra. A similar observation but opposite peak shift was observed in the Br 2d peak, mainly attributed to an electron transfer from CsPbBr3 to TiO2 (Figure 5c). Such electron transfer is responsible for constructing the S-scheme heterojunction among TiO2/CsPbBr3, which efficiently separates the photogenerated charges to promote CO2 reduction. The electron transfer from CsPbBr3 to TiO2 was further validated by work function values calculated from the energy difference of vacuum and fermi levels, as shown in Figure 5d–f. Due to the lower Fermi level of TiO2 than CsPbBr3 QDs (work function (ϕ) of CsPbBr3, 5.79 eV and TiO2, 7.18 or 7.08 eV), the electron flow would be favorable from CsPbBr3 to TiO2 for enabling the phases at the similar Fermi level and created an internal electric field at the interface of TiO2/CsPbBr3. A similar observation was observed in DFT calculations. Therefore, such an improved electron transfer is responsible for the enhanced catalytic performance of hybrid (9.02 µmol g−1 h−1) than the pristine CsPbBr3 and TiO2 (4.94 and 4.68 µmol g−1 h−1), respectively, towards CO formation. The improved CO2 adsorption on CsPbBr3 QDs and S-scheme heterojunction formation was ascribed to superior photocatalytic activity.
Next, 3D structures, e.g., 3D microporous graphene, exhibited possible support material for catalytic reactions [80,81]. Such a 3D network structure can provide more CO2 reactive sites and provide fast charge transport across multidimensional networks. Kaung and co-workers fabricated the hierarchical ternary nanocomposite of CsPbBr3 with ZnO nanowire/3D graphene through a multi-step process for photocatalytic CO2 reduction [81]. First, in-situ 1D ZnO/2D RGO macropores with a high specific surface area were fabricated on a film. Then, as-prepared CsPbBr3 was used for the synthesis of ternary composite through the centrifugation cast method. SEM and TEM images revealed that the CsPbBr3 NCs are well decorated in the ZnO nanowires over RGO. Similarly, optoelectronic and surface properties showed improved light harvesting in the visible region and improved CO2 adsorption on the catalyst’s surface, respectively. As a result, the ternary composite exhibited 52.02 µmol g−1 h−1 of CH4 evolution (96.7% selectivity) along with CO formation. The electron pathway for CO2 reduction was achieved via CsPbBr3 to 1D ZnO to 3D RGO.
Apart from metal oxides, CsPbBr3 can also be anchored with g-C3N4 due to its superior properties, such as visible-light active, tunable band potentials, and rich active surface area [82,83]. The combination of CsPbBr3 and g-C3N4 facilitates efficient charge transport through their closely connected interface. In this regard, Xu et al. attached CsPbBr3 QDs to amino-functionalized g-C3N4 nanosheet through N-Br bonding [83]. The 20 wt.% contained QDs anchored on g-C3N4 achieved a superior photocatalytic CO2 reduction towards CO formation with a rate of 149 μmol h−1 g−1 in acetonitrile/water solvent. The surface functionalization of g-C3N4 with abundant NHx was shown to help build a bridge between g-C3N4 and CsPbBr3. Such N-Br bonding was confirmed by XPS analysis, which is responsible for the enhanced charge separation and decreased electron-hole recombination rate. Resulting, the composite showed 15- and 3-fold improved catalytic activity than pristine CsPbBr3 QDs and g-C3N4. Later, in 2019 a similar study was reported by Zhang and co-workers for CsPbBr3/g-C3N4 containing TiO species (TiO-CN) [84]. The well-defined composite of CsPbBr3@TiO-CN was able to undergo CO2 reduction to produce a 129 μmol g−1 of CO under 10 h visible light irradiation. The interaction among CsPbBr3 and g-C3N4 was established via N-Br and O-Br, reducing the recombination of the electron/hole pair. The electrons generated in the conduction band of CsPbBr3 transfer to TiO-CN nanosheet and react with adsorbed CO2 molecules. At the same time, water oxidation was carried out by holes accumulated at the valence band of CsPbBr3.
The core-shell combinations have been reported as an alternative option to improve halide perovskite’s stability and activity towards CO2 reduction, where coating the surface of perovskite may also upsurge water stability. Hence, various materials, including metal oxides, polymers, silica, zeolites, and metal–organic framework (MOF), have been successfully utilized. These days, MOFs are widely employed for photocatalytic applications due to their unique properties such as high specific surface area, more catalytic active sites, and tunable structural flexibility. In this way, CsPbBr3@ZIF (zeolitic imidazolate framework) was reported for an efficient CO2 photoreduction [85]. In this study, CsPbBr3 was coated with the Zn-based metal-organic system ZIF-8 and the Co-based ZIF-67 by an in-situ approach that activates the CO2 molecule, as shown in Figure 6a. The HAADF-STEM images and elemental mapping confirmed the formation of the CsPbBr3@ZIF core-shell structure, as shown in Figure 6b–e. It has been reported that ZIF coating increases the stability of CsPbBr3 due to its weak hydrophobic nature. The gas-phase photocatalytic CO2 reduction with water vapor showed CH4 and CO evolution; and, the CH4 formation was increased with the increase in irradiation time and achieved 100% selectivity. The increased catalytic activity was demonstrated by composite with ZIF-67 achieving 10.53 μmol g–1 of CH4 evolution (Figure 6f). The electron consumption rate for ZIF-8 and ZIF-67 composites was 15.498 and 29.630 μmol g–1 h–1. Moreover, the CsPbBr3@ZIF catalyst showed stability for six consecutive cycles, which proves its excellent proficiency (Figure 6g). Therefore, such studies could lead to fabricated highly stable hybrid composites of perovskite materials. Later on, few more studies have been reported on a similar class of hybrid perovskites. For instance, Wang and co-workers successfully developed CsPbBr3 QDs/UiO-66(NH2) nano junction and employed it for visible-light-active CO2 reduction [86]. TEM images confirmed the construction of a nano junction between CsPbBr3 and UiO-66(NH2). The optimized catalyst was able to produce 98.57 μmol g−1 of CO and 3.08 μmol g−1 of CH4. Interestingly, the specific surface of pristine UiO-66(NH2) was 709.02 m2 g−1, which was more than the nanocomposite (465.68 m2 g−1); the catalyst exhibited much more catalytic activity than the bare samples. The catalyst’s reusability was reported for three cycles, proving its high chemical stability and photo resistivity. The suitable VB and CB potentials of CsPbBr3 and HOMO-LUMO of UiO-66(NH2) were well matched for CO2 reduction and water oxidation to generate H+ and O2, resulting to form CO. The photocatalytic reactions were carried out in ethyl acetate/H2O combination; therefore, H2 evolution could be possible. However, the absence of any co-catalyst restricts the H2 evolution and selectively produces CO more efficiently along with CH4.
Metal complexes have long been recognized for CO2 photoreduction due to their exciting characteristics such as high selectivity and CO2 conversion activity, and structural flexibility [87,88,89,90]. The conjugated structures of metal complexes are commonly employed as a multi-electron transporter in the catalytic process. Moreover, the structural flexibility of these materials proved beneficial for tailoring catalytic activity and product selectivity. Previously, metal complexes were employed for visible-light-driven CO2 reduction combined with various organic photosensitizers [91]. In the study of Kaung and co-workers in 2020, similar class of material anchoring CsPbBr3 with Re(CO)3Br(dcbpy) (dcbpy¼4,4′- dicarboxy-2,2′-bipyridine) complex has been reported [92]. The interface between Re-complex and CsPbBr3 was established through the carboxyl group, which is responsible for the fast electron transfer to boost the catalytic activity towards CO2 reduction. Hence, the optimized catalyst showed 23 times higher electron consumption rate than CsPbBr3 towards CO evolution. However, arduous synthetic procedures of photosensitizer or the use of precious metals restrict them for large-scale applications. Later, the combination of CsPbBr3 perovskite NCs with (Ni(tpy)), a hybrid transition metal complex, was developed by Gaponik et al. and cast-off for the conversion of photocatalytic CO2 into CO/CH4 [93]. The synthesis of the hybrid composite includes multi-steps, synthesis of (i) organic ligands of CsPbBr3, (ii) ligand exchange, and (iii) assembly of CsPbBr3-Ni(tpy) by immobilization. The charge transfer between the composite was confirmed by TRPL decay and transient absorption spectroscopy, where the electron transfer from CsPbBr3 to Ni(tpy) was observed. Thus, under the light irradiation, the catalyst undergoes CO2 reduction and achieved 1724 μmol g−1 of CO/CH4 formation. The catalytic activity was shown to be 26 times higher than the pristine CsPbBr3 and AQE of 0.23% for CO and CH4 evolution under monochromatic light (450 nm). Also, the catalyst showed stability over 16 h, and post catalytic analysis confirmed its high stability. Therefore, Ni(tpy) offers more catalytic sites for the CO2 molecule and improves the catalytic performance.
After MAPbBrx and CsPbBrx, Que and co-workers introduced a novel FAPbBr3 as an alternative option for traditional perovskites [94]. The synthesis of FAPbBr3 was carried out by a hot injection method, and the results were compared with CsPbBr3 synthesized similarly. As shown in Figure 7a–e, the XRD patterns and optical properties of as-prepared FAPbBr3 are almost identical to that of CsPbBr3, and they possess identical morphology with cubic shape. Despite, FAPbBr3 showed an enormous improvement in the CO evolution (main product) under the CO2 photoreduction compared to CsPbBr3, achieving 181.25 μmol g−1 h−1, which was ≈17 times greater than CsPbBr3 (Figure 7f). The significant cyclic stability was observed in FAPbBr3, preserving more than 165 μmol g−1 h−1 of CO evolution after three cycles (Figure 7g). Such high catalytic efficiency in FAPbBr3 was due to the improved lifetime of 7003 ps compared to CsPbBr3 with 956 ps.

4. Lead-Free Halide Perovskites

Throughout the years, Pb-based perovskites have been proved as the most efficient materials for photocatalytic CO2 reduction applications due to their excellent photophysical properties. Nevertheless, Pb perovskites’ high toxicity may restrict large-scale applications in the near future [95]. Li and colleagues’ recent research revealed that the biological impact of Pb-perovskite is unsafe, which shows that Pb could reach the human food chain by plants from perovskites leakage into the ground [96]. Lead exposure can cause a severe problem to human health, including nausea, clumsiness, muscle weakness, and clouded consciousness [97,98]. Therefore, eliminating Pb from perovskite structure should be the primary concern to use them for long-term applications [99,100]. To this end, numerous attempts are being made to replace Pb from halide perovskite structure with other potential candidates, including Sn, Sb, Bi, Cu, In, and Pd. In this contribution, Chu et al. published a review article on lead-free halide double perovskites for various applications covering photodetector, X-ray detector, LEDs, solar cells, and photocatalysis [98]. To date, a limited number of studies were carried out on Pb-free halide perovskites for photocatalytic CO2 reduction [101].
Recently, halide double perovskite materials have been recognized as the ideal alternative for toxic lead halide perovskites [98]. Numerous experiments have demonstrated excellent optoelectronic features of halide double perovskites, which are also suitable for photocatalytic CO2 applications [102,103]. In 2018, Zhou et al. demonstrated a highly crystalline Cs2AgBiBr6 double perovskite NCs synthesized through a hot injection process [104]. To acquire the highly crystalline Cs2AgBiBr6, the temperature was optimized, and it was observed that 200 °C is a suitable temperature to get a pure form of double perovskite. The significant role of OLA and OA ligands towards the formation of crystalline Cs2AgBiBr6 was studied, where it was observed that in the absence of these ligands, bulk Cs2AgBiBr6 was formed. The bandgap was calculated to be 2.52 eV by Tauc’s plot, and band potentials were measured by combining the results of VB-XPS and bandgap values. The enlarged band gap was observed in NCs as compared to bulk Cs2AgBiBr6, which was attributed to its quantum confinement effect. The stability of the NCs was studied in different solvents ranging from polar, partial polar, non-polar, and protonic solvents. Results revealed that the NCs were quickly decomposed in polar solvents like DMF or acetone and highly stable in mild/non-polar solvents for 3 weeks. The high ligand density on the surface of NCs may block electron/hole transportation and decrease the catalytic activity. Therefore, to decline the ligand density, the NCs were washed with absolute ethanol, which was supposed to improve the catalytic activity. The XPS and FTIR revealed that the surface ligands were wholly removed by the washing process (Figure 8a, b), while TGA results further confirm the removal of organic residues (Figure 8c, d). As shown in Figure 9a, b, the catalytic activity of CO2 reduction towards CO and CH4 evolution was much higher in the NCs washed by absolute ethanol than the Cs2AgBiBr6 NCs without the washed one in 6 h. The band potentials of the NCs are well suitable for the reduction of CO2 to produce CO/CH4, as shown in Figure 9c. The stability of catalysts was examined by post catalytic analysis using TEM, XRD, and XPS analysis. The results revealed that the catalyst’s structure and the surface did not differ from those of fresh samples. The further extension for the development of the Z-scheme Cs2AgBiBr6@g-C3N4 was carried out by Wang and coworkers [105].
The Z-scheme combination was achieved by the in-situ method, mixing g-C3N4 precursor to Cs2AgBiBr6 nanoparticles in dichloromethane/toluene. The optimized catalyst achieved 2.0 µmol g−1 h−1 of activity for CO and CH4 production with CH4 selectivity over 70%. The construction of Z-scheme among perovskite and g-C3N4 improves the redox ability of the system. After that, Sn-based halide perovskites fell into the spotlight as Pb-free materials. Wang et al. successfully developed the novel Cs2SnI6/SnS2 nanosheet combination in 2019 [106]. Such a heterojunction mixture of perovskite NCs with metal dichalcogenide (SnS2) nanosheets greatly increases the lifetime of photogenerated electrons from 1290 to 3080 ps, observed from transient absorption measurements. DFT studies confirmed the type-II band alignment in Cs2SnI6/SnS2 heterojunction, which was further supported by UPS measurement. Such a heterojunction was responsible for improved electron transportation through Cs2SnI6 and SnS2 interface, and hole extraction by Cs2SnI6 from SnS2, defeating the electron-hole recombination. As a result, the 5.4 times improved activity was observed in the Cs2SnI6(1.0)/SnS2 sample (CH4, 6.09 µmol g−1) compared to the pristine SnS2 and the stability of 3 cycles. No changes in the XRD pattern and UV-Vis-NIR spectra were observed in the samples tested after CO2 reduction. Apart from Sn-based perovskites, Bi-based materials are considered the best replacement for Pb-materials [107]. In this regard, Bhosale et al. developed a system anchored series of non-toxic, Bi-based halide perovskites, such as Rb3Bi2I9, Cs3Bi2I9, and MA3Bi2I9 by an ultrasonic, top-down method (Figure 10a). The catalyst showed 12 h of stability after seven days of aging under UV illumination, confirmed by XRD patterns. The catalytic responses were acquired at a gas-solid interface under UV irradiation. The time-dependent CH4 evolution was measured for 10 h illumination, and increased CH4 production was observed in all the samples (Figure 10b). The comparative catalyst activity for Bi-based perovskite was observed in the order of Cs3Bi2I9 > Rb3Bi2I9 > MA3Bi2I9 towards CO and Rb3Bi2I9 > Cs3Bi2I9 > MA3Bi2I9 for CH4 evolution after 10 h UV illumination (Figure 10c).
Later, halide perovskite confined with Sb metal center (i.e., Cs3Sb2Br9) was developed by Lu et al. and showed 10 times better activity than CsPbBr3 NCs [108]. The effect of the ligand in the hot injection synthesis of Cs3Sb2Br9 from CsPbX3 was studied. It was revealed that the use of saturated octanoic acid by replacing unsaturated oleic acid produces pure Cs3Sb2Br9 NCs due to the temperature expansion up to 230 °C. Moreover, these ligands were purified/removed by simple hexane/acetone washings before CO2 reduction tests. Resulting, the photocatalytic CO2 reduction of Cs3Sb2Br9 was carried out in the presence of dried octadecene solvent. The octadecene plays an important role; (i) it has low volatility and (ii) increases the solubility of CO2 compared to typical solvents like acetonitrile or ethyl acetate. After 4 h light illumination, Cs3Sb2Br9 generated 510 µmol g−1 of CO, which was over 10 times greater than various halide perovskites. The DFT calculations revealed that the sites Cs3Sb2Br9 on the (1000) and (0001) surfaces play an essential role in forming COOH* and CO* intermediates. Hence, such studies may help establish a practical, large-scale, Pb-free photocatalyst for the CO2 reduction process in the near future. The summary of all the studied catalysts has been presented in Table 2.

5. Summary and Outlook

Halide perovskites have been considered an advanced and high-performance material for many applications over the last few years. Their superior optoelectronic properties make them suitable for photocatalytic CO2 reduction. Pb-halide perovskites have proven to be the finest materials for CO2 conversion due to their high catalytic activity, high stability towards humidity, and long-term photostability. However, due to Pb’s high toxicity, the research focus is shifting towards the development of non-toxic, Pb-free halide perovskites. Therefore, in this review, the Pb and Pb-free halide perovskite’s recent progress towards photocatalytic CO2 reduction has been involved. We have covered the halide perovskites and their hybrid heterostructures/composites formed with metal co-catalyst, graphene, metal complexes, MOFs, and other 0D or 2D semiconductors. Although halide perovskites achieved significant success for photoreduction of CO2; still the catalytic efficiency is restricted in the µmol range, which keeps them away from large-scale use. Almost all the reported halide perovskites undergo CO2 reduction to form C1 products like CO and CH4. Therefore, more focus should be given to generate higher hydrocarbons, which are industrially important. Therefore, understanding the reaction mechanism towards the formation of higher-ordered organic chemicals via CO2 reduction of perovskite structure is crucial. Moreover, it is necessary to emphasize improving the structural and chemical stability of these materials. Improved catalytic efficiency and stability can be achieved by combining halide perovskite with other efficient semiconductors, improving the optical behavior, and charge separation ability. Therefore, developing sustainable, scalable, and low-cost halide perovskites will be a tremendous challenge for real applications. We believe that the process of unlocking more perovskite materials with improved optoelectronic features should continue to make them perfect for better CO2 reduction performance.

Author Contributions

C.B.H. and S.-I.I. conceptualized, wrote, and edited the manuscript. N.S.P. has revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful for the Ministry of Science and ICT and the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science and ICT for financial support under the grant numbers 2017R1E1A1A01074890 and 2015M1A2A2074670, respectively.

Acknowledgments

We gratefully acknowledge the Ministry of Science and ICT’s support and the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science. The authors also acknowledge the Flux Photon Corporation for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for selective photoreduction of CO2 into solar fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  2. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO–TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kim, K.; Razzaq, A.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid mesoporous Cu2ZnSnS4(CZTS)–TiO2 photocatalyst for efficient photocatalytic conversion of CO2 into CH4 under solar irradiation. RSC Adv. 2016, 6, 38964–38971. [Google Scholar] [CrossRef]
  4. Parayil, S.K.; Razzaq, A.; Park, S.-M.; Kim, H.R.; Grimes, C.A.; In, S.-I. Photocatalytic conversion of CO2 to hydrocarbon fuel using carbon and nitrogen co-doped sodium titanate nanotubes. Appl. Catal. A Gen. 2015, 498, 205–213. [Google Scholar] [CrossRef]
  5. Kim, H.R.; Razzaq, A.; Grimes, C.A.; In, S.-I. Heterojunction pnp Cu2O/S-TiO2/CuO: Synthesis and application to photocatalytic conversion of CO2 to methane. J. CO2 Util. 2017, 20, 91–96. [Google Scholar] [CrossRef]
  6. Sorcar, S.; Hwang, Y.; Lee, J.; Kim, H.; Grimes, K.M.; Grimes, C.A.; Jung, J.-W.; Cho, C.-H.; Majima, T.; Hoffmann, M.R.; et al. CO2, water, and sunlight to hydrocarbon fuels: A sustained sunlight to fuel (Joule-to-Joule) photoconversion efficiency of 1%. Energy Environ. Sci. 2019, 12, 2685–2696. [Google Scholar] [CrossRef] [Green Version]
  7. Ali, S.; Flores, M.C.; Razzaq, A.; Sorcar, S.; Hiragond, C.B.; Kim, H.R.; Park, Y.H.; Hwang, Y.; Kim, H.S.; Kim, H.; et al. Gas Phase Photocatalytic CO2 Reduction,“A Brief Overview for Benchmarking”. Catalysts 2019, 9, 727. [Google Scholar] [CrossRef] [Green Version]
  8. Hiragond, C.; Ali, S.; Sorcar, S.; In, S.-I. Hierarchical Nanostructured Photocatalysts for CO2 Photoreduction. Catalysts 2019, 9, 370. [Google Scholar] [CrossRef] [Green Version]
  9. Sorcar, S.; Thompson, J.; Hwang, Y.; Park, Y.H.; Majima, T.; Grimes, C.A.; Durrant, J.R.; In, S.-I. High-rate solar-light photoconversion of CO2 to fuel: Controllable transformation from C1 to C2 products. Energy Environ. Sci. 2018, 11, 3183–3193. [Google Scholar] [CrossRef]
  10. Sorcar, S.; Hwang, Y.; Grimes, C.A.; In, S.-I. Highly enhanced and stable activity of defect-induced titania nanoparticles for solar light-driven CO2 reduction into CH4. Mater. Today 2017, 20, 507–515. [Google Scholar] [CrossRef]
  11. Ali, S.; Lee, J.; Kim, H.; Hwang, Y.; Razzaq, A.; Jung, J.-W.; Cho, C.-H.; In, S.-I. Sustained, Photocatalytic CO2 Reduction to CH4 in a Continuous Flow Reactor by Earth-Abundant Materials: Reduced Titania-Cu2O Z-Scheme Heterostructures. Appl. Catal. B Environ. 2020, 279, 119344. [Google Scholar] [CrossRef]
  12. Kumar, A.; Thakur, P.R.; Sharma, G.; Naushad, M.; Rana, A.; Mola, G.T.; Stadler, F.J. Carbon nitride, metal nitrides, phosphides, chalcogenides, perovskites and carbides nanophotocatalysts for environmental applications. Environ. Chem. Lett. 2019, 17, 655–682. [Google Scholar] [CrossRef]
  13. Das, S.; Daud, W.M.A.W. A review on advances in photocatalysts towards CO2 conversion. Rsc Adv. 2014, 4, 20856–20893. [Google Scholar] [CrossRef]
  14. Razzaq, A.; Sinhamahapatra, A.; Kang, T.-H.; Grimes, C.A.; Yu, J.-S.; In, S.-I. Efficient solar light photoreduction of CO2 to hydrocarbon fuels via magnesiothermally reduced TiO2 photocatalyst. Appl. Catal. B Environ. 2017, 215, 28–35. [Google Scholar] [CrossRef]
  15. Akhundi, A.; Habibi-Yangjeh, A.; Abitorabi, M.; Rahim Pouran, S. Review on photocatalytic conversion of carbon dioxide to value-added compounds and renewable fuels by graphitic carbon nitride-based photocatalysts. Catal. Rev. 2019, 61, 595–628. [Google Scholar] [CrossRef]
  16. Wu, H.; Li, X.; Tung, C.; Wu, L. Semiconductor Quantum Dots: An Emerging Candidate for CO2 Photoreduction. Adv. Mater. 2019, 31, 1900709. [Google Scholar] [CrossRef] [PubMed]
  17. Jiao, X.; Zheng, K.; Liang, L.; Li, X.; Sun, Y.; Xie, Y. Fundamentals and challenges of ultrathin 2D photocatalysts in boosting CO2 photoreduction. Chem. Soc. Rev. 2020, 49, 6592–6604. [Google Scholar] [CrossRef]
  18. Nguyen, T.P.; Nguyen, D.L.T.; Nguyen, V.-H.; Le, T.-H.; Vo, D.-V.N.; Trinh, Q.T.; Bae, S.-R.; Chae, S.Y.; Kim, S.Y.; Le, Q. Van Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels. Nanomaterials 2020, 10, 337. [Google Scholar] [CrossRef] [Green Version]
  19. Kim, M.; Razzaq, A.; Kim, Y.K.; Kim, S.; In, S.-I. Synthesis and characterization of platinum modified TiO2-embedded carbon nanofibers for solar hydrogen generation. RSC Adv. 2014, 4, 51286–51293. [Google Scholar] [CrossRef]
  20. Razzaq, A.; In, S.-I. TiO2 Based Nanostructures for Photocatalytic CO2 Conversion to Valuable Chemicals. Micromachines 2019, 10, 326. [Google Scholar] [CrossRef] [Green Version]
  21. Zubair, M.; Razzaq, A.; Grimes, C.A.; In, S.-I. Cu2ZnSnS4 (CZTS)-ZnO: A noble metal-free hybrid Z-scheme photocatalyst for enhanced solar-spectrum photocatalytic conversion of CO2 to CH4. J. CO2 Util. 2017, 20, 301–311. [Google Scholar] [CrossRef]
  22. Razzaq, A.; Ali, S.; Asif, M.; In, S.-I. Layered Double Hydroxide (LDH) Based Photocatalysts: An Outstanding Strategy for Efficient Photocatalytic CO2 Conversion. Catalysts 2020, 10, 1185. [Google Scholar] [CrossRef]
  23. Hiragond, C.B.; Lee, J.; Kim, H.; Jung, J.-W.; Cho, C.-H.; In, S.-I. A novel N-doped graphene oxide enfolded reduced titania for highly stable and selective gas-phase photocatalytic CO2 reduction into CH4: An in-depth study on the interfacial charge transfer mechanism. Chem. Eng. J. 2020, 127978. [Google Scholar] [CrossRef]
  24. Zeng, S.; Kar, P.; Thakur, U.K.; Shankar, K. A review on photocatalytic CO2 reduction using perovskite oxide nanomaterials. Nanotechnology 2018, 29, 52001. [Google Scholar] [CrossRef] [PubMed]
  25. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide perovskite photovoltaics: Background, status, and future prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef] [PubMed]
  26. Johnsson, M.; Lemmens, P. Crystallography and chemistry of perovskites. Handb. Magn. Adv. Magn. Mater. 2007. [Google Scholar] [CrossRef] [Green Version]
  27. Xue, J.; Wang, R.; Yang, Y. The surface of halide perovskites from nano to bulk. Nat. Rev. Mater. 2020, 5, 807–827. [Google Scholar] [CrossRef]
  28. Teh, Y.W.; Chee, M.K.T.; Kong, X.Y.; Yong, S.-T.; Chai, S.-P. An insight into perovskite-based photocatalysts for artificial photosynthesis. Sustain. Energy Fuels 2020, 4, 973–984. [Google Scholar] [CrossRef]
  29. Hemminger, J.C.; Carr, R.; Somorjai, G.A. The photoassisted reaction of gaseous water and carbon dioxide adsorbed on the SrTiO3 (111) crystal face to form methane. Chem. Phys. Lett. 1978, 57, 100–104. [Google Scholar] [CrossRef]
  30. Luo, C.; Zhao, J.; Li, Y.; Zhao, W.; Zeng, Y.; Wang, C. Photocatalytic CO2 reduction over SrTiO3: Correlation between surface structure and activity. Appl. Surf. Sci. 2018, 447, 627–635. [Google Scholar] [CrossRef]
  31. Kwak, B.S.; Do, J.Y.; Park, N.-K.; Kang, M. Surface modification of layered perovskite Sr2TiO4 for improved CO2 photoreduction with H2O to CH4. Sci. Rep. 2017, 7, 1–15. [Google Scholar] [CrossRef] [PubMed]
  32. Li, P.; Ouyang, S.; Xi, G.; Kako, T.; Ye, J. The effects of crystal structure and electronic structure on photocatalytic H2 evolution and CO2 reduction over two phases of perovskite-structured NaNbO3. J. Phys. Chem. C 2012, 116, 7621–7628. [Google Scholar] [CrossRef] [Green Version]
  33. Wang, Y.; Liu, M.; Chen, W.; Mao, L.; Shangguan, W. Ag loaded on layered perovskite H2SrTa2O7 to enhance the selectivity of photocatalytic CO2 reduction with H2O. J. Alloys Compd. 2019, 786, 149–154. [Google Scholar] [CrossRef]
  34. Fan, J.; Jia, B.; Gu, M. Perovskite-based low-cost and high-efficiency hybrid halide solar cells. Photonics Res. 2014, 2, 111–120. [Google Scholar] [CrossRef] [Green Version]
  35. Kulkarni, S.A.; Mhaisalkar, S.G.; Mathews, N.; Boix, P.P. Perovskite nanoparticles: Synthesis, properties, and novel applications in photovoltaics and LEDs. Small Methods 2019, 3, 1800231. [Google Scholar] [CrossRef] [Green Version]
  36. Kubicek, M.; Bork, A.H.; Rupp, J.L.M. Perovskite oxides–a review on a versatile material class for solar-to-fuel conversion processes. J. Mater. Chem. A 2017, 5, 11983–12000. [Google Scholar] [CrossRef]
  37. Chen, P.; Ong, W.; Shi, Z.; Zhao, X.; Li, N. Pb-Based Halide Perovskites: Recent Advances in Photo (electro) catalytic Applications and Looking Beyond. Adv. Funct. Mater. 2020, 30, 1909667. [Google Scholar] [CrossRef]
  38. Tabish, A.; Varghese, A.M.; Wahab, M.A.; Karanikolos, G.N. Perovskites in the Energy Grid and CO2 Conversion: Current Context and Future Directions. Catalysts 2020, 10, 95. [Google Scholar] [CrossRef] [Green Version]
  39. Singh, M.; Sinha, I. Halide perovskite-based photocatalysis systems for solar-driven fuel generation. Sol. Energy 2020, 208, 296–311. [Google Scholar] [CrossRef]
  40. Shyamal, S.; Pradhan, N. Halide Perovskite Nanocrystal Photocatalysts for CO2 Reduction: Successes and Challenges. J. Phys. Chem. Lett. 2020, 11, 6921–6934. [Google Scholar] [CrossRef]
  41. Han, C.; Zhu, X.; Martin, J.S.; Lin, Y.; Spears, S.; Yan, Y. Recent Progress in Engineering Metal Halide Perovskites for Efficient Visible-Light-Driven Photocatalysis. ChemSusChem 2020, 13, 4005–4025. [Google Scholar] [CrossRef] [PubMed]
  42. 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]
  43. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO2 to ethanol. Appl. Catal. B Environ. 2020, 266, 118618. [Google Scholar] [CrossRef]
  44. Li, K.; Peng, B.; Peng, T. Recent advances in heterogeneous photocatalytic CO2 conversion to solar fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  45. Lingampalli, S.R.; Ayyub, M.M.; Rao, C.N.R. Recent progress in the photocatalytic reduction of carbon dioxide. ACS Omega 2017, 2, 2740–2748. [Google Scholar] [CrossRef] [Green Version]
  46. Qin, D.; Zhou, Y.; Wang, W.; Zhang, C.; Zeng, G.; Huang, D.; Wang, L.; Wang, H.; Yang, Y.; Lei, L. Recent advances in two-dimensional nanomaterials for photocatalytic reduction of CO2: Insights into performance, theories and perspective. J. Mater. Chem. A 2020, 8, 19156–19195. [Google Scholar] [CrossRef]
  47. Thompson, W.A.; Sanchez Fernandez, E.; Maroto-Valer, M.M. Review and Analysis of CO2 Photoreduction Kinetics. ACS Sustain. Chem. Eng. 2020, 8, 4677–4692. [Google Scholar] [CrossRef] [Green Version]
  48. Chang, X.; Wang, T.; Gong, J. CO 2 photo-reduction: Insights into CO 2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  49. Peng, C.; Reid, G.; Wang, H.; Hu, P. Perspective: Photocatalytic reduction of CO2 to solar fuels over semiconductors. J. Chem. Phys. 2017, 147, 30901. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, Y.; Mori, T.; Ye, J. Polymeric carbon nitrides: Semiconducting properties and emerging applications in photocatalysis and photoelectrochemical energy conversion. Sci. Adv. Mater. 2012, 4, 282–291. [Google Scholar] [CrossRef] [Green Version]
  51. 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]
  52. Kishan, K.Y.; Bisht, N.; Hiragond, C.; Dey, A.; Khanna, P.K.; More, P. V Room temperature thermoelectric performance of Methyl Ammonium Lead Iodide Perovskite and their MWCNT-PANI composites. Mater. Today Chem. 2020, 17, 100275. [Google Scholar] [CrossRef]
  53. Zhu, X.; Lin, Y.; San Martin, J.; Sun, Y.; Zhu, D.; Yan, Y. Lead halide perovskites for photocatalytic organic synthesis. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, Y.; Yang, M.; Li, Z.; Crisp, R.; Zhu, K.; Beard, M.C. Comparison of recombination dynamics in CH3NH3PbBr3 and CH3NH3PbI3 perovskite films: Influence of exciton binding energy. J. Phys. Chem. Lett. 2015, 6, 4688–4692. [Google Scholar] [CrossRef]
  55. Wu, Y.; Wang, P.; Zhu, X.; Zhang, Q.; Wang, Z.; Liu, Y.; Zou, G.; Dai, Y.; Whangbo, M.; Huang, B. Composite of CH3NH3PbI3 with reduced graphene oxide as a highly efficient and stable visible-light photocatalyst for hydrogen evolution in aqueous HI solution. Adv. Mater. 2018, 30, 1704342. [Google Scholar] [CrossRef]
  56. Wu, Y.; Wang, P.; Guan, Z.; Liu, J.; Wang, Z.; Zheng, Z.; Jin, S.; Dai, Y.; Whangbo, M.-H.; Huang, B. Enhancing the Photocatalytic Hydrogen Evolution Activity of Mixed-Halide Perovskite CH3NH3PbBr3–xIx Achieved by Bandgap Funneling of Charge Carriers. ACS Catal. 2018, 8, 10349–10357. [Google Scholar] [CrossRef]
  57. Cardenas-Morcoso, D.; Gualdrón-Reyes, A.F.; Ferreira Vitoreti, A.B.; García-Tecedor, M.; Yoon, S.J.; Solis de la Fuente, M.; Mora-Seró, I.; Gimenez, S. Photocatalytic and Photoelectrochemical Degradation of Organic Compounds with All-Inorganic Metal Halide Perovskite Quantum Dots. J. Phys. Chem. Lett. 2019, 10, 630–636. [Google Scholar] [CrossRef] [Green Version]
  58. Gao, G.; Xi, Q.; Zhou, H.; Zhao, Y.; Wu, C.; Wang, L.; Guo, P.; Xu, J. Novel inorganic perovskite quantum dots for photocatalysis. Nanoscale 2017, 9, 12032–12038. [Google Scholar] [CrossRef]
  59. Mu, Y.; Zhang, W.; Guo, X.; Dong, G.; Zhang, M.; Lu, T. 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]
  60. Wu, L.; Mu, Y.; Guo, X.; Zhang, W.; Zhang, Z.; Zhang, M.; Lu, T. Encapsulating Perovskite Quantum Dots in Iron-Based Metal–Organic Frameworks (MOFs) for Efficient Photocatalytic CO2 Reduction. Angew. Chemie Int. Ed. 2019, 58, 9491–9495. [Google Scholar] [CrossRef]
  61. Hou, J.; Cao, S.; Wu, Y.; Gao, Z.; Liang, F.; Sun, Y.; Lin, Z.; Sun, L. Inorganic colloidal perovskite quantum dots for robust solar CO2 reduction. Chem. Eur. J. 2017, 23, 9481–9485. [Google Scholar] [CrossRef] [PubMed]
  62. Guo, S.-H.; Zhou, J.; Zhao, X.; Sun, C.-Y.; You, S.-Q.; Wang, X.-L.; Su, Z.-M. Enhanced CO2 photoreduction via tuning halides in perovskites. J. Catal. 2019, 369, 201–208. [Google Scholar] [CrossRef]
  63. 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]
  64. 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]
  65. Liu, Y.-W.; Guo, S.-H.; You, S.-Q.; Sun, C.-Y.; Wang, X.-L.; Zhao, L.; Su, Z.-M. Mn-doped CsPb(Br/Cl)3 mixed-halide perovskites for CO2 photoreduction. Nanotechnology 2020, 31, 215605. [Google Scholar] [CrossRef]
  66. Dong, G.-X.; Zhang, W.; Mu, Y.-F.; Su, K.; Zhang, M.; Lu, T.-B. A halide perovskite as a catalyst to simultaneously achieve efficient photocatalytic CO2 reduction and methanol oxidation. Chem. Commun. 2020, 56, 4664–4667. [Google Scholar] [CrossRef]
  67. Chen, Y.-X.; Xu, Y.-F.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B. Solvent selection and Pt decoration towards enhanced photocatalytic CO2 reduction over CsPbBr3 perovskite single crystals. Sustain. Energy Fuels 2020, 4, 2249–2255. [Google Scholar] [CrossRef]
  68. Zhu, J.; Zhu, Y.; Huang, J.; Hou, L.; Shen, J.; Li, C. Synthesis of monodisperse water-stable surface Pb-rich CsPbCl3 nanocrystals for efficient photocatalytic CO2 reduction. Nanoscale 2020, 12, 11842–11846. [Google Scholar] [CrossRef]
  69. Hiragond, C.B.; Kim, H.; Lee, J.; Sorcar, S.; Erkey, C.; In, S.-I. Electrochemical CO2 Reduction to CO Catalyzed by 2D Nanostructures. Catalysts 2020, 10, 98. [Google Scholar] [CrossRef] [Green Version]
  70. Ali, S.; Razzaq, A.; In, S.-I. Development of graphene based photocatalysts for CO2 reduction to C1 chemicals: A brief overview. Cat. Today 2019, 335, 39–54. [Google Scholar] [CrossRef]
  71. 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]
  72. Kumar, S.; Regue, M.; Isaacs, M.A.; Freeman, E.; Eslava, S. All-Inorganic CsPbBr3 Nanocrystals: Gram-Scale Mechanochemical Synthesis and Selective Photocatalytic CO2 Reduction to Methane. ACS Appl. Energy Mater. 2020, 3, 4509–4522. [Google Scholar] [CrossRef]
  73. Wang, X.; Li, K.; He, J.; Yang, J.; Dong, F.; Mai, W.; Zhu, M. Defect in reduced graphene oxide tailored selectivity of photocatalytic CO2 reduction on Cs4PbBr6 pervoskite hole-in-microdisk structure. Nano Energy 2020, 78, 105388. [Google Scholar] [CrossRef]
  74. Mu, Y.; Zhang, W.; Dong, G.; Su, K.; Zhang, M.; Lu, T. Ultrathin and Small-Size Graphene Oxide as an Electron Mediator for Perovskite-Based Z-Scheme System to Significantly Enhance Photocatalytic CO2 Reduction. Small 2020, 16, 2002140. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, J.; Wang, J.; Li, N.; Du, X.; Ma, J.; He, C.; Li, Z. Direct Z-Scheme 0D/2D Heterojunction of CsPbBr3 Quantum Dots/Bi2WO6 Nanosheets for Efficient Photocatalytic CO2 Reduction. ACS Appl. Mater. Interfaces 2020, 12, 31477–31485. [Google Scholar] [CrossRef] [PubMed]
  76. Xu, Y.-F.; Yang, M.-Z.; Chen, H.-Y.; Liao, J.-F.; Wang, X.-D.; Kuang, D.-B. Enhanced solar-driven gaseous CO2 conversion by CsPbBr3 nanocrystal/Pd nanosheet Schottky-junction photocatalyst. ACS Appl. Energy Mater. 2018, 1, 5083–5089. [Google Scholar] [CrossRef]
  77. Pan, A.; Ma, X.; Huang, S.; Wu, Y.; Jia, M.; Shi, Y.; Liu, Y.; Wangyang, P.; He, L.; Liu, Y. CsPbBr3 Perovskite Nanocrystal Grown on MXene Nanosheets for Enhanced Photoelectric Detection and Photocatalytic CO2 Reduction. J. Phys. Chem. Lett. 2019, 10, 6590–6597. [Google Scholar] [CrossRef] [Green Version]
  78. Xu, Y.; Wang, X.; Liao, J.; Chen, B.; Chen, H.; Kuang, D. Amorphous-TiO2-Encapsulated CsPbBr3 Nanocrystal Composite Photocatalyst with Enhanced Charge Separation and CO2 Fixation. Adv. Mater. Interfaces 2018, 5, 1801015. [Google Scholar] [CrossRef]
  79. 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, 1–9. [Google Scholar] [CrossRef]
  80. Lu, Y.; Ma, Y.; Zhang, T.; Yang, Y.; Wei, L.; Chen, Y. Monolithic 3D cross-linked polymeric graphene materials and the likes: Preparation and their redox catalytic applications. J. Am. Chem. Soc. 2018, 140, 11538–11550. [Google Scholar] [CrossRef]
  81. Jiang, Y.; Liao, J.-F.; Xu, Y.-F.; Chen, H.-Y.; Wang, X.-D.; Kuang, D.-B. Hierarchical CsPbBr3 nanocrystal-decorated ZnO nanowire/macroporous graphene hybrids for enhancing charge separation and photocatalytic CO2 reduction. J. Mater. Chem. A 2019, 7, 13762–13769. [Google Scholar] [CrossRef]
  82. Lin, Z.; Wang, X. Nanostructure engineering and doping of conjugated carbon nitride semiconductors for hydrogen photosynthesis. Angew. Chemie 2013, 125, 1779–1782. [Google Scholar] [CrossRef]
  83. Ou, M.; Tu, W.; Yin, S.; Xing, W.; Wu, S.; Wang, H.; Wan, S.; Zhong, Q.; Xu, R. Amino-assisted anchoring of CsPbBr3 perovskite quantum dots on porous g-C3N4 for enhanced photocatalytic CO2 reduction. Angew. Chemie 2018, 130, 13758–13762. [Google Scholar] [CrossRef]
  84. Guo, X.-X.; Tang, S.-F.; Mu, Y.-F.; Wu, L.-Y.; Dong, G.-X.; Zhang, M. Engineering a CsPbBr3-based nanocomposite for efficient photocatalytic CO2 reduction: Improved charge separation concomitant with increased activity sites. RSC Adv. 2019, 9, 34342–34348. [Google Scholar] [CrossRef] [Green Version]
  85. Kong, Z.-C.; Liao, J.-F.; Dong, Y.-J.; Xu, Y.-F.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. Core@shell CsPbBr3@Zeolitic imidazolate framework nanocomposite for efficient photocatalytic CO2 reduction. ACS Energy Lett. 2018, 3, 2656–2662. [Google Scholar] [CrossRef]
  86. Wan, S.; Ou, M.; Zhong, Q.; Wang, X. Perovskite-type CsPbBr3 quantum dots/UiO-66 (NH2) nanojunction as efficient visible-light-driven photocatalyst for CO2 reduction. Chem. Eng. J. 2019, 358, 1287–1295. [Google Scholar] [CrossRef]
  87. Hawecker, J.; Lehn, J.-M.; Ziessel, R. Efficient photochemical reduction of CO2 to CO by visible light irradiation of systems containing Re(bipy)(CO)3X or Ru(bipy)32+–Co2+ combinations as homogeneous catalysts. J. Chem. Soc. Chem. Commun. 1983, 536–538. [Google Scholar] [CrossRef]
  88. Hawecker, J.; Lehn, J.; Ziessel, R. Photochemical and electrochemical reduction of carbon dioxide to carbon monoxide mediated by (2, 2′-bipyridine) tricarbonylchlororhenium (I) and related complexes as homogeneous catalysts. Helv. Chim. Acta 1986, 69, 1990–2012. [Google Scholar] [CrossRef]
  89. Hori, H.; Johnson, F.P.A.; Koike, K.; Ishitani, O.; Ibusuki, T. Efficient photocatalytic CO2 reduction using [Re(bpy)(CO)3{P(OEt)3}]+. J. Photochem. Photobiol. A Chem. 1996, 96, 171–174. [Google Scholar] [CrossRef]
  90. Asai, Y.; Katsuragi, H.; Kita, K.; Tsubomura, T.; Yamazaki, Y. Photocatalytic CO2 reduction using metal complexes in various ionic liquids. Dalt. Trans. 2020, 49, 4277–4292. [Google Scholar] [CrossRef]
  91. Arias-Rotondo, D.M.; McCusker, J.K. The photophysics of photoredox catalysis: A roadmap for catalyst design. Chem. Soc. Rev. 2016, 45, 5803–5820. [Google Scholar] [CrossRef] [PubMed]
  92. Kong, Z.-C.; Zhang, H.-H.; Liao, J.-F.; Dong, Y.-J.; Jiang, Y.; Chen, H.-Y.; Kuang, D.-B. Immobilizing Re(CO)3Br(dcbpy) Complex on CsPbBr3 Nanocrystal for Boosted Charge Separation and Photocatalytic CO2 Reduction. Sol. RRL 2020, 4, 1900365. [Google Scholar] [CrossRef]
  93. Chen, Z.; Hu, Y.; Wang, J.; Shen, Q.; Zhang, Y.; Ding, C.; Bai, Y.; Jiang, G.; Li, Z.; Gaponik, N. Boosting Photocatalytic CO2 Reduction on CsPbBr3 Perovskite Nanocrystals by Immobilizing Metal Complexes. Chem. Mater. 2020, 32, 1517–1525. [Google Scholar] [CrossRef]
  94. Que, M.; Zhao, Y.; Pan, L.; Yang, Y.; He, Z.; Yuan, H.; Chen, J.; Zhu, G. Colloidal Formamidinium Lead Bromide Quantum Dots for Photocatalytic CO2 Reduction. Mater. Lett. 2020, 282, 128695. [Google Scholar] [CrossRef]
  95. 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]
  96. Li, J.; Cao, H.-L.; Jiao, W.-B.; Wang, Q.; Wei, M.; Cantone, I.; Lü, J.; Abate, A. Biological impact of lead from halide perovskites reveals the risk of introducing a safe threshold. Nat. Commun. 2020, 11, 1–5. [Google Scholar] [CrossRef] [Green Version]
  97. Flora, G.; Gupta, D.; Tiwari, A. Toxicity of lead: A review with recent updates. Interdiscip. Toxicol. 2012, 5, 47–58. [Google Scholar] [CrossRef]
  98. Chu, L.; Ahmad, W.; Liu, W.; Yang, J.; Zhang, R.; Sun, Y.; Yang, J.; Li, X. Lead-free halide double perovskite materials: A new superstar toward green and stable optoelectronic applications. Nano-Micro Lett. 2019, 11, 16. [Google Scholar] [CrossRef] [Green Version]
  99. Chakraborty, S.; Xie, W.; Mathews, N.; Sherburne, M.; Ahuja, R.; Asta, M.; Mhaisalkar, S.G. Rational design: A high-throughput computational screening and experimental validation methodology for lead-free and emergent hybrid perovskites. ACS Energy Lett. 2017, 2, 837–845. [Google Scholar] [CrossRef]
  100. Kamat, P.V.; Bisquert, J.; Buriak, J. Lead-free perovskite solar cells. ACS Energy Lett. 2017, 2, 904–905. [Google Scholar] [CrossRef] [Green Version]
  101. Fan, Q.; Biesold-McGee, G.V.; Ma, J.; Xu, Q.; Pan, S.; Peng, J.; Lin, Z. Lead-Free Halide Perovskite Nanocrystals: Crystal Structures, Synthesis, Stabilities, and Optical Properties. Angew. Chemie Int. Ed. 2020, 59, 1030–1046. [Google Scholar] [CrossRef]
  102. Volonakis, G.; Filip, M.R.; Haghighirad, A.A.; Sakai, N.; Wenger, B.; Snaith, H.J.; Giustino, F. Lead-free halide double perovskites via heterovalent substitution of noble metals. J. Phys. Chem. Lett. 2016, 7, 1254–1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Slavney, A.H.; Hu, T.; Lindenberg, A.M.; Karunadasa, H.I. A bismuth-halide double perovskite with long carrier recombination lifetime for photovoltaic applications. J. Am. Chem. Soc. 2016, 138, 2138–2141. [Google Scholar] [CrossRef]
  104. Zhou, L.; Xu, Y.; Chen, B.; Kuang, D.; Su, C. Synthesis and Photocatalytic Application of Stable Lead-Free Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14, 1703762. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, Y.; Huang, H.; Zhang, Z.; Wang, C.; Yang, Y.; Li, Q.; Xu, D. Lead-Free Perovskite Cs2AgBiBr6@g-C3N4 Z-scheme System for Improving CH4 Production in Photocatalytic CO2 Reduction. Appl. Catal. B Environ. 2020, 282, 119570. [Google Scholar] [CrossRef]
  106. Wang, X.-D.; Huang, Y.-H.; Liao, J.-F.; Jiang, Y.; Zhou, L.; Zhang, X.-Y.; Chen, H.-Y.; Kuang, D.-B. In situ construction of a Cs2SnI6 perovskite nanocrystal/SnS2 nanosheet heterojunction with boosted interfacial charge transfer. J. Am. Chem. Soc. 2019, 141, 13434–13441. [Google Scholar] [CrossRef] [PubMed]
  107. Bhosale, S.S.; Kharade, A.K.; Jokar, E.; Fathi, A.; Chang, S.; Diau, E.W.-G. Mechanism of Photocatalytic CO2 Reduction by Bismuth-Based Perovskite Nanocrystals at the Gas-Solid Interface. J. Am. Chem. Soc. 2019, 141, 20434–20442. [Google Scholar] [CrossRef]
  108. Lu, C.; Itanze, D.S.; Aragon, A.G.; Ma, X.; Li, H.; Ucer, K.B.; Hewitt, C.; Carroll, D.L.; Williams, R.T.; Qiu, Y. Synthesis of lead-free Cs3Sb2Br9 perovskite alternative nanocrystals with enhanced photocatalytic CO2 reduction activity. Nanoscale 2020, 12, 2987–2991. [Google Scholar] [CrossRef]
Figure 1. (a–d) CsPbBr3 quantum dots (QDs) TEM images showing the difference in particle size, (e) photocatalytic CO2 reduction using an optimized sample of CsPbBr3, (f) particle size effect on the CO2 reduction activity, (g) Time-resolved photoluminescence (TRPL) decay of different samples, and (h) band diagram showing mechanism of CO2 reduction to chemical fuels. Reproduced from [61], with permission from John Wiley and Sons, 2017.
Figure 1. (a–d) CsPbBr3 quantum dots (QDs) TEM images showing the difference in particle size, (e) photocatalytic CO2 reduction using an optimized sample of CsPbBr3, (f) particle size effect on the CO2 reduction activity, (g) Time-resolved photoluminescence (TRPL) decay of different samples, and (h) band diagram showing mechanism of CO2 reduction to chemical fuels. Reproduced from [61], with permission from John Wiley and Sons, 2017.
Nanomaterials 10 02569 g001
Figure 2. The free energy diagram showing the pathway of CO2 reduction on pristine and Co/Fe doped CsPbBr3. Reproduced from [63], Royal Society of Chemistry, 2019.
Figure 2. The free energy diagram showing the pathway of CO2 reduction on pristine and Co/Fe doped CsPbBr3. Reproduced from [63], Royal Society of Chemistry, 2019.
Nanomaterials 10 02569 g002
Figure 3. The comparative (a) PL spectra, (b) PL decay, (c) transient absorption kinetic plots (at an excitation wavelength of 400 nm) among CsPbBr3 NC and composite samples, (d) schematic illustration and band alignment of CsPbBr3/Pd composite for CO2 reduction, (e) photocatalytic CO2 reduction performance, and (f) quantum efficiency of different samples. Reproduced from [76], with permission from Royal Society of Chemistry, 2019.
Figure 3. The comparative (a) PL spectra, (b) PL decay, (c) transient absorption kinetic plots (at an excitation wavelength of 400 nm) among CsPbBr3 NC and composite samples, (d) schematic illustration and band alignment of CsPbBr3/Pd composite for CO2 reduction, (e) photocatalytic CO2 reduction performance, and (f) quantum efficiency of different samples. Reproduced from [76], with permission from Royal Society of Chemistry, 2019.
Nanomaterials 10 02569 g003
Figure 4. (a) An in−situ synthetic procedure for CsPbBr3/MXene, (bd) TEM and HR-TEM images of CsPbBr3/MXene-20 composite and (ei) EDX elemental mapping for the respective elements. Reproduced from [77], with permission from American Chemical Society, 2019.
Figure 4. (a) An in−situ synthetic procedure for CsPbBr3/MXene, (bd) TEM and HR-TEM images of CsPbBr3/MXene-20 composite and (ei) EDX elemental mapping for the respective elements. Reproduced from [77], with permission from American Chemical Society, 2019.
Nanomaterials 10 02569 g004
Figure 5. (ac) Ex-situ and in-situ XPS of TiO2 (T) and TiO2/CsPbBr3 (TC2) samples (TC2-UV is in-situ XPS under UV light irradiation), and electrostatic potentials for (d) anatase TiO2 (101), (e) rutile TiO2 (110) and (f) CsPbBr3 (001) facets. Reproduced from [79], with permission from Springer Nature, 2020.
Figure 5. (ac) Ex-situ and in-situ XPS of TiO2 (T) and TiO2/CsPbBr3 (TC2) samples (TC2-UV is in-situ XPS under UV light irradiation), and electrostatic potentials for (d) anatase TiO2 (101), (e) rutile TiO2 (110) and (f) CsPbBr3 (001) facets. Reproduced from [79], with permission from Springer Nature, 2020.
Nanomaterials 10 02569 g005
Figure 6. (a) Schematic illustration of synthesis and utilization of CsPbBr3/ZIFs for CO2 reduction, (b,c) TEM of CsPbBr3@ZIF-8, (d) HAADF-STEM of CsPbBr3@ZIF-27, (e) TEM of CsPbBr3@ZIF-27, (f) CO2 reduction results for pristine and composite, and (g) Stability test for CsPbBr3@ZIF-67. Reproduced from [85], with permission from American chemical society, 2018.
Figure 6. (a) Schematic illustration of synthesis and utilization of CsPbBr3/ZIFs for CO2 reduction, (b,c) TEM of CsPbBr3@ZIF-8, (d) HAADF-STEM of CsPbBr3@ZIF-27, (e) TEM of CsPbBr3@ZIF-27, (f) CO2 reduction results for pristine and composite, and (g) Stability test for CsPbBr3@ZIF-67. Reproduced from [85], with permission from American chemical society, 2018.
Nanomaterials 10 02569 g006
Figure 7. (a,b) The XRD patterns, (c) UV-Vis/PL spectra and (d,e) TEM images of FAPbBr3, and CsPbBr3, (f) Results of comparative photocatalytic CO2 reduction in FAPbBr3 and CsPbBr3, and (g) reusability test of FAPbBr3. Reproduced from [94], with permission from Elsevier, 2020.
Figure 7. (a,b) The XRD patterns, (c) UV-Vis/PL spectra and (d,e) TEM images of FAPbBr3, and CsPbBr3, (f) Results of comparative photocatalytic CO2 reduction in FAPbBr3 and CsPbBr3, and (g) reusability test of FAPbBr3. Reproduced from [94], with permission from Elsevier, 2020.
Nanomaterials 10 02569 g007
Figure 8. (a) XPS, (b) FTIR, and (c,d) TGA analysis of Cs2AgBiBr6 before and after the washing. Reproduced from [104], with permission from John Wiley and Sons, 2018.
Figure 8. (a) XPS, (b) FTIR, and (c,d) TGA analysis of Cs2AgBiBr6 before and after the washing. Reproduced from [104], with permission from John Wiley and Sons, 2018.
Nanomaterials 10 02569 g008
Figure 9. (a) CO and CH4 formation, (b) times course product formation in different samples of Cs2AgBiBr6, and (c) Schematic illustration of CO2 reduction on the surface of Cs2AgBiBr6. Reproduced from [104], with permission from John Wiley and Sons, 2018.
Figure 9. (a) CO and CH4 formation, (b) times course product formation in different samples of Cs2AgBiBr6, and (c) Schematic illustration of CO2 reduction on the surface of Cs2AgBiBr6. Reproduced from [104], with permission from John Wiley and Sons, 2018.
Nanomaterials 10 02569 g009
Figure 10. (a) Schematic illustration of the synthesis of Bi−based perovskites by top-down method, (b) Time−dependent CH4 production in Bi−based perovskites, and (c) yields of CO and CH4 production in different samples. Reproduced from [107], with permission from American Chemical Society, 2019.
Figure 10. (a) Schematic illustration of the synthesis of Bi−based perovskites by top-down method, (b) Time−dependent CH4 production in Bi−based perovskites, and (c) yields of CO and CH4 production in different samples. Reproduced from [107], with permission from American Chemical Society, 2019.
Nanomaterials 10 02569 g010
Table 1. The catalytic pathway of CO2 reduction towards various main products with their redox potentials (at pH = 7).
Table 1. The catalytic pathway of CO2 reduction towards various main products with their redox potentials (at pH = 7).
No.ReactionE0redox Vs. NHEProduct
1CO2 + e → CO2●−−1.90 VCO2●−
2CO2 + 2H+ + 2e → CO + H2O−0.53 VCarbon monoxide
3CO2 + 2H+ + 2e → HCOOH−0.61 VFormic acid
4CO2 + 4H+ + 4e → HCHO + H2O−0.48 VFormaldehyde
5CO2 + 6H+ + 6e → CH3OH + H2O−0.38 VMethanol
6CO2 + 8H+ + 8e → CH4 + 2H2O−0.24 VMethane
72CO2 + 8H+ + 8e → CH3COOH + 2H2O−0.31 VAcetic acid
82CO2 + 14H+ + 14e → C2H6 + 4H2O−0.51 VEthane
Table 2. The progress of Pb and Pb-free perovskites for CO2 reduction with details including catalysts [reference], reaction medium, light source, product yield and time, and stability (*EA = ethyl acetate, MeOH = methanol, AN = acetonitrile, IPN = isopropanol, TCM = trichloromethane, OC = octadecene).
Table 2. The progress of Pb and Pb-free perovskites for CO2 reduction with details including catalysts [reference], reaction medium, light source, product yield and time, and stability (*EA = ethyl acetate, MeOH = methanol, AN = acetonitrile, IPN = isopropanol, TCM = trichloromethane, OC = octadecene).
NoCatalystMediumLight SourceProduct, Yield, and Reaction TimeCatalytic Stability
Lead-based halide perovskites
1Fe/CH3NH3PbI3 (MAPbI3) QDs
[60]
EA/H2O300 W Xe-lamp with standard 400 nm filterCO + CH4, 1559 µmol g−1
CO (34%) and CH4 (66%)
80 h
2CsPbBr3 QDs
[61]
EA/H2O300 W Xe-lamp with standard AM 1.5 filterCO, 20.9 µmol g−1
(average electron yield)
8 h
3CsPb(Br0.5/Cl0.5)3
[62]
EA300 W Xe-lamp with AM 1.5 filterCO, 767 µmol g−1 (9 h)
CH4, 108 µmol g−1 (9 h)
9 h
4Co- and Fe-CsPbBr3
[63]
A theoretical study, DFT calculations with DMol3 program
5Fe(II)-CsPbBr3
[64]
EA/H2O300 W Xe-lamp (150 mW cm−2 light intensity)CO, 6.1 µmol g−1 h−1 (3 h)
CH4, 3.2 µmol g−1 h−1 (3 h)
-
6Mn/CsPb(Br/Cl)3 [65]EA300 W Xe-lamp with AM 1.5 filterCO, 1917 µmol g−1 (9 h)
CH4, 82 µmol g−1 (9 h)
9 h
7Co-CsPbBr3/Cs4PbBr6
[59]
H2OXe-lamp irradiation
with a 400 nm filter (100 mW cm−2 light intensity)
CO, 239 µmol g−1 (20 h)
CH4, 7 µmol g−1 (20 h)
-
8Co-CsPbBr3/Cs4PbBr6
[66]
AN/H2O/MeOH300 W Xe-lamp (light intensity of 100 mW m−2)CO, 1835 µmol g−1 (15 h)-
9Pt/CsPbBr3
[67]
EA150 W Xe-lamp with 380 nm cut off filter CO, 5.6 µmol g−1 h−130 h
10Ni and Mn-doped CsPbCl3 NCs
[68]
CO2/ H2O300 W Xe-lamp with AM 1.5 filterNi = CO, 169.37 μmol g−1 h−1
Mn = 152.49 μmol g−1 h−1
6 h
(3 runs)
11CsPbBr3 QDs/GO
[71]
EA100 W Xe-lamp with AM 1.5 filterCO, 58.7 µmol g−1 (12 h)
CH4, 29.6 µmol g−1 (12 h)
H2, 1.58 µmol g−1 (12 h)
12 h
12Cs4PbBr6/rGO
[73]
EA/H2O300 W Xe-lamp with 420 nm filter (light intensity, 100 mW cm−2)CO, 11.4 μmol g−1 h−160 h
13Cu-RGO-CsPbBr3
[72]
CO2/H2OXe-lamp irradiation
with a 400 nm filter
CH4 12.7 μmol g–1 h–1 (4 h)12 h
(3 cycles)
14CsPbBr3/USGO/α-Fe2O3
[74]
ACN/H2O300 W Xe-lamp with 420 nm filter (light intensity, 100 mW cm−2)CO, 73.8 μmol g−1 h−12 cycles
15CsPbBr3/Bi2WO6
[75]
300 W Xe-lamp with 420 nm filter (100 mW cm−2 light intensity)CH4/CO, 503 μmol g–14 Runs
(pristine samples)
16CsPbBr3/Pd Nanosheet
[76]
H2O vaporA 150 W Xe-lamp (Zolix) equipped with an AM 1.5 G and 420 nm optical filter (100 mW cm−2 light intensity)CO, 12.63 µmol g−1 (3 h)
CH4, 10.41 µmol g−1 (3 h)
Electron consumption rate, 101.39 µmol g−1 (3 h)
-
17CsPbBr3/MXene Nanosheets
[77]
EA300 W Xe-lamp with 420 nm cut-off filterCO, 26.32 µmol g−1 h−1
CH4, 7.25 µmol g−1 h−1
-
18Amorphous-TiO2/CsPbBr3 NCs
[78]
EA/IPN150 W Xe-lamp with an AM 1.5 G filterCO, 11.71 μmol g −1
CH4, 20.15 μmol g −1
H2, 4.38 μmol g −1
30 h
19TiO2/CsPbBr3
[79]
ACN/H2O300 W Xe-arc lampCO, 9.02 μmol g–1 h–116 h
20CsPbBr3 NCs- ZnO nanowire/graphene
[81]
CO2/H2OA 150 W Xe-lamp with an AM 1.5 G and 420 nm optical filter (100 mW cm−2 light intensity)CH4, 6.29 µmol g−1 h−1 (3 h)
CO, 0.8 µmol g−1 h−1 (3 h)
Photoelectron consumption rate, 52.02 µmol g−1 h−1 (3 h)
4 cycles
21CsPbBr3 QDs/g-C3N4
[83]
ACN/H2O300 W Xe-lamp with a 420 nm cut-off filterCO, 149 μmol g−1 h−13 Runs
22CsPbBr3/g-C3N4 containing TiO species
[84]
EA/H2OXe-lamp with a 400 nm cut off filter (100 mW cm−2 light intensity) CO, 129 μmol g−1 (10 h) -
23CsPbBr3@Zeolitic Imidazolate
[85]
CO2/ H2O100 W Xe-lamp with AM 1.5 G filter (light intensity was 150 mW cm−2)The electron consumption rate for CH4, 29.630 μmol g–1 h–1 (3 h)6 cycles
24CsPbBr3 QDs/UiO-66(NH2) nanojunction
[86]
EA/H2O300 W Xe-lamp with a 420 nm UV-cut filterCO, 98.57 μmo g−1
CH4, 3.08 μmol g−1 (12 h)
3 cycles
25[Ni(terpy)2]2+ (Ni(tpy)) CsPbBr3 NCs
[93]
EA/H2O300 W Xe-lamp (Solaredge 700, 100 mW cm−2), λ > 400 nmCO + CH4, 1724 μmol g−1
(4 h)
16 h
26CsPbBr3-Re(CO)3Br(dcbpy) (dcbpy¼4,4′- dicarboxy-2,2′-bipyridine)
[92]
Toluene/IPN150 W Xe-lamp (≥420 nm)CO, 509.14 μmol g−1 (15 h)-
27FAPbBr3 QDs
[94]
EA/H2O300 W Xe-lamp (light intensity of 100 mW cm−2)CO, 181.25 μmol g−1 h−1-
Lead-free halide perovskites
28Cs2AgBiBr6
[104]
EA100 W Xe-lamp with an AM 1.5 G filter CO, 14.1 μmol g−1 (6 h)
CH4, 9.6 μmol g−1 (6 h)
6 h
29Cs2AgBiBr6@g-C3N4 Z-scheme
[105]
EA/MeOHXe-lamp (80 mW cm−2 light intensity) CO + CH4, 2.0 μmol g−1 h−112 h
30Cs2SnI6/SnS2 Nanosheet
[106]
CO2/H2O/MeOH150 mW cm−2 Xe-lamp with 400 nm filterCH4, 6.09 μmol g−1 (3 h)9 h
(3 cycles)
31Bi-based perovskite NCs [107]
  • Rb3Bi2I2
  • Cs3Bi2I9
  • MA3Bi2T9
TCM32 W UV-lamp, 305 nmCs3Bi2I9 = CO, 77.6 μmol g−1
and CH4, 14.9 μmol g−1 (10 h)
Rb3Bi2I9 = CO, 18.2 μmol g−1 and CH4, 17.0 μmol g−1 (10 h)
MA3Bi2I9 = CO, 7.2 μmol g−1 and CH4, 9.8 μmol g−1 (10 h)
-
32Cs3Sb2Br9
[108]
Dried OC300 W Xe-lamp with AM 1.5 irradiationCO, 510 μmol g−1 (4 h)9 h
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hiragond, C.B.; Powar, N.S.; In, S.-I. Recent Developments in Lead and Lead-Free Halide Perovskite Nanostructures towards Photocatalytic CO2 Reduction. Nanomaterials 2020, 10, 2569. https://doi.org/10.3390/nano10122569

AMA Style

Hiragond CB, Powar NS, In S-I. Recent Developments in Lead and Lead-Free Halide Perovskite Nanostructures towards Photocatalytic CO2 Reduction. Nanomaterials. 2020; 10(12):2569. https://doi.org/10.3390/nano10122569

Chicago/Turabian Style

Hiragond, Chaitanya B., Niket S. Powar, and Su-Il In. 2020. "Recent Developments in Lead and Lead-Free Halide Perovskite Nanostructures towards Photocatalytic CO2 Reduction" Nanomaterials 10, no. 12: 2569. https://doi.org/10.3390/nano10122569

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