Next Article in Journal
Coke Deposition and Structural Changes of Pellet V2O5/NaY-SiO2 in Air Regeneration: The Effects of Temperature on Regeneration
Previous Article in Journal
Electrocatalysts Based on Novel Carbon Forms for the Oxidation of Sulphite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol

Department of Industrial and Engineering Chemistry, Institute of Chemical Technology, Mumbai-Marathwada Campus, Jalna 431203, India
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(1), 94; https://doi.org/10.3390/catal12010094
Submission received: 6 December 2021 / Revised: 8 January 2022 / Accepted: 11 January 2022 / Published: 14 January 2022
(This article belongs to the Section Photocatalysis)

Abstract

:
Constantly increasing hydrocarbon fuel combustion along with high levels of carbon dioxide emissions has given rise to a global energy crisis and environmental alterations. Photocatalysis is an effective technique for addressing this energy and environmental crisis. Clean and renewable solar energy is a very favourable path for photocatalytic CO2 reduction to value-added products to tackle problems of energy and the environment. The synthesis of various products such as CH4, CH3OH, CO, EtOH, etc., has been expanded through the photocatalytic reduction of CO2. Among these products, methanol is one of the most important and highly versatile chemicals widely used in industry and in day-to-day life. This review emphasizes the recent progress of photocatalytic CO2 hydrogenation to CH3OH. In particular, Metal organic frameworks (MOFs), mixed-metal oxide, carbon, TiO2 and plasmonic-based nanomaterials are discussed for the photocatalytic reduction of CO2 to methanol. Finally, a summary and perspectives on this emerging field are provided.

Graphical Abstract

1. Introduction

Today’s global issues and challenges include the energy crisis and environmental concerns [1]. Carbon dioxide comes primarily from the combustion of carbon sources such as fossil fuels and other natural sources. Developing renewable energy technologies to reduce pollutant emissions has become a significant area of research for the development of a sustainable planet.
In light of this, various advanced nanomaterials for CO2 reduction have been reported, including alkali hydroxide [2,3], CuNi@g-C3N4/TiO2 [4,5], Pd50-Ru50/MXene [6,7], CoSA-Ti3C2Tx [8,9], CuSAs/TCNFs carbon nanofibers [10], and copper selenide (Cu2xSe(y) nanocatalysts) [11]. Gawande and his co-workers recently reviewed advanced Ag-based nanomaterials used for various photocatalytic applications, including CO2 hydrogenation processes [12]. One promising approach to addressing issues related to climate change and the energy crisis is photocatalytic CO2 reduction into value-added chemicals. Photocatalysis is considered to be a promising method for CO2 conversion into valuable products, such as methanol, methane, formaldehyde, ethanol, and higher hydrocarbons [13]. Photosynthesis is an ideal method for effectively resolving issues relating to energy and the environment by reducing CO2 into value-added chemicals and fuels [14]. The photocatalytic performance of a photocatalyst is highly dependent upon its electronic band structure and bandwidth energy. For an effective photocatalyst, the bandgap energy must be less than 3 eV to expand the light absorption in the visible area and use solar power efficiently. To date, a variety of photocatalysts including P- and F-co-doped carbon nitride (PFCN) [15], RuSA–mC3N4 [16], Cu-ZIF [17,18], single Cu2O particle [19,20], g-C3N4-TiO2 [21], (Pd/Pt)SA/g-C3N4 [22], O-doped g-C3N4 (OCN-Tube) [23], Cu-TiO2 [24], Ni-nanocluster loaded on TiO2 (Ni/TiO2[Vo]) [25], aerogel flow-reactor [26], porous-g-C3N4/TiO2-nanotube [27,28,29], carbon-doped TiO2 [30,31,32], RGO-NH2-MIL-125(Ti) [33], Cu porphyrin-based MOF [34], Zn2GeO4/ZIF-8 nanocomposite [35,36], TZTZ-TA-CMP [37], graphene quantum dots [38] and In2O3–CuO [39] have been reported for photocatalytic CO2 reduction to valuable chemicals. Figure 1 indicates that the interest of researchers is continuously growing in the field of “Photocatalytic reduction of CO2 to methanol”. In the last decade, there has been a significant increase in the publication of research articles. Recently, carbon, TiO2, MOFs, and mixed-metal-oxide-based supporting materials have been utilized for photocatalytic hydrogenation of carbon dioxide into methanol. In this context, highly efficient solar-convertible materials to be reserved into chemicals are desired.
Due to its role as a greenhouse gas, the conversion of CO2 to alternative chemicals has enabled many options. However, CO2 is highly thermodynamically stable, which means that converting it into valuable chemicals requires a large amount of energy. A promising photo-reduced product of CO2 is methanol. Methanol has multiple applications, including fuel transportation, biodiesel transesterification, and electricity generation [40,41,42]. The CO2 conversion to methanol is a highly suitable method for reducing CO2 emissions into the atmosphere. Methanol is greener than gasoline and has a high energy density. In addition to being a feasible clean fuel, methanol is also an important feedstock for chemical industries. The internal combustion engines can directly use it if stored at atmospheric pressure (atm P) due to its high octane number [43,44]. In light of all the above advantages of methanol, we examine the various strategies to enhance the photocatalytic CO2 conversion to CH3OH using carbon-based, TiO2 based, MOFs based, mixed metal-oxide, and plasmonic-based photocatalysts.

2. Scope and Focus of This Review

CO2 emissions are one of the greatest environmental concerns globally today. It is extremely necessary to convert this atmospheric CO2 into valuable fuel or chemicals through catalysis. Photocatalysis is one of the ideal paths for CO2 conversion into value-added products.
There has not been a review on photocatalytic selective hydrogenation of CO2 to methanol yet. Though there are some reviews on CO2 reduction into value-added chemicals reported, none of these reviews focused solely on methanol [43,44,45,46,47,48,49,50,51,52,53,54]. Thus, in this review, we highlighted the photocatalytic CO2 reduction to selectively solar fuel such as methanol by using photoactive supporting materials, including carbon, TiO2, MOF, mixed metal-oxide, and plasmonic based photocatalysts (Figure 2). We believe that a detailed overview of the catalytic performance of various photocatalysts for CO2 reduction to methanol would be useful to a broad community of scientists with interests in nanotechnology, materials chemistry, inorganic chemistry, organic chemistry, and chemical engineering.

3. Applications

Photocatalytic CO2 reduction is a promising method for converting CO2 into valuable fuels and chemicals by utilising solar energy. An everyday variety of products utilize utilise methanol as a chemical building block, such as paints, plastics, and construction materials. In this section, a variety of photocatalysts for CO2 reduction to methanol are included, including MOF-based, mixed-metal oxide-based, carbon-based, TiO2-based and plasmonic-based photocatalysts.

3.1. Photocatalyst Based on MOFs

The metal–organic framework (MOF) is a hybrid (inorganic–organic) crystalline porous material that consists of metal ions surrounded by organic linkers. Due to an internal hollow structure, it has a remarkably large internal surface area, since the metal ions serve as nodes that bind the linker arms together [55]. In contrast to other porous materials, MOFs exhibit unparalleled structural diversity–atomic structural uniformity, uniform pore structures, tunable porosity, as well as flexibility in network topology, and chemical utility. MOFs’ cage-like structure is currently being exploited in numerous fields, including purification, gas separation and storage, liquid separation, sensing, gas storage, and catalysis [49]. The various types of MOF-based photocatalysts, such as MOF-based, MOF composites, MOF-derived, MOFs as support, and single-site MOFs, are employed for the reduction of CO2.
In recent years, MOF-based materials for photoreduction of CO2 have attracted significant research interest [50]. MOF materials are easily designed with convenient metallic sites, specific heteroatoms, and an orderly structure of functional organic ligands [54]. They can efficiently increase the efficiency of photocatalytic activity and electron–hole separation. MOF porosity can help to expose channels for reactant adsorption and more active sites, thus resulting in an excellent catalytic performance. This can increase the efficiency of charge transfer and solar power while inhibiting the recombination of photo-induced electrons. Based on the above merits, the researchers attempt to use various MOFs for the photocatalytic reduction of CO2.
MOFs have porous and channel structures that provide anchored centres for photocatalysis, giving them their structural and functional characteristics. Wang et al. reported composites (Cd0.2Zn0.8S/UiO-66-NH2) with various UiO-66-NH2 compositions via the solvothermal method, which is used for CO2 photoreduction and hydrogen evaluation under visible light [55]. The electron microscopic (TEM/HR-TEM) images of CZS/UN20 confirmed that Cd0.2Zn0.8S NPs were evenly distributed on the surface of the UiO-66NH2 cubes, and the lattice spacing (0.313 nm), which corresponds to the (111) plane of Cd0.2Zn0.8S (Figure 3a,b). The composite (Cd0.2Zn0.8S/UiO-66-NH2) showed the largest photocurrent density, implying effective photo-generated charge transfer (between Cd0.2Zn0.8S and UiO-66-NH2), which is reliable for improved photocatalytic performance (Figure 3c). The UiO-66-NH2 content significantly influences the photoactivity of Cd0.2Zn0.8S. The CZS@UN20 sample shows the higher methanol evolution rate among the as-prepared samples (6.8 µmol h−1 g−1) under irradiation of visible light; it was larger (3.4 times) compared to pure Cd0.2Zn0.8S (Figure 3d). Because of the transfer and charge separation (between Cd0.2Zn0.8S and UiO-66-NH2), the photo-induced electrons absorbed by these materials could be transferred over the surface for the reduction of CO2. Apart from photocatalytic performance, the lifetime of the catalyst is very important. Furthermore, the photocatalyst (Cd0.2Zn0.8S/UiO-66-NH2) showed excellent stability after four recycles in the system of photoreduction CO2 (Figure 3e). A plausible interface electron transfer behaviour and a related mechanism of photocatalytic CO2 reduction were demonstrated in Figure 3f.
The Z-scheme heterojunction has a high electron–hole pair separation efficiency and redox ability and a broad light response range. The Z-scheme heterojunction is a wonderful alternative for converting CO2 into value-adding compounds because of the advantages listed above. The artificial Z-scheme photocatalyst is typically made up of two connected semiconductor photocatalysts: one for oxidation and the other for reduction [56]. The Z-scheme gets its name from the fact that it connects the two photosystems in a fashion that looks like the letter “Z”. In particular, the engineered Z-scheme direct photocatalysts that mimic the natural photosynthetic system provide a number of advantages, including improved light uptake, spatially separated reductive and oxidative active sites, and well-preserved strong redox capacity. Furthermore, photogenerated separation was made possible by the heterojunction at the interfaces, which boosted the charge participation in catalysed conversion reactions [57]. Heterojunctions (Z-scheme) are an effective way to isolate photogenerated electron holes and improve photocatalytic activity in semiconductors. On the other hand, heterojunction-based MOFs are rarely documented. In a similar context, Liu et al. have recently designed a Z-scheme O-ZnO/rGO/UiO-66-NH2 (OZ/R/U) heterojunction that was obtained by pairing UiO-66-NH2 and rGO (reduced graphene oxide) with O-ZnO (oxygen-poor) by a facile solvothermal method [58]. The SEM and TEM images of the OZ/R/U catalyst displayed that the UiO-66-NH2 and O-ZnO were well dispersed over the rGO NS with close contact (Figure 4a,b). The OZ/R/U catalyst has a much higher photocurrent intensity than the OZnO and UiO-66-NH2 catalysts. As a result, it confirms the high charge transfer and separation ability of the OZ/R/U (Figure 4c). The optimal rGO content was 1.5 wt% for the photoactivity of the composite OZnO/rGO/UiO-66-NH2 photocatalyst (Figure 4d). Under visible light irradiation, the ternary photocatalyst demonstrated excellent photocatalytic performance for CO2 reduction into methanol and formic acid, as well as good stability. The results showed that this ternary composite could effectively reduce CO2 to HCOOH and CH3OH and that activity was far better than ZnO/rGO/UiO-66-NH2 and O-ZnO/UiO-66-NH2. Under visible light irradiation, the yield of HCOOH and CH3OH over the ternary composite (O-ZnO/rGO/UiO-66-NH2) achieved 6.41 and 34.83 μmol g−1 h−1, respectively (Figure 4e). The photoinduced electrons on the CB of O-ZnO are transferred to the VB of UiO-66-NH2 through rGO and combined with the photoinduced holes, resulting in the accumulation of electrons on the CB of UiO-66-NH2 and the accumulation of holes over the VB of O-ZnO. The CB level of UiO-66-NH2 is more negative than the reduction potential of CO2. The electrons accumulated over CB of UiO-66-NH2 can readily reduce CO2 to methanol. This Z-scheme charge transfer process of the above two semiconductors can reduce the rate of recombination of photogenerated charge carriers while maintaining the photogenerated electrons’ high reduction capacity at more negative CB of UiO-66-NH2 and the strong oxidation capacity of the photoinduced holes over more positive VB of O-ZnO. This Z-scheme photocatalytic system improved charge separation efficiency by transferring electrons from the CB of O-ZnO to the rGO NS and then recombining those electrons to holes in the VB of UiO-66-NH2 (Figure 4f). This boosted composite photocatalyst activity was assigned to the development of a photocatalytic system (Z-scheme), which inhibits photogenerated charge carriers’ recombination, while maintaining the high reduction capacity of UiO-66-NH2 and the strong oxidizing ability of O-ZnO (Figure 4f).
The copper single-atoms (CuSA) on the UiO-66-NH2 support (CuSA@UiO-66-NH2) have been developed for photoreduction of CO2 into liquid fuels in water medium (Figure 5a) [59] under visible-light irradiation. The morphology of the CuSA@UiO-66-NH2 photocatalyst was studied by HR-TEM and AC-STEM analysis, which clearly verified the occurrence of a Cu single-atom catalyst marked with red circles (Figure 5b–d). The images of elemental mapping of (CuSA@UiO-66-NH2), furthermore, demonstrated that the atoms N, C, O, Cu and Zr are uniformly dispersed on the support (Figure 5e). Remarkably, it was observed that the photocurrent density of CuSA@UiO-66-NH2 was more than UiO-66-NH2, and the CuNPs@UiO-66-NH2 catalysts showed that a larger number of carriers were produced under visible light illumination (Figure 5f). Remarkably, the developed CuSA@UiO-66-NH2 reached the solar-induced CO2 conversion to ethanol and methanol with an evolution rate of about 4.22 μmol h−1 g−1 and 5.33 μmol h−1 g−1, respectively. The yields were significantly higher as compared to pure counterparts (Figure 5g).
Liu et al. developed (ZIF-8/g-C3N4) composites by developing various compositions of nanoclusters of ZIF-8 over the surface of g-C3N4 [60]. The ZIF-8 nanoclusters were deposited over g-C3N4 nanotubes (NT); the overall tubular morphology remained the same (Figure 6a). Additionally, some nanoparticles protruded from the tube wall of g-C3N4 nanotubes. The corresponding TEM image indicated that the ZIF-8 was effectively inserted on the surface of the g-C3N4 NT (Figure 6b). After ZIF-8 surface grafting, charge separation and transfer efficiency were retarded. Further confirmation of this conclusion came from the photoelectrochemical measurements. Among the three samples, a TCN sample achieved the highest stable photocurrent value, indicating the rapid charge separation (Figure 6c,d). Due to the incorporation of ZIF-8 nanoclusters, ZIF-8@g-C3N4 composites could adsorb more CO2 than the g-C3N4 NT without compromising light absorption ability. The (ZIF-8/g-C3N4) composites had improved photocatalytic functioning for the reduction of CO2 due to the higher charge separation efficacy and CO2 capture ability of ZIF-8 from the g-C3N4 NT, where the highest methanol evolution rate arrived at 0.75 mmol h−1 g−1 under the ZIF-8/g-C3N4 composite (Figure 6e). The methanol production rate for bulk g-C3N4 and g-C3N4 NT was 0.24 and 0.49 mmol h−1g−1 respectively, under similar conditions. However, the methanol was not produced on pristine ZIF-8 nanocrystals. Remarkably, the number of grafted ZIF-8 nanoclusters was critical for the conversion of ZIF-8 and synergetic nanostructural patterns. By combining nanostructure semiconductor and MOF grafts, the ZIF-8/g-C3N4 photocatalyst exhibited an enhanced capacity of CO2 adsorption, light-harvesting ability, and charge separation efficiency; it therefore, exhibited an excellent increment in effective photocatalytic methanol generation. Typically, a higher CO2 adsorption capability accompanies a greater efficiency of photocatalytic CO2 reduction for a semiconductor photocatalyst.
In a membrane reactor, Maina et al. demonstrated the controlled incorporation of TiO2 nanoparticles and CuII-doped TiO2 nanoparticles over ZIF-8 films via the rapid thermal deposition (RTD) method (Figure 7a) [61]. The presence of TiO2 NPs in the MOF matrix was emphasized with the red square (Figure 7c). Under UV irradiation, the Cu-TiO2@ZIF-8 hybrid film exhibited powerful photocatalytic activity. The results indicate that, compared with only the quantity generated by pure ZIF-8 film single, the yields of CH3OH and CO raised by 50% and 188%, respectively (Figure 7d). Furthermore, when Cu@TiO2 nanoparticles are loaded on MOF, the photocatalytic CO2 reduction to CO and CH3OH is considerably enhanced (Figure 7e). When the loading of Cu@TiO2 nanoparticles was 7 μg, the Cu-TiO2@ZIF-8 displayed excellent catalytic efficacy. Compared to the pure ZIF-8 film, CO and CH3OH yields were 23.3% and 70% respectively. This is attributed to a synergistic effect caused by semiconductor NPs’ ability to generate photogenerated electrons under light illumination and the CO2 adsorption potential of MOFs.
In 2018, Cardoso et al. prepared a MOF-based Ti/TiO2NT-ZIF-8 photocatalyst by growing films of ZIF-8 over Ti@TiO2 nanotube electrodes through a layer-by-layer method (Figure 8a) [62]. Figure 8b shows the addition of ZIF-8 composites to the TiO2 NTs. Figure 8c displays the image captured by TEM for Ti/TiO2NT-ZIF-8: ZIF-8 nanoparticles can be observed between and within TiO2 nanotubes. The photo-electrocatalytic CO2 reduction by (Ti/TiO2NT-ZIF-8) electrodes was executed in sodium sulphate saturated with CO2 at a constant potential (+0.1 V) under UV-visible light illumination. Notably, 0.7 mmol/L of methanol and 10 mmol/L of ethanol were generated in 3 h (Figure 8d). Because of the lack of ZIF-8 films, the Ti4+ species in Ti@TiO2NT may have a lower ability for CO2 absorption (Figure 8e).
As a consequence of the interaction between Cu/ZnOx nanoparticles and their ligands and Zr6 SBUs, these ultra-small well-mixed NPs do not agglomerate as much and do not undergo phase separation. The Cu/ZnOx@MOF catalysts produce extremely higher activity with yield (2.59 g MeOH kg Cu−1 h−1), and higher selectivity for the CO2 hydrogenation to obtain methanol, and strong stability for over 100 h. The catalyst was prepared by the in situ reduction of post-synthetic metalized UiO-bpy (Figure 9a). The ideal precatalyst structure of Zn@UiO-bpy-Cu via post-synthetic Cu co-ordination to bpy is shown in Figure 9b. As shown in Figure 9c, the CuZn@UiO-bpy has enriched sites that enhance the yield of methanol. The Cu and ZnOx combine well to form Zn and Zr, which leads to improved hydrogenation of CO2. The STEM-HAADF images of ultrasmall Cu/ZnOx and CuZn@UiO-bpy catalyst clearly showed the Cu and Zn were uniformly dispersed and mixed across the entire MOF particle and nanoparticles of 0.5−2.0 nm in size (Figure 9d,e). The UiO-bpy proved to have an effective selectivity for methanol production (Figure 9f).
It was crucial to enhance the stability of metal-oxide QDs for reaction patterns containing H2O so that they could be employed in photocatalytic CO2 reduction. Recently, Li et al. designed the (g-C3N4/CuO@MIL-125(Ti)) photocatalyst for the reduction of CO2 with the proximity of H2O by encapsulating CuO quantum dots with pores of MOFs of MIL-125 (Ti) via a complexation-oxidation method (Figure 10a) [64]. In aqueous reaction systems, the photocatalyst displays significantly enhanced stability due to the protection offered by the MIL-125(Ti) framework. Because of the close contact between the CuO QDs and the Ti active site in MIL-125 (Ti), electron transfers between the confined CuO QDs and the Ti active site were observed (Figure 10b). The presence of water greatly increases the photocatalytic activity of g-C3N4/CuO@MIL-125(Ti) composites for the photoreduction of CO2. Among the prepared catalysts, the resultant 2.5% g-C3N4/1%CuO@MIL-125(Ti) photocatalyst performed better for photocatalytic reduction of CO2 to CH3OH, CO, CH3CHO, and CH3CH2OH, implying water (reductant), with yields of 997.2, 180.1, 531.5, and 1505.7 umol/g, respectively (Figure 10c). Figure 10d–g depicts the structural morphologies of the 2.5% g-C3N4/1% CuO@MIL-125(Ti) photocatalyst. This study developed an efficient method for enhancing charge separation efficiency and stability of metallic-oxide QD-adapted photocatalyst. Furthermore, MIL-125(Ti), due to its high specific surface area and porous structure, provides an encapsulating structure for CuO QDs to prevent aggregation and provides excellent stability and recyclability. Moreover, the unique heterostructure of the composite ensures high light absorption and efficient electron transport from g-C3N4 nanosheets and MIL-125(Ti) to QDs as well. The recent research progress of MOF-based materials/photocatalysts for photocatalytic reduction of CO2 is summarised in Table 1.

3.2. Mixed-Metal-Oxide-Based Photocatalyst

Mixed metal oxides (MMOs), containing two or more types of metals and oxygen, have been broadly employed as photocatalysts for the reduction of CO2. The semiconducting properties of their aqueous suspensions irradiated with visible light have been a prominent subject of research. Mixed metal oxide (MMO) is one of the most important photocatalysts, with features that differ from ordinary oxides in circumstances such as acid-base, redox, and surface area. These have been widely explored for several catalytic applications due to their excellent chemical-thermal stability compared to single oxides [70,71,72].

Transition and Non-Transition Metals

Singhal et al. reported Ni/InTaO4 as supported catalysts for the CO2 photoreduction to methanol under visible light irradiation by the facile sol–gel method. Loading of 1 wt% Ni on InTaO4-produced methanol with a higher yield and slowed down recombination by reducing the bandgap slightly. The bandgap of InTaO4 and Ni/InTaO4 catalysts was noted at 2.6 eV and 2.54 eV, respectively. The SEM image shows the agglomerated nature of the catalyst, and EDX confirmed the existence of elements (In, Ta, O, and Ni). TEM displays the size of nanomaterials at 50–80 nm. The HR-TEM demonstrated the crystallinity of InTaO4, which was a good match for the XRD pattern. XPS measurement of the catalyst showed the presence of Ni and NiO due to four peaks at different binding energies. The catalyst displayed sharp diffraction peaks, indicating monoclinic InTaO4. The material crystalline phase is not altered by Ni loading [73]. Recently, the bimetallic oxide was derived from MOFs developed by Cheng et al. [74].
Kumar et al. proposed a highly efficient rGO-covered magnetically separable core-shell-structured microsphere photocatalyst (rGO@CuZnO@Fe3O4) for CO2 reduction below visible light illumination (Figure 11a) [75]. In order to allow simultaneous CO2 reduction and water oxidation, the semiconductor should have a wide bandgap. Neither Fe3O4 nor ZnO by themselves can convert CO2 to methanol because electrons can only move from the valence band (VB) to conduction band (CB) under the influence of UV light. The morphology of the surface and core-shell structure of microspheres were examined by FE–SEM tomography. The structure of rGO@CuZnO@Fe3O4 remains the same after GO reduction to rGO (Figure 11b). The HR-TEM image clearly showed the 1D lattice plane at about 0.26 nm interplanar distance, corresponding to the (002) plane of ZnO. Additionally, the amorphous zone (0.64 nm thickness) related to rGO was visual (Figure 11c,d). The rGO@CuZnO@Fe3O4 catalyst has shown excellent catalytic activity for a methanol yield (2656 µmol g−1) among the various catalysts prepared (Figure 11e). The advantages of this catalyst include a high yield of methanol, no sacrificial donor requirement, ease of recovery, and effective recycling (Figure 11f). The methanol yield increased with Cu loading from 0.25 to 1.0 wt%; however, the photocatalytic performance did not improve with a further increase in Cu content. The enhancement in the reduction in methanol yield due to Cu content can be attributed to the trapping of photo-generated electrons, which reduces the recombination process of electron holes (Figure 11g). This catalyst has a low price, higher electron mobility, and environmental preservation than ZnO. The large surface area and excellent charge-carrier mobility of rGO can lead to the development of a highly efficient, reusable photocatalyst for CO2 reduction.
Gao et al. synthesized separate Vv-rich and Vv-poor o-BiVO4 atomic layers with the thickness of a unit of a cell at the gram-scale using an intermediate lamellar hybrid approach (Figure 12a) [76]. The single-unit-cell o-BiVO4 layers with a high content of vanadium vacancies (Vv) are high in the production of methanol(398.3 μmol g−1 h−1) compared to atomic layers of Vv-poor o-BiVO4 (Figure 12c). At the atomic level, a correlation between defect sites and photoreduction of CO2 was investigated. A density-functional theory calculation shows that vanadium (V) vacancies introduce a new defect level and greater hole concentration around Fermi levels, resulting in enhanced electronic conductivity and photoabsorption (Figure 12b). The increased surface photovoltage of layers with vanadium voids is confirmed by the enhanced carrier lifetime of the o-BiVO4 layers, which is shown through time-persistent fluorescence spectra.
Yu et al. prepared a novel efficient visible light 2D photocatalyst (Ti3C2/Bi2WO6 nanosheets) via in situ progress of ultrathin nanosheets of Bi2WO6 with a surface of Ti3C2 nanosheets (Figure 13a) [77]. The generation of the 2D/2D heterojunction was characterized by TEM studies; the results demonstrated the clean ultrathin nanosheets along with lucent features, nanoparticles of Ti3C2@ Bi2WO6 nanosheets, and the element composition and distribution (Figure 13b–d). The Ti3C2@Bi2WO6 heterostructured hybrids with distinct atomic layers had much higher photoreduction activity than pure Bi2WO6, with a 4.6 times higher overall yield of CH3OH and CH4 than pure Bi2WO6 (Figure 13g). Under solar light irradiation, photo-induced electrons are excited and jump from the VB to the CB of Bi2WO6. The Bi2WO6 has a more negative CB potential compared to the EF of Ti3C2 with terminal -O; photoinduced electrons can then be transferred from Bi2WO6 to Ti3C2 via an ultra-thin layered heterojunction. The photo-induced electrons that have accumulated on the surface of Ti3C2 can then interact with the CO2 molecules that have been adsorbed. The photo-induced electrons can be quickly transferred from the bulk of Bi2WO6 to the heterojunction interface up to the Ti3C2 surface because of the unique atomic layer 2D/2D heterostructure. Thus, a significant enhancement in the efficiency of photocatalytic CO2 reduction was observed. This 2D photocatalyst possesses a high contact area and a small distance of charge transport. As a result, electrons are efficiently transferred from the Bi2WO6 (photocatalyst) to the Ti3C2 (cocatalyst) (Figure 13h). The enhancement of photocatalytic functioning is due to excellent CO2 adsorption ability and effective separation of charge carrier. The present work demonstrated that Ti3C2 nanosheets can be employed as effective cocatalysts for the photoreduction of CO2.
Viswanathan et al. synthesized a photocatalyst (Fe-N/Na(1−x)LaxTaO(3+x)) via a hydrothermal method. Using a variety of approaches, researchers evaluated the effect of doping or co-doping LNTO (lantana) with iron (Fe) and nitrogen (N) on the photo-physical properties of catalysts. The dopant iron (Fe-3d) energy level of the orbital was lower than the Ta-5d energy level. The new energy level is under the conduction band. The Fe3+ ion (dopant) retains the capacity to capture and de-capture the charge carriers. It can readily capture photogenerated electrons from the CB of NaTaO3. A decrease in the PL line intensity suggests that dopants/co-dopants (Fe or N) can reduce recombination, therefore enhancing the charge carrier’s lifetime. The enhancement in activity is because of co-doping with N and Fe over lantana (LNTO). The XPS and XRD results show that the Fe3+ occupies the Ta5+ ions site, La3+ ions Na+ ion sites, and Nitrogen (N) takes O2− sites. N and Fe co-doping in a Ta matrix results in the confining of the bandgap because of the formation of secondary energy levels between the bandgap and the absorption of visible light. The presence of dopants (La3+ and Fe3+) in the tantalate (Ta) lattice assures electro-neutrality. The charge trapping and de-trapping of (Fe3/Fe4+) ions efficiently decreases the recombination of the charge carrier. The co-doping of LNTO through N and Fe increases the absorption of visible light because of the synergetic effect. It hinders charge carrier recombination and boosts charge transfer interfacial, enhancing photocatalytic CO2 reduction to CH3OH in aqueous alkaline media higher than that of pure NaTaO3 [78].
Kumar et al. developed a Z-scheme heterostructured (rGO/InVO4/Fe2O3) photocatalyst for the reduction of CO2 to methanol over visible light illumination by the deposition–precipitation method. The existence of a few dark spots over the sheet surface suggested the presence of iron oxide (Fe2O3) in the nanocomposites. In Raman spectra, the presence of comparable bands of all constituents at 468, 924, 1329, and 1587 cm−1 proved the existence of Fe2O3, InVO4, and rGO in the composites. In comparison to Fe2O3 and InVO4, notably rGO/InVO4/Fe2O3 had a lower PL intensity, indicating less charge recombination and thus greater photocatalytic activity. The XRD pattern indicated the appearance of Fe2O3 and InVO4 into the composite The photo-induced electron transfer occurred from the CB of Fe2O3 into the VB of InVO4 and further to the CB of InVO4 over the Z-scheme system for the CO2 photoreduction. Triethylamine (TEA) was employed as a sacrificial electron donor to achieve methanol yields as high as 16.9 mmol gcat−1 employing the Z-scheme-based photocatalyst. The recycling ability of the photocatalyst demonstrated greater stability and efficient recyclability. The synergistic effect of the ternary rGO/InVO4/Fe2O3 photocatalyst leads to effective separation of charge carriers and charge mobility over the surface of the catalysts, which results in an effective reduction of CO2 and an enhancement in the yield of methanol [79].
Yan et al. [80] developed a polymorph of indium oxide as a photocatalyst for the formation of CO and CH3OH. Zhang et al. reported an efficient, highly selective, and stable photocatalyst (In2O3−x(OH)y) with a rod-like nano-crystallized structure for hydrogenation of gaseous carbon dioxide to methanol and an evolution rate (0.06 mmol gcat −1 h−1) under solar irradiation [81]. The SEM image showed the rod length of the nanocrystal superstructure to be about 2.6 μm (Figure 14a). The nanoporous nature of the materials produced, made up of nanocrystalline superstructures, has been confirmed by TEM studies (Figure 14b,c). The photocatalytic stability of the In2O3−x(OH)y nano-crystallized structure sample was illustrated. The stabilized rate of methanol within the first 15 h led to a significant enhancement in the 16th h optimized at 250 °C in the light for 20 h (Figure 14d). DFT measurements of the energy profile for the production of methanol via CO2 hydrogenation over In2O3−x(OH)y (111) were investigated. An acetal intermediate was formed by adding a hydride (H2CO2*). The lowest energy issue was chosen as the probable intermediate (Figure 14e).

3.3. Photocatalysts Based on Carbon Materials

Carbon-based photocatalysts have become increasingly popular for the reduction of CO2 because of their excellent physicochemical and electrochemical properties. Several excellent carbon-based supports such as graphene, CNT, carbon dots, graphitic carbon-nitride, and conducting polymers have been employed for several applications over the years. The use of different carbonaceous materials as supports for photocatalysis is highly beneficial, since carbon has fair photocatalytic activity, high surface area, high electrical conductivity, higher dispersion, and visible light absorption capability [82,83].
Gusain et al. synthesized rGO-CuO nanocomposites for CO2 reduction to methanol over visible light illumination. The deposition of CuO nanorods over rGO enhanced photocatalytic efficiency and produced a higher yield of methanol compared to pure CuO nanorods [84]. In a similar line, Liu et al. prepared cuprous oxide decorated to various morphologies of rGO(Cu2O/rGO photocatalysts) for CO2 reduction under visible light. As-prepared and well characterized rhombic dodecahedral Cu2O/rGO catalysts exhibited the maximum methanol yield [85].
Yu et al. reported a ternary composite (Ag2CrO4/g-C3N4/GO) for the photoreduction of CO2 to methane and methanol by using silver chromate NPs (photosensitizer) and graphene oxide (cocatalyst) [86]. The dispersion of Ag2CrO4 nanoparticles over the surface of the g-C3N4 sheet was investigated through SEM and TEM imaging (Figure 15a,b). The triplet composites displayed an increased conversion of CO2 with a TOF of 0.30 h−1, which was higher than bare g-C3N4 under sunlight irradiation. The photocatalytic activity was enhanced because of expanded light absorption, greater adsorption of CO2, and efficient charge separation. Ag2CrO4 nanoparticles can improve the g-C3N4 light absorption, and GO cocatalyst (electron acceptor), besides promoting charge transfer, also provides higher adsorption of CO2 and catalytic sites. Notably, charge transfer takes place between Ag2CrO4 and g-C3N4 via a direct Z-scheme mechanism, which enhances the photocatalytic system’s redox ability while promoting charge separation (Figure 15e).
Jain et al. developed a cobalt phthalocyanine embedded over g-C3N4 as a hybrid photocatalyst (denoted as g-C3N4/CoPc-COOH) for selective reduction of CO2 into CH3OH by employing a sacrificial electron donor (triethylamine (TEA) in the presence of visible light (Figure 16a) [87]. The immobilization of the CoPc-COOH complex over the g-C3N4 support was confirmed by HR-TEM (Figure 16b). The resultant g-C3N4/CoPc-COOH hybrid provided greatly increased affinity of CO2 through active Co2+ sites, charge separation through the g-C3N4 surface, as well as large surface areas for efficient CO2 conversion. They have demonstrated and investigated that the hybrid photocatalyst offered a much improved yield of methanol than the reported catalysts. The higher methanol yield over the hybrid photocatalyst was observed after 24 h, employing TEA as an electron contributor (Figure 16c). Notably, the durable interaction between –COOH with heteroatoms of g-C3N4 support makes a robust hybrid photocatalyst that ultimately inhibits metal leaching during photoreduction.
The Gong group investigated the production of a family of polymeric C3N4/CdSe QDs (p-CNCS) with various particle sizes of CdSe for CO2 photoreduction into methanol, targeting maximum selectivity and activity by the quantum confinement effect [88]. Based on the quantum confinement effect, the energy of the CB electrons was modified to a suitable value, which was deficient for H2 formation but sufficient for methanol formation, and which enhances activity and selectivity for methanol production from the photocatalytic CO2 reduction reaction (CO2RR). CdSe QDs were packed on p-C3N4 via an impregnation method, producing p-CNCS as a photocatalyst (Figure 17b–d). The correlation between the band energy and photocatalytic activity have been discussed. Moreover, the high surface area and appropriate dangling bonds of p-C3N4 (Polymeric carbon nitride nanosheet) provide sufficient loading sites for CdSe quantum dots, allowing a strong interaction between p-C3N4 and CdSe to the heterojunction (Figure 17e,f). The heterojunction within 0D/2D materials could transport photoinduced holes and electrons to CdSe and p-C3N4, enhancing charge separation and increasing CdSe stability.
Bafaqeer et al. synthesized 2D/2D heterojunction composite catalyst (2D/2D ZnV2O6/pCN) via combing ZnV2O6 and g-C3N4 (protonated -CN) for selective photoreduction of CO2 with H2O via the solvothermal method [89]. The HR-TEM was employed to examine the crystalline structure of the pCN merged with ZnV2O6 nanosheets (Figure 18b,c). The surface charge conversion with protonation of g-C3N4 acts as a moderator and trapped photo-excited electrons. The performances of as-prepared catalysts were studied in gas-phase and liquid photocatalytic methods under light irradiation (UV/visible). The methanol formation rate of the 2D/2D ZnV2O6/pCN(100%) composite catalyst was 3742 μmol g-cat−1 higher than that of the pure ZnV2O6 and pCN catalysts (Figure 18d,e). When bombarded with light, the pCN in the heterostructured photocatalysts functions as a sensitizer, absorbing photons and exciting electron and hole pairs. Photoexcited electrons on pCN might be transported to ZnV2O6 CB because the CB edge potential of pCN (1.12 eV) is more negative than that of ZnV2O6 (0.87 eV). The protonation of g-C3N4 can operate as an excellent acceptor and trap for photoexcited electrons, allowing the photo-induced electrons to be quickly transported to the ZnV2O6 and converted to methanol. The photoexcited electron–hole pairs could be effectively separated in this way. As a result, the created junction between pCN and ZnV2O6 in heterostructured photocatalysts reduces electron and hole recombination during the charge carrier transfer process, resulting in improved photocatalytic activity of pCN/ZnV2O6 heterojunctions. In addition to the synergistic effect of ZnV2O6 and pCN nanosheet heterojunctions, combining pCN with ZnV2O6 resulted in increased activity. Furthermore, the hierarchical structure, rich 2D coupling surfaces, increased interfacial contact, and charge separation could all help to improve photoactivity and product selectivity. The photocatalytic stability of the 2D/2D ZnV2O6/pCN composite catalyst with a moderator was excellent.
Kumar et al. synthesized CNT-TiO2 photocatalysts for photoreduction of CO2 and water splitting by joined sonothermal and hydrothermal method [90]. Figure 19a–d shows the effect of CNT on TiO2 morphology. The (101) plane of anatase TiO2 (lattice spacing 0.342 nm) is confirmed by HRTEM. The CNT content increases the attachment of spherical TiO2. The prepared CNT-TiO2 photocatalysts showed increased photocatalytic activity for the reduction of CO2 compared to bare TiO2 in the presence of visible light. As prepared catalysts, the 2.0 CNT-TiO2 showed better performance for the yield of methanol under UV light (Figure 19g,h). According to a computational study, the binding of CNT to TiO2 NPs was desirable at (101) surfaces rather than (001) facets (Figure 19e). The photoexcitation of this composite over visible light leads to charge transfer within CNT and TiO2 and the formation of isolated charge carriers, while UV light excitation leads to charge transfer in all directions from TiO2 to CNT and CNT to TiO2 (Figure 19i).
Tang et al. investigated a special variety of CD (carbon-dots) as a hole receiver for mCD/CN composite when produced via the flexible microwave method [38]. The co-residence and morphological information of both mCD and CN structures were investigated through HR-TEM. The as-prepared CN was graphene-like nanosheets, and mCD has a graphitic structure with a featured-spacing of about 0.23 nm (Figure 20a,b). The mCD/CN composite was c.a. 12 times more active than sCD/CN (sonication method) for the conversion of CO2. Remarkably, the mCD/CN nanocomposite generated methanol and oxygen from H2O and CO2 with a selectivity of around 100% methanol (Figure 20d) and an internal quantum efficiency of 2.1% in the visible area, which was validated through isotopic labelling. Furthermore, the unique mCD captured holes from CN (carbon nitride) and hindered methanol adsorption, which resulted in the oxidization of water instead of methanol and the enhancement of the selectivity for reduction of CO2 to alcohols (Figure 20c).

3.4. TiO2 Based Photocatalysts

Titanium dioxide (TiO2) is a wide-bandgap semiconductor. In a variety of energy and environmental applications, TiO2 is employed as a photocatalyst because of its high stability, efficient photoactivity, low price, and safety for people and the environment. TiO2 has properties such as nontoxicity, easy availability, chemical stability, and the capability to oxidize in the presence of radiation. TiO2-based photocatalysts will be able to solve the main global problems related to pollution of the environment and renewable energy requirements [91,92].
Nogueira et al. reported nanocomposite TiO2/Ti4O9/Cu2O for photocatalytic CO2 reduction through loading of Cu2O on TiO2-pillared tetratitanate (K2Ti4O9) [93]. The structural and compositional characteristics of as-prepared catalysts were investigated through XRD, XPS, and TEM images (Figure 21a–d). The TEM and HR-TEM revealed that the size of the Cu2O and TiO2 nanoparticles were approximately 5 and 10 nm, respectively (Figure 21d,e). The enhancement in the photoactivity of the as-prepared photocatalyst was because of the synergistic effect produced by pillaring K2Ti4O9 (layered semiconductor) with TiO2 and the eventual loading of Cu2O nanoparticles as cocatalyst. The increment in the surface area of the photocatalyst TiO2/Ti4O9/Cu2O was from (25 to 145 m2 g−1) analysed through nitrogen adsorption–desorption isotherms (NADI). The methanol production was double that of the pure K2Ti4O9 due to the enhanced activity of the photocatalyst (Figure 21f). Incorporating Cu2O nanoparticles onto mesoporous solid surfaces extended solar radiation absorption, mobility of electron and charge separation at the surface.
Jiang and co-workers synthesized hybrid carbon@TiO2 composite hollow spheres for CO2 photo-reduction by the simple and benign method of employing colloidal carbon spheres (Figure 22a) [94]. The nanostructure, hollow spherical structure, and elemental mapping of the as-prepared composite were confirmed by FESEM, TEM and STEM analysis (Figure 22b–f). The obtained carbon@TiO2 composite nanostructure displayed good photoreduction of CO2 operation and selectivity over artificial solar light compared to TiO2 (P25). The methanol generation rate was (9.1 μmol g−1 h−1) greater than pure TiO2 (Figure 22g). The carbon composition of the carbon@TiO2 composites considerably affected photocatalytic activity. The improvement in the photocatalytic performance of the carbon@TiO2 photocatalyst was because of enhanced specific surface area (110 m2 g−1), the CO2 absorption capacities, and a native photothermal impact near the photocatalyst due to the carbon. According to the electrochemical impedance spectra (EIS), carbon composition could affect the efficiency of charge transfer of carbon@TiO2 composites.
Yadav et al. successfully loaded Au nanoparticles on S8-TiO2 (S8-TiO2 (40)-Au photocatalyst) via the sol immobilization method. The as-prepared catalysts employed or CO2 photoreduction to methanol and HER under visible light irradiation [95]. Sharma et al. developed a nanocomposite photocatalyst NiO-TiO2/ACF via the sol–gel method for photoreduction of CO2 with H2O into methanol by immobilization of Ni loaded TiO2 on initiated carbon fibers (ACFs) under UV and visible light illumination (Figure 23a) [96]. The SEM images of ACF, ACF-TiO2 and NiO-TiO2/ACF photocatalyst indicated that the even distribution of NiO/TiO2 nanoparticles was found over the surface of ACF with high crystallinity. The ACF surface was almost clean after pre-treatment with a hierarchical porous structure capable of acting as active sites for adsorption (Figure 23b–d). The yield of methanol over NiO-TiO2/ACF photocatalyst was 986.3 and 755.1 μmol g−1 under visible light and UV, respectively (Figure 23e). Increasing the photocatalytic activity of Ni-loaded TiO2 was achieved with the use of the ACF support, which inhibited electron–hole recombination. The ACFs (surface area: 163.9 m2/g) and NiO were used to increase CO2 adsorption capacity and also to alter the electronic absorption properties of TiO2. Moreover, NiO co-doped TiO2 beat the recombination rate issue of electron–hole and induced the formation of Ti3+ and oxygen vacancies for CO2 conversion into CH3OH over UV/visible light irradiation (Figure 23f).
Xu et al. synthesized hybrid TiO2/Ni(OH)2 photocatalysts for photoreduction of CO2 by deposition of perpendicularly arranged Ni(OH)2 nanosheets on the TiO2 fibers by electrospinning and the wet-chemical precipitation method (Figure 24a) [97]. The FESEM, TEM, and HR-TEM images of TiO2/Ni(OH)2 photocatalyst revealed that the hierarchical nanostructure. The thickness of nanosheets of Ni(OH)2 was nearly 20 nm and was well deposited on TiO2 surface (Figure 24b–d). The TiO2/Ni(OH)2 composite nanofibre exhibited unusually enhanced activity of CO2 photoreduction than the bare TiO2 fibers. The yield of CH3OH over this hybrid photocatalyst was increased after loading of 15 wt% Ni(OH)2 and irradiation time (Figure 24e,f). The TiO2/Ni(OH)2 photocatalyst displayed an increased activity of CO2 reduction, selectivity, the efficiency of charge separation, and increased the density of the CO2 on photocatalyst surface due to the existence of Ni(OH)2 nanosheets as cocatalyst. The clusters of Ni(OH)2/Ni can act as electron sinks and stimulate the photo-generated electron separation from TiO2 to a cluster of Ni(OH)2/Ni and toward CO2 molecules, which can play a vital role as active sites and during the photocatalytic process, reducing photo-excited charge carrier recombination (Figure 24g). Later on, modified TiO2/rGO/CeO2 composite photocatalyst was investigated for CO2 photoconversion by the ultrasonication method [98].

3.5. Photocatalysts with Plasmonic Properties

Recently, plasmonic photocatalysis has enabled rapid advances in improving the photocatalytic efficiency for CO2 reduction under irradiation with visible light. It employs precious metal NPs dispersed on semiconductor photocatalysts and has outstanding properties such as localized plasmonic surface resonance (LSPR), which contributes to the strong absorption of visible light and the excitation of active charge carriers. The conduction of electrons on the nanoparticle surface offers plasmonic materials extraordinary optical characteristics. The interaction between free electrons in metal nanoparticles and incident light is known as the plasmonic effect. Plasmonic metal nanoparticles with high light absorptivity have been shown to represent a new class of photocatalysts with features that differ dramatically from those of typical semiconductor photocatalysts. Plasmonic nanoparticles have unique optical, electrical, and thermal properties [99,100,101,102,103,104,105,106,107]. In this section, we focused on plasmonic photocatalysts for the photoreduction of CO2 to methanol.
Becerra et al. developed an efficient method for the preparation of plasmonic photocatalyst (Aux@ZIF-67) [69]. The decoration of Au NPs over ZIF-67 was investigated for the reduction of CO2 under irradiation with sunlight. The plasmonic system showed greatly enhanced photocatalytic activity for CO2 reduction to CH3OH with excellent selectivity. The plasmonic Au NPs (size 30–40 nm) significantly enhanced the absorption of visible light, increased separation of charge, and contributed to selectivity. This nanocomposite demonstrated the advantages of high specific surface area (SSA) as well as electrochemical properties. An optimal production rate of methanol (~2.5 mmol g−1 h−1) was achieved for the conversion of CO2. This deposited Au NPs favours both the activity and the selectivity of synthesized plasmonic photocatalyst because of the injection of energetic electrons on the surface of the ZIF-67 derived from the plasmonic response. The localized surface plasmon resonance (LSPR) at the edges of rod- and triangular nanostructures decreases the recombination of electron–hole pairs during the photocatalytic reduction of CO2. The production of CH3OH (~2.5 mmol g−1 h−1)and C2H5OH (0.5 mmol g−1 h−1) were achieved from Au nanorods and Au nanotriangle, respectively. This Au@ ZIF-67 plasmonic system has been significantly improved in terms of photocatalytic activity to reduce CO2 and improve cost-efficiency. Yadav et al. reported plasmonic photocatalyst (Au/TixSi1−xO2) for photoreduction of CO2 employing an LED light source by deposition of plasmonic Au NPS on mesoporous titania with isolated silica (TixSi1−xO2). Following a DFT study, an excellent adsorption of CO2 on the surface was observed, which could have been due to the incorporation of silica sites. The methanol production was 1835 μmol gcat−1 using Au/TixSi1−xO2 material with 28 mol% Si in titania lattice and 1.0 wt% Au nanoparticle deposition [108].
Ye et al. reported plasmonic catalysts (Cu/ZnO) for photoreduction of CO2 into methanol that promoted visible light illumination under atmospheric pressure. The rich Cu-ZnO interfaces are recognized as active sites for the production of CH3OH (Figure 25a). The production rate of methanol rose from 1.38 to 2.13 μmol g−1 min−1, and the noticeable activation energy was reduced from 82.4 to 49.4 kJ mol−1 (Figure 25b). N2 sorption isotherms specified that the physicochemical characteristics of the decreased Cu/ZnO catalyst were mesoporous in nature (Figure 25c). The mechanism fuses the photo-generated hot electrons onto Cu NPs, and these electrons can transfer to the ZnO via interfaces of metal-support (Figure 25d). The activation of reaction intermediates co-operatively promoted by hot-electrons over Cu and ZnO leads to the photo-promoted synthesis of methanol [109].
Fan et al. established a plasmonic photocatalyst (Cu/TiO2) by employing both hydrothermal and microwave-assisted processes [110]. Both Cu NPs and unique TiO2 film display the properties of light harvesting as per LSPR. The charge carrier recombination was decreased by the deposition of Cu NPs, that were found during fluorescence quenching. The photocatalytic activity of Cu/TiO2 films was excellent because of charge transfer efficiency and LSPR absorption of Cu NPs. Recently, Wang et al. prepared plasmonic photocatalysts (Ag NPs/ACFs) by deposition of plasmonic Ag NPs over acid-ached carbon fibers (ACFs) coupled with ultrasonication for photoreduction of CO2 to methanol under visible light irradiation [111].

4. Conclusions and Future Perspectives

In this review, we discussed the recent research advances of photocatalytic CO2 reduction into methanol. Photocatalysts such as MOFs, mixed-metal oxide, carbon, TiO2 and plasmonic-based photocatalytic reduction of CO2 to methanol were systematically summarized. The photocatalytic activity, photo-excited charge transfers and separation efficiency, CO2 capture capacities, and stability can be enhanced by supporting synthetic procedures, semiconductor, metal, and ligand replacement, and incorporation of photoactive responsive units. A detailed overview of the catalytic performance of various photocatalysts for CO2 reduction to methanol would be helpful to researchers. The selectivity and catalytic activity that are mostly directed by the electronic environment of the metals could be attuned by a combination of support, doping of heteroatoms, and employing a system with multiple metals. The photoreduction of CO2 to valuable chemicals and to fuels has gradually become important because of its efficiency in simultaneously solving global warming and energy crisis problems. Indeed, photoactive materials retain their specific benefits with light-driven CO2 reduction, and all of these efficient variation strategies will offer directions to the rational design of photocatalysts with improved catalytic performance. It is hoped that this review will play a crucial and interesting role in serving future developments in this important field.

Author Contributions

G.Y.S. collected the complete literature and wrote the complete review. A.S.M. collected some part of the literature and wrote the respective part. M.B.G. edited and revised the draft and supervised the complete work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

Gajanan Y. Shinde would like to thank the Council of Scientific and Industrial Research (CSIR), Delhi, for financial support through the NET-JRF fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, C.; Wang, W.; Liu, B.; Qiao, J.; Lv, L.; Gao, X.; Zhang, X.; Xu, D.; Liu, W.; Liu, J.; et al. Recent Advances in MOF-based Nanocatalysts for Photo-Promoted CO2 Reduction Applications. Catalysts 2019, 9, 658. [Google Scholar] [CrossRef] [Green Version]
  2. Sen, R.; Goeppert, A.; Kar, S.; Prakash, G.K.S. Hydroxide Based Integrated CO2 Capture from Air and Conversion to Methanol. J. Am. Chem. Soc. 2020, 142, 4544–4549. [Google Scholar] [CrossRef]
  3. Singh, A.K.; Montoya, J.H.; Gregoire, J.M.; Persson, K.A. Robust and synthesizable photocatalysts for CO2 reduction: A data-driven materials discovery. Nat. Commun. 2019, 10, 443. [Google Scholar] [CrossRef] [Green Version]
  4. Rathi, A.K.; Kmentová, H.; Naldoni, A.; Goswami, A.; Gawande, M.B.; Varma, R.S.; Kment, Š.; Zbořil, R. Significant Enhancement of Photoactivity in Hybrid TiO2/g-C3N4 Nanorod Catalysts Modified with Cu–Ni-Based Nanostructures. ACS Appl. Nano Mater. 2018, 1, 2526–2535. [Google Scholar] [CrossRef]
  5. Bharath, G.; Rambabu, K.; Hai, A.; Othman, I.; Ponpandian, N.; Banat, F.; Loke Show, P. Hybrid Pd50-Ru50/MXene (Ti3C2Tx) nanocatalyst for effective hydrogenation of CO2 to methanol toward climate change control. Chem. Eng. J. 2021, 414, 128869. [Google Scholar] [CrossRef]
  6. Chen, Y.-H.; Qi, M.-Y.; Li, Y.-H.; Tang, Z.-R.; Wang, T.; Gong, J.; Xu, Y.-J. Activating two-dimensional Ti3C2Tx-MXene with single-atom cobalt for efficient CO2 photoreduction. Cell Rep. Phys. Sci. 2021, 2, 100371. [Google Scholar] [CrossRef]
  7. Yang, H.; Wu, Y.; Li, G.; Lin, Q.; Hu, Q.; Zhang, Q.; Liu, J.; He, C. Scalable Production of Efficient Single-Atom Copper Decorated Carbon Membranes for CO2 Electroreduction to Methanol. J. Am. Chem. Soc. 2019, 141, 12717–12723. [Google Scholar] [CrossRef]
  8. Liang, L.; Lei, F.; Gao, S.; Sun, Y.; Jiao, X.; Wu, J.; Qamar, S.; Xie, Y. Single Unit Cell Bismuth Tungstate Layers Realizing Robust Solar CO2 Reduction to Methanol. Angew. Chem. Int. Ed. 2015, 54, 13971–13974. [Google Scholar] [CrossRef]
  9. Rezvani, A.; Abdel-Mageed, A.M.; Ishida, T.; Murayama, T.; Parlinska-Wojtan, M.; Behm, R.J. CO2 Reduction to Methanol on Au/CeO2 Catalysts: Mechanistic Insights from Activation/Deactivation and SSITKA Measurements. ACS Catal. 2020, 10, 3580–3594. [Google Scholar] [CrossRef]
  10. Zhu, J.; Ciolca, D.; Liu, L.; Parastaev, A.; Kosinov, N.; Hensen, E.J.M. Flame Synthesis of Cu/ZnO–CeO2 Catalysts: Synergistic Metal–Support Interactions Promote CH3OH Selectivity in CO2 Hydrogenation. ACS Catal. 2021, 11, 4880–4892. [Google Scholar] [CrossRef]
  11. Yang, D.; Zhu, Q.; Chen, C.; Liu, H.; Liu, Z.; Zhao, Z.; Zhang, X.; Liu, S.; Han, B. Selective electroreduction of carbon dioxide to methanol on copper selenide nanocatalysts. Nat. Commun. 2019, 10, 677. [Google Scholar] [CrossRef]
  12. Sharma, R.K.; Yadav, S.; Dutta, S.; Kale, H.B.; Warkad, I.R.; Zbořil, R.; Varma, R.S.; Gawande, M.B. Silver nanomaterials: Synthesis and (electro/photo) catalytic applications. Chem. Soc. Rev. 2021, 50, 11293–11380. [Google Scholar] [CrossRef] [PubMed]
  13. Al Jitan, S.; Palmisano, G.; Garlisi, C. Synthesis and Surface Modification of TiO2-Based Photocatalysts for the Conversion of CO2. Catalysts 2020, 10, 227. [Google Scholar] [CrossRef] [Green Version]
  14. Yang, X.; Wang, D. Photocatalysis: From Fundamental Principles to Materials and Applications. ACS Appl. Energy Mater. 2018, 1, 6657–6693. [Google Scholar] [CrossRef]
  15. Kumar, S.; Gawande, M.B.; Kopp, J.; Kment, S.; Varma, R.S.; Zboril, R. P- and F-co-doped Carbon Nitride Nanocatalysts for Photocatalytic CO2 Reduction and Thermocatalytic Furanics Synthesis from Sugars. ChemSusChem 2020, 13, 5231–5238. [Google Scholar] [CrossRef]
  16. Sharma, P.; Kumar, S.; Tomanec, O.; Petr, M.; Zhu Chen, J.; Miller, J.T.; Varma, R.S.; Gawande, M.B.; Zbořil, R. Carbon Nitride-Based Ruthenium Single Atom Photocatalyst for CO2 Reduction to Methanol. Small 2021, 17, 2006478. [Google Scholar] [CrossRef]
  17. Goyal, S.; Shaharun, M.S.; Jayabal, G.S.; Kait, C.F.; Abdullah, B.; Wei, L.J. Photocatalytic Reduction of CO2 to Methanol Using a Copper-Zirconia Imidazolate Framework. Catalysts 2021, 11, 346. [Google Scholar] [CrossRef]
  18. Izadpanah Ostad, M.; Niknam Shahrak, M.; Galli, F. Photocatalytic carbon dioxide reduction to methanol catalyzed by ZnO, Pt, Au, and Cu nanoparticles decorated zeolitic imidazolate framework-8. J. CO2 Util. 2021, 43, 101373. [Google Scholar] [CrossRef]
  19. Wu, Y.A.; McNulty, I.; Liu, C.; Lau, K.C.; Liu, Q.; Paulikas, A.P.; Sun, C.-J.; Cai, Z.; Guest, J.R.; Ren, Y.; et al. Facet-dependent active sites of a single Cu2O particle photocatalyst for CO2 reduction to methanol. Nat. Energy 2019, 4, 957–968. [Google Scholar] [CrossRef]
  20. Zeng, Z.; Yan, Y.; Chen, J.; Zan, P.; Tian, Q.; Chen, P. Boosting the Photocatalytic Ability of Cu2O Nanowires for CO2 Conversion by MXene Quantum Dots. Adv. Funct. Mater. 2019, 29, 1806500. [Google Scholar] [CrossRef]
  21. Adekoya, D.O.; Tahir, M.; Amin, N.A.S. g-C3N4/(Cu/TiO2) nanocomposite for enhanced photoreduction of CO2 to CH3OH and HCOOH under UV/visible light. J. CO2 Util. 2017, 18, 261–274. [Google Scholar] [CrossRef]
  22. Gao, G.; Jiao, Y.; Waclawik, E.R.; Du, A. Single Atom (Pd/Pt) Supported on Graphitic Carbon Nitride as an Efficient Photocatalyst for Visible-Light Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2016, 138, 6292–6297. [Google Scholar] [CrossRef] [Green Version]
  23. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical Porous O-Doped g-C3 N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef]
  24. Albo, J.; Qadir, M.I.; Samperi, M.; Fernandes, J.A.; de Pedro, I.; Dupont, J. Use of an optofluidic microreactor and Cu nanoparticles synthesized in ionic liquid and embedded in TiO2 for an efficient photoreduction of CO2 to methanol. Chem. Eng. 2021, 404, 126643. [Google Scholar] [CrossRef]
  25. Billo, T.; Fu, F.Y.; Raghunath, P.; Shown, I.; Chen, W.F.; Lien, H.T.; Shen, T.H.; Lee, J.F.; Chan, T.S.; Huang, K.Y.; et al. Ni-Nanocluster Modified Black TiO2 with Dual Active Sites for Selective Photocatalytic CO2 Reduction. Small 2018, 14, 1702928. [Google Scholar] [CrossRef]
  26. Rechberger, F.; Niederberger, M. Translucent nanoparticle-based aerogel monoliths as 3-dimensional photocatalysts for the selective photoreduction of CO2 to methanol in a continuous flow reactor. Mater. Horiz. 2017, 4, 1115–1121. [Google Scholar] [CrossRef]
  27. Wu, J.; Feng, Y.; Logan, B.E.; Dai, C.; Han, X.; Li, D.; Liu, J. Preparation of Al–O-Linked Porous-g-C3N4/TiO2-Nanotube Z-Scheme Composites for Efficient Photocatalytic CO2 Conversion and 2,4-Dichlorophenol Decomposition and Mechanism. ACS Sustain. Chem. Eng. 2019, 7, 15289–15296. [Google Scholar] [CrossRef]
  28. Zeng, G.; Qiu, J.; Li, Z.; Pavaskar, P.; Cronin, S.B. CO2 Reduction to Methanol on TiO2-Passivated GaP Photocatalysts. ACS Catal. 2014, 4, 3512–3516. [Google Scholar] [CrossRef]
  29. Guo, H.; Wan, S.; Wang, Y.; Ma, W.; Zhong, Q.; Ding, J. Enhanced photocatalytic CO2 reduction over direct Z-scheme NiTiO3/g-C3N4 nanocomposite promoted by efficient interfacial charge transfer. Chem. Eng. J. 2021, 412, 128646. [Google Scholar] [CrossRef]
  30. Nor, N.U.M.; Amin, N.A.S. Glucose precursor carbon-doped TiO2 heterojunctions for enhanced efficiency in photocatalytic reduction of carbon dioxide to methanol. J. CO2 Util. 2019, 33, 372–383. [Google Scholar] [CrossRef]
  31. Liu, J.-Y.; Gong, X.-Q.; Alexandrova, A.N. Mechanism of CO2 Photocatalytic Reduction to Methane and Methanol on Defected Anatase TiO2 (101): A Density Functional Theory Study. J. Phys. Chem. C 2019, 123, 3505–3511. [Google Scholar] [CrossRef]
  32. Bahadori, E.; Tripodi, A.; Villa, A.; Pirola, C.; Prati, L.; Ramis, G.; Dimitratos, N.; Wang, D.; Rossetti, I. High pressure CO2 photoreduction using Au/TiO2: Unravelling the effect of co-catalysts and of titania polymorphs. Catal. Sci. Technol. 2019, 9, 2253–2265. [Google Scholar] [CrossRef]
  33. Olowoyo, J.O.; Saini, U.; Kumar, M.; Valdés, H.; Singh, H.; Omorogie, M.O.; Babalola, J.O.; Vorontsov, A.V.; Kumar, U.; Smirniotis, P.G. Reduced graphene oxide/NH2-MIL-125(Ti) composite: Selective CO2 photoreduction to methanol under visible light and computational insights into charge separation. J. CO2 Util. 2020, 42, 101300. [Google Scholar] [CrossRef]
  34. Liu, Y.; Yang, Y.; Sun, Q.; Wang, Z.; Huang, B.; Dai, Y.; Qin, X.; Zhang, X. Chemical adsorption enhanced CO2 capture and photoreduction over a copper porphyrin based metal organic framework. ACS Appl. Mater. Interfaces 2013, 5, 7654–7658. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, Q.; Low, Z.-X.; Li, L.; Razmjou, A.; Wang, K.; Yao, J.; Wang, H. ZIF-8/Zn2GeO4 nanorods with an enhanced CO2 adsorption property in an aqueous medium for photocatalytic synthesis of liquid fuel. J. Mater. Chem. A 2013, 1, 11563–11569. [Google Scholar] [CrossRef]
  36. Alhaddad, M.; Shawky, A. Pt-decorated ZnMn2O4 nanorods for effective photocatalytic reduction of CO2 into methanol under visible light. Ceram. Int. 2021, 47, 9763–9770. [Google Scholar] [CrossRef]
  37. Wang, Y.; Liu, H.; Pan, Q.; Ding, N.; Yang, C.; Zhang, Z.; Jia, C.; Li, Z.; Liu, J.; Zhao, Y. Construction of Thiazolo[5,4-d]thiazole-based Two-Dimensional Network for Efficient Photocatalytic CO2 Reduction. ACS Appl. Mater. Interfaces 2020, 12, 46483–46489. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, Y.; Liu, X.; Han, X.; Godin, R.; Chen, J.; Zhou, W.; Jiang, C.; Thompson, J.F.; Mustafa, K.B.; Shevlin, S.A.; et al. Unique hole-accepting carbon-dots promoting selective carbon dioxide reduction nearly 100% to methanol by pure water. Nat. Commun. 2020, 11, 2531. [Google Scholar] [CrossRef]
  39. Chen, K.; Zhao, X.; Zhang, X.-J.; Zhang, W.-S.; Wu, Z.-F.; Wang, H.-Y.; Han, D.-X.; Niu, L. Enhanced photocatalytic CO2 reduction by constructing an In2O3–CuO heterojunction with CuO as a cocatalyst. Catal. Sci. Technol. 2021, 11, 2713–2717. [Google Scholar] [CrossRef]
  40. McFarlan, A. Techno-economic assessment of pathways for electricity generation in northern remote communities in Canada using methanol and dimethyl ether to replace diesel. Renew. Sustain. Energy Rev. 2018, 90, 863–876. [Google Scholar] [CrossRef]
  41. Bunyakiat, K.; Makmee, S.; Sawangkeaw, R.; Ngamprasertsith, S. Continuous Production of Biodiesel via Transesterification from Vegetable Oils in Supercritical Methanol. Energy Fuels 2006, 20, 812–817. [Google Scholar] [CrossRef]
  42. Williams, R.H.; Larson, E.D.; Katofsky, R.E.; Chen, J. Methanol and hydrogen from biomass for transportation. Energy Sustain. Dev. 1995, 1, 18–34. [Google Scholar] [CrossRef]
  43. Bowker, M. Methanol Synthesis from CO2 Hydrogenation. ChemCatChem 2019, 11, 4238–4246. [Google Scholar] [CrossRef] [Green Version]
  44. Din, I.U.; Shaharun, M.S.; Alotaibi, M.A.; Alharthi, A.I.; Naeem, A. Recent developments on heterogeneous catalytic CO2 reduction to methanol. J. CO2 Util. 2019, 34, 20–33. [Google Scholar] [CrossRef]
  45. Yang, Z.; Qi, Y.; Wang, F.; Han, Z.; Jiang, Y.; Han, H.; Liu, J.; Zhang, X.; Ong, W.J. State-of-the-art advancements in photo-assisted CO2 hydrogenation: Recent progress in catalyst development and reaction mechanisms. J. Mater. Chem. A 2020, 8, 24868–24894. [Google Scholar] [CrossRef]
  46. 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]
  47. Adekoya, D.; Tahir, M.; Amin, N.A.S. Recent trends in photocatalytic materials for reduction of carbon dioxide to methanol. Renew. Sustain. Energy Rev. 2019, 116, 109389. [Google Scholar] [CrossRef]
  48. Fu, Z.; Yang, Q.; Liu, Z.; Chen, F.; Yao, F.; Xie, T.; Zhong, Y.; Wang, D.; Li, J.; Li, X.; et al. Photocatalytic conversion of carbon dioxide: From products to design the catalysts. J. CO2 Util. 2019, 34, 63–73. [Google Scholar] [CrossRef]
  49. Li, D.; Kassymova, M.; Cai, X.; Zang, S.-Q.; Jiang, H.-L. Photocatalytic CO2 reduction over metal-organic framework-based materials. Coord. Chem. Rev. 2020, 412, 213262. [Google Scholar] [CrossRef]
  50. Lei, Z.; Xue, Y.; Chen, W.; Qiu, W.; Zhang, Y.; Horike, S.; Tang, L. MOFs-Based Heterogeneous Catalysts: New Opportunities for Energy-Related CO2 Conversion. Adv. Energy Mater. 2018, 8, 1801587. [Google Scholar] [CrossRef]
  51. Chen, Y.; Jia, G.; Hu, Y.; Fan, G.; Tsang, Y.H.; Li, Z.; Zou, Z. Two-dimensional nanomaterials for photocatalytic CO2 reduction to solar fuels. Sustain. Energy Fuels 2017, 1, 1875–1898. [Google Scholar] [CrossRef]
  52. Wang, C.; Sun, Z.; Zheng, Y.; Hu, Y.H. Recent progress in visible light photocatalytic conversion of carbon dioxide. J. Mater. Chem. A 2019, 7, 865–887. [Google Scholar] [CrossRef]
  53. Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. State of the art and perspectives in heterogeneous catalysis of CO2 hydrogenation to methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. [Google Scholar] [CrossRef] [PubMed]
  54. Maina, J.W.; Pozo-Gonzalo, C.; Kong, L.; Schütz, J.; Hill, M.; Dumée, L.F. Metal organic framework based catalysts for CO2 conversion. Mater. Horiz. 2017, 4, 345–361. [Google Scholar] [CrossRef]
  55. Su, Y.; Zhang, Z.; Liu, H.; Wang, Y. Cd0.2Zn0.8S@UiO-66-NH2 nanocomposites as efficient and stable visible-light-driven photocatalyst for H2 evolution and CO2 reduction. Appl. Catal. B Environ. 2017, 200, 448–457. [Google Scholar] [CrossRef]
  56. Zhang, G.; Wang, Z.; Wu, J. Construction of a Z-scheme heterojunction for high-efficiency visible-light-driven photocatalytic CO2 reduction. Nanoscale 2021, 13, 4359–4389. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  58. Meng, J.; Chen, Q.; Lu, J.; Liu, H. Z-Scheme Photocatalytic CO2 Reduction on a Heterostructure of Oxygen-Defective ZnO/Reduced Graphene Oxide/UiO-66-NH2 under Visible Light. ACS Appl. Mater. Interfaces 2019, 11, 550–562. [Google Scholar] [CrossRef]
  59. Wang, G.; He, C.-T.; Huang, R.; Mao, J.; Wang, D.; Li, Y. Photoinduction of Cu Single Atoms Decorated on UiO-66-NH2 for Enhanced Photocatalytic Reduction of CO2 to Liquid Fuels. J. Am. Chem. Soc. 2020, 142, 19339–19345. [Google Scholar] [CrossRef]
  60. Liu, S.; Chen, F.; Li, S.; Peng, X.; Xiong, Y. Enhanced photocatalytic conversion of greenhouse gas CO2 into solar fuels over g-C3N4 nanotubes with decorated transparent ZIF-8 nanoclusters. Appl. Catal. B Environ. 2017, 211, 1–10. [Google Scholar] [CrossRef]
  61. Maina, J.W.; Schutz, J.A.; Grundy, L.; Des Ligneris, E.; Yi, Z.; Kong, L.; Pozo-Gonzalo, C.; Ionescu, M.; Dumee, L.F. Inorganic Nanoparticles/Metal Organic Framework Hybrid Membrane Reactors for Efficient Photocatalytic Conversion of CO2. ACS Appl. Mater. Interfaces 2017, 9, 35010–35017. [Google Scholar] [CrossRef] [PubMed]
  62. Cardoso, J.C.; Stulp, S.; de Brito, J.F.; Flor, J.B.S.; Frem, R.C.G.; Zanoni, M.V.B. MOFs based on ZIF-8 deposited on TiO2 nanotubes increase the surface adsorption of CO2 and its photoelectrocatalytic reduction to alcohols in aqueous media. Appl. Catal. B Environ. 2018, 225, 563–573. [Google Scholar] [CrossRef] [Green Version]
  63. An, B.; Zhang, J.; Cheng, K.; Ji, P.; Wang, C.; Lin, W. Confinement of Ultrasmall Cu/ZnOx Nanoparticles in Metal–Organic Frameworks for Selective Methanol Synthesis from Catalytic Hydrogenation of CO2. J. Am. Chem. Soc. 2017, 139, 3834–3840. [Google Scholar] [CrossRef] [PubMed]
  64. Li, N.; Liu, X.; Zhou, J.; Chen, W.; Liu, M. Encapsulating CuO quantum dots in MIL-125(Ti) coupled with g-C3N4 for efficient photocatalytic CO2 reduction. Chem. Eng. J. 2020, 399, 125782. [Google Scholar] [CrossRef]
  65. Li, J.; Luo, D.; Yang, C.; He, S.; Chen, S.; Lin, J.; Zhu, L.; Li, X. Copper(II) imidazolate frameworks as highly efficient photocatalysts for reduction of CO2 into methanol under visible light irradiation. J. Solid State Chem. 2013, 203, 154–159. [Google Scholar] [CrossRef]
  66. Chen, S.; Yu, J.; Zhang, J. Enhanced photocatalytic CO2 reduction activity of MOF-derived ZnO/NiO porous hollow spheres. J. CO2 Util. 2018, 24, 548–554. [Google Scholar] [CrossRef]
  67. Goyal, S.; Shaharun, M.; Kait, C.; Abdullah, B.; Ameen, M. Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst. Catalysts 2018, 8, 581. [Google Scholar] [CrossRef] [Green Version]
  68. Nadeem, S.; Mumtaz, A.; Mumtaz, M.; Abdul Mutalib, M.I.; Shaharun, M.S.; Abdullah, B. Visible light driven CO2 reduction to methanol by Cu-porphyrin impregnated mesoporous Ti-MCM-48. J. Mol. Liq. 2018, 272, 656–667. [Google Scholar] [CrossRef]
  69. Becerra, J.; Nguyen, D.-T.; Gopalakrishnan, V.-N.; Do, T.-O. Plasmonic Au Nanoparticles Incorporated in the Zeolitic Imidazolate Framework (ZIF-67) for the Efficient Sunlight-Driven Photoreduction of CO2. ACS Appl. Energy Mater. 2020, 3, 7659–7665. [Google Scholar] [CrossRef]
  70. Burange, A.S.; Gawande, M.B. Role of Mixed Metal Oxides in Heterogeneous Catalysis. In Encyclopedia of Inorganic and Bioinorganic Chemistry; Wiley: Hoboken, NJ, USA, 2016; pp. 1–19. [Google Scholar]
  71. Gawande, M.B.; Pandey, R.K.; Jayaram, R.V. Role of mixed metal oxides in catalysis science—Versatile applications in organic synthesis. Catal. Sci. Technol. 2012, 2, 1113–1125. [Google Scholar] [CrossRef]
  72. Thompson, W.A.; Olivo, A.; Zanardo, D.; Cruciani, G.; Menegazzo, F.; Signoretto, M.; Maroto-Valer, M.M. Systematic study of TiO2/ZnO mixed metal oxides for CO2 photoreduction. RSC Adv. 2019, 9, 21660–21666. [Google Scholar] [CrossRef] [Green Version]
  73. Singhal, N.; Goyal, R.; Kumar, U. Visible-Light-Assisted Photocatalytic CO2 Reduction over InTaO4: Selective Methanol Formation. Energy Fuels 2017, 31, 12434–12438. [Google Scholar] [CrossRef]
  74. Wang, Z.; Jiao, X.; Chen, D.; Li, C.; Zhang, M. Porous Copper/Zinc Bimetallic Oxides Derived from MOFs for Efficient Photocatalytic Reduction of CO2 to Methanol. Catalysts 2020, 10, 1127. [Google Scholar] [CrossRef]
  75. Kumar, P.; Joshi, C.; Barras, A.; Sieber, B.; Addad, A.; Boussekey, L.; Szunerits, S.; Boukherroub, R.; Jain, S.L. Core–shell structured reduced graphene oxide wrapped magnetically separable rGO@CuZnO@Fe3O4 microspheres as superior photocatalyst for CO2 reduction under visible light. Appl. Catal. B Environ. 2017, 205, 654–665. [Google Scholar] [CrossRef]
  76. Gao, S.; Gu, B.; Jiao, X.; Sun, Y.; Zu, X.; Yang, F.; Zhu, W.; Wang, C.; Feng, Z.; Ye, B.; et al. Highly Efficient and Exceptionally Durable CO2 Photoreduction to Methanol over Freestanding Defective Single-Unit-Cell Bismuth Vanadate Layers. J. Am. Chem. Soc. 2017, 139, 3438–3445. [Google Scholar] [CrossRef]
  77. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  78. Jeyalakshmi, V.; Mahalakshmy, R.; Ramesh, K.; Rao, P.V.C.; Choudary, N.V.; Thirunavukkarasu, K.; Krishnamurthy, K.R.; Viswanathan, B. Metal oxides as photo catalysts: Modified sodium tantalate as catalyst for photo reduction of carbon dioxide. Mol. Catal. 2018, 451, 105–113. [Google Scholar] [CrossRef]
  79. Kumar, A.; Prajapati, P.K.; Pal, U.; Jain, S.L. Ternary rGO/InVO4/Fe2O3 Z-Scheme Heterostructured Photocatalyst for CO2 Reduction under Visible Light Irradiation. ACS Sustain. Chem. Eng. 2018, 6, 8201–8211. [Google Scholar] [CrossRef]
  80. Yan, T.; Wang, L.; Liang, Y.; Makaremi, M.; Wood, T.E.; Dai, Y.; Huang, B.; Jelle, A.A.; Dong, Y.; Ozin, G.A. Polymorph selection towards photocatalytic gaseous CO2 hydrogenation. Nat. Commun. 2019, 10, 2521. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, L.; Ghoussoub, M.; Wang, H.; Shao, Y.; Sun, W.; Tountas, A.A.; Wood, T.E.; Li, H.; Loh, J.Y.Y.; Dong, Y.; et al. Photocatalytic Hydrogenation of Carbon Dioxide with High Selectivity to Methanol at Atmospheric Pressure. Joule 2018, 2, 1369–1381. [Google Scholar] [CrossRef]
  82. Kandy, M.M. Carbon-based photocatalysts for enhanced photocatalytic reduction of CO2 to solar fuels. Sustain. Energy Fuels 2020, 4, 469–484. [Google Scholar] [CrossRef]
  83. Lin, J.; Tian, W.; Zhang, H.; Duan, X.; Sun, H.; Wang, S. Graphitic Carbon Nitride-Based Z-Scheme Structure for Photocatalytic CO2 Reduction. Energy Fuels 2021, 35, 7–24. [Google Scholar] [CrossRef]
  84. Gusain, R.; Kumar, P.; Sharma, O.P.; Jain, S.L.; Khatri, O.P. Reduced graphene oxide–CuO nanocomposites for photocatalytic conversion of CO2 into methanol under visible light irradiation. Appl. Catal. B Environ. 2016, 181, 352–362. [Google Scholar] [CrossRef]
  85. Liu, S.-H.; Lu, J.-S.; Pu, Y.-C.; Fan, H.-C. Enhanced photoreduction of CO2 into methanol by facet-dependent Cu2O/reduce graphene oxide. J. CO2 Util. 2019, 33, 171–178. [Google Scholar] [CrossRef]
  86. Xu, D.; Cheng, B.; Wang, W.; Jiang, C.; Yu, J. Ag2CrO4/g-C3N4/graphene oxide ternary nanocomposite Z-scheme photocatalyst with enhanced CO2 reduction activity. Appl. Catal. B Environ. 2018, 231, 368–380. [Google Scholar] [CrossRef]
  87. Kumar, A.; Prajapati, P.K.; Aathira, M.S.; Bansiwal, A.; Boukherroub, R.; Jain, S.L. Highly improved photoreduction of carbon dioxide to methanol using cobalt phthalocyanine grafted to graphitic carbon nitride as photocatalyst under visible light irradiation. J. Colloid Interface Sci. 2019, 543, 201–213. [Google Scholar] [CrossRef]
  88. Li, A.; Wang, T.; Li, C.; Huang, Z.; Luo, Z.; Gong, J. Adjusting the Reduction Potential of Electrons by Quantum Confinement for Selective Photoreduction of CO2 to Methanol. Angew. Chem. Int. Ed. 2019, 58, 3804–3808. [Google Scholar] [CrossRef]
  89. Bafaqeer, A.; Tahir, M.; Amin, N.A.S. Well-designed ZnV2O6/g-C3N4 2D/2D nanosheets heterojunction with faster charges separation via pCN as mediator towards enhanced photocatalytic reduction of CO2 to fuels. Appl. Catal. B Environ. 2019, 242, 312–326. [Google Scholar] [CrossRef]
  90. Olowoyo, J.O.; Kumar, M.; Jain, S.L.; Babalola, J.O.; Vorontsov, A.V.; Kumar, U. Insights into Reinforced Photocatalytic Activity of the CNT-TiO2 Nanocomposite for CO2 Reduction and Water Splitting. J. Phys. Chem. C 2019, 123, 367–378. [Google Scholar] [CrossRef]
  91. Xu, H.; Ouyang, S.; Liu, L.; Reunchan, P.; Umezawa, N.; Ye, J. Recent advances in TiO2-based photocatalysis. J. Mater. Chem. A 2014, 2, 12642–12661. [Google Scholar] [CrossRef]
  92. Kou, J.; Lu, C.; Wang, J.; Chen, Y.; Xu, Z.; Varma, R.S. Selectivity Enhancement in Heterogeneous Photocatalytic Transformations. Chem. Rev. 2017, 117, 1445–1514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Alves Melo Júnior, M.; Morais, A.; Nogueira, A.F. Boosting the solar-light-driven methanol production through CO2 photoreduction by loading Cu2O on TiO2-pillared K2Ti4O9. Microporous Mesoporous Mater. 2016, 234, 1–11. [Google Scholar] [CrossRef]
  94. Wang, W.; Xu, D.; Cheng, B.; Yu, J.; Jiang, C. Hybrid carbon@TiO2 hollow spheres with enhanced photocatalytic CO2 reduction activity. J. Mater. Chem. A 2017, 5, 5020–5029. [Google Scholar] [CrossRef]
  95. Yadav, R.; Sinha, A.K. Titania Cowrapped α-Sulfur Composite as a Visible Light Active Photocatalyst for Hydrogen Evolution Using in Situ Methanol from CO2 as a Sacrificial Agent. ACS Sustain. Chem. Eng. 2017, 5, 6736–6745. [Google Scholar] [CrossRef]
  96. Sharma, A.; Lee, B.-K. Photocatalytic reduction of carbon dioxide to methanol using nickel-loaded TiO2 supported on activated carbon fiber. Catal. Today 2017, 298, 158–167. [Google Scholar] [CrossRef]
  97. Meng, A.; Wu, S.; Cheng, B.; Yu, J.; Xu, J. Hierarchical TiO2/Ni(OH)2 composite fibers with enhanced photocatalytic CO2 reduction performance. J. Mater. Chem. A 2018, 6, 4729–4736. [Google Scholar] [CrossRef]
  98. Seeharaj, P.; Kongmun, P.; Paiplod, P.; Prakobmit, S.; Sriwong, C.; Kim-Lohsoontorn, P.; Vittayakorn, N. Ultrasonically-assisted surface modified TiO2/rGO/CeO2 heterojunction photocatalysts for conversion of CO2 to methanol and ethanol. Ultrason. Sonochem. 2019, 58, 104657. [Google Scholar] [CrossRef]
  99. Zhu, Q.; Xuan, Y.; Zhang, K.; Chang, K. Enhancing photocatalytic CO2 reduction performance of g-C3N4-based catalysts with non-noble plasmonic nanoparticles. Appl. Catal. B Environ. 2021, 297, 120440. [Google Scholar] [CrossRef]
  100. Collado, L.; Reynal, A.; Coronado, J.M.; Serrano, D.P.; Durrant, J.R.; de la Peña O’Shea, V.A. Effect of Au surface plasmon nanoparticles on the selective CO2 photoreduction to CH4. Appl. Catal. B Environ. 2015, 178, 177–185. [Google Scholar] [CrossRef]
  101. Tahir, M.; Tahir, B.; Amin, N.A.S. Synergistic effect in plasmonic Au/Ag alloy NPs co-coated TiO2 NWs toward visible-light enhanced CO2 photoreduction to fuels. Appl. Catal. B Environ. 2017, 204, 548–560. [Google Scholar] [CrossRef]
  102. Mahyoub, S.A.; Hezam, A.; Qaraah, F.A.; Namratha, K.; Nayan, M.B.; Drmosh, Q.A.; Ponnamma, D.; Byrappa, K. Surface Plasmonic Resonance and Z-Scheme Charge Transport Synergy in Three-Dimensional Flower-like Ag–CeO2–ZnO Heterostructures for Highly Improved Photocatalytic CO2 Reduction. ACS Appl. Energy Mater. 2021, 4, 3544–3554. [Google Scholar] [CrossRef]
  103. Zhang, X.; Ke, X.; Yao, J. Recent development of plasmon-mediated photocatalysts and their potential in selectivity regulation. J. Mater. Chem. A 2018, 6, 1941–1966. [Google Scholar] [CrossRef]
  104. Vu, N.N.; Kaliaguine, S.; Do, T.O. Plasmonic Photocatalysts for Sunlight-Driven Reduction of CO2: Details, Developments, and Perspectives. ChemSusChem 2020, 13, 3967–3991. [Google Scholar] [CrossRef]
  105. Jun, H.; Choi, S.; Yang, M.Y.; Nam, Y.S. A ruthenium-based plasmonic hybrid photocatalyst for aqueous carbon dioxide conversion with a high reaction rate and selectivity. J. Mater. Chem. A 2019, 7, 17254–17260. [Google Scholar] [CrossRef]
  106. Verma, R.; Belgamwar, R.; Polshettiwar, V. Plasmonic Photocatalysis for CO2 Conversion to Chemicals and Fuels. ACS Mater. Lett. 2021, 3, 574–598. [Google Scholar] [CrossRef]
  107. Becerra, J.; Gopalakrishnan, V.N.; Quach, T.-A.; Do, T.-O. Enhancing CO2 photoreduction over ZIF-based reticular materials by morphology control of Au plasmonic nanoparticles. Sustain. Energy Fuels 2022. [Google Scholar] [CrossRef]
  108. Yadav, R.; Amoli, V.; Singh, J.; Tripathi, M.K.; Bhanja, P.; Bhaumik, A.; Sinha, A.K. Plasmonic gold deposited on mesoporous TixSi1−xO2 with isolated silica in lattice: An excellent photocatalyst for photocatalytic conversion of CO2 into methanol under visible light irradiation. J. CO2 Util. 2018, 27, 11–21. [Google Scholar] [CrossRef]
  109. Wang, Z.-J.; Song, H.; Pang, H.; Ning, Y.; Dao, T.D.; Wang, Z.; Chen, H.; Weng, Y.; Fu, Q.; Nagao, T.; et al. Photo-assisted methanol synthesis via CO2 reduction under ambient pressure over plasmonic Cu/ZnO catalysts. Appl. Catal. B Environ. 2019, 250, 10–16. [Google Scholar] [CrossRef]
  110. Liu, E.; Qi, L.; Bian, J.; Chen, Y.; Hu, X.; Fan, J.; Liu, H.; Zhu, C.; Wang, Q. A facile strategy to fabricate plasmonic Cu modified TiO2 nano-flower films for photocatalytic reduction of CO2 to methanol. Mater. Res. Bull. 2015, 68, 203–209. [Google Scholar] [CrossRef]
  111. Wang, R.; Ding, J.; Zhong, Q.; Hu, J.; Guo, H.; Liu, Q. Plasmonic Ag Nanoparticles Decorated Acid-Aching Carbon Fibers for Enhanced Photocatalytic Reduction of CO2 into CH3OH Under Visible-Light Irradiation. Catal. Lett. 2021, 151, 3079–3088. [Google Scholar] [CrossRef]
Figure 1. The number of publications for the keyword search “Photocatalytic reduction of CO2 to methanol” as found in Web of Science (dated: 12 November 2021).
Figure 1. The number of publications for the keyword search “Photocatalytic reduction of CO2 to methanol” as found in Web of Science (dated: 12 November 2021).
Catalysts 12 00094 g001
Figure 2. The supporting materials used in photocatalytic CO2 reduction to fuels.
Figure 2. The supporting materials used in photocatalytic CO2 reduction to fuels.
Catalysts 12 00094 g002
Figure 3. (a,b) TEM and HR-TEM images of CZS@UN20; (c) Photocurrent spectra; (d) Methanol production; (e) Cycle runs for reduction of CO2 over the CZS@UN20 photocatalyst; (f) Mechanism for photocatalytic (visible light) CO2 reduction. Reproduced from [55] with permission.
Figure 3. (a,b) TEM and HR-TEM images of CZS@UN20; (c) Photocurrent spectra; (d) Methanol production; (e) Cycle runs for reduction of CO2 over the CZS@UN20 photocatalyst; (f) Mechanism for photocatalytic (visible light) CO2 reduction. Reproduced from [55] with permission.
Catalysts 12 00094 g003
Figure 4. (a,b) SEM and HR-TEM images of OZ/R/U catalyst; (c) Photocurrent spectra; (d) The impact of rGO on photoactivity of the ternary composite (OZ/R/U); (e) Photocatalytic activity; (f) Photocatalytic mechanism over O-ZnO/rGO/UiO-66-NH2 composites. Reproduced from [58] with permission.
Figure 4. (a,b) SEM and HR-TEM images of OZ/R/U catalyst; (c) Photocurrent spectra; (d) The impact of rGO on photoactivity of the ternary composite (OZ/R/U); (e) Photocatalytic activity; (f) Photocatalytic mechanism over O-ZnO/rGO/UiO-66-NH2 composites. Reproduced from [58] with permission.
Catalysts 12 00094 g004
Figure 5. (a) Schematic representation for the synthesis of (CuSA@UiO-66-NH2) photocatalyst; (b,c) The (TEM and HR-TEM) images of CuSA@UiO-66-NH2 photocatalyst; (d) AC-STEM image of CuSA@UiO-66-NH2; (e) EDS mapping; (f) Photocurrent spectra; (g) Photocatalytic activity. Reproduced from [59] with permission.
Figure 5. (a) Schematic representation for the synthesis of (CuSA@UiO-66-NH2) photocatalyst; (b,c) The (TEM and HR-TEM) images of CuSA@UiO-66-NH2 photocatalyst; (d) AC-STEM image of CuSA@UiO-66-NH2; (e) EDS mapping; (f) Photocurrent spectra; (g) Photocatalytic activity. Reproduced from [59] with permission.
Catalysts 12 00094 g005
Figure 6. (a,b) FESEM and TEM images of TCNZ8 sample (ZIF-8/g-C3N4 photocatalyst); (c) Photocurrent spectra; (d) The steady-state PL spectra; (e) Photocatalytic activity. Reproduced from [60] with permission.
Figure 6. (a,b) FESEM and TEM images of TCNZ8 sample (ZIF-8/g-C3N4 photocatalyst); (c) Photocurrent spectra; (d) The steady-state PL spectra; (e) Photocatalytic activity. Reproduced from [60] with permission.
Catalysts 12 00094 g006
Figure 7. (a) A schematic representation for the synthesis of Cu-TiO2/ZIF-8 membranes; (b) SEM image of porous substrate; (c) The high-resolution (HR-TEM) image of the TiO2@ZIF-8 hybrid; (d) Effect of Cu−TiO2 nanoparticle loading on product yields; (e) The impact of membrane content on yield of product. Reproduced from [61] with permission.
Figure 7. (a) A schematic representation for the synthesis of Cu-TiO2/ZIF-8 membranes; (b) SEM image of porous substrate; (c) The high-resolution (HR-TEM) image of the TiO2@ZIF-8 hybrid; (d) Effect of Cu−TiO2 nanoparticle loading on product yields; (e) The impact of membrane content on yield of product. Reproduced from [61] with permission.
Catalysts 12 00094 g007
Figure 8. (a) Sketch of preparation of ZIF-8 nanoparticles over TiO2 nanotube (Ti@TiO2NT-ZIF-8); (b) SEM image of incorporated ZIF-8 on Ti/TiO2NT; (c) TEM image of Ti/TiO2NT-ZIF-8 catalyst; (d) Generation of methanol by photo-electrocatalytic technique; (e) Product selectivity; (f) Mechanism of CO2 reduction using Ti/TiO2NT-ZIF-8 electrode under UV-visible irradiation. Reproduced from [62] with permission.
Figure 8. (a) Sketch of preparation of ZIF-8 nanoparticles over TiO2 nanotube (Ti@TiO2NT-ZIF-8); (b) SEM image of incorporated ZIF-8 on Ti/TiO2NT; (c) TEM image of Ti/TiO2NT-ZIF-8 catalyst; (d) Generation of methanol by photo-electrocatalytic technique; (e) Product selectivity; (f) Mechanism of CO2 reduction using Ti/TiO2NT-ZIF-8 electrode under UV-visible irradiation. Reproduced from [62] with permission.
Catalysts 12 00094 g008
Figure 9. (a) The preparation of CuZn@UiO-bpy catalyst; (b) The precatalyst structure (Zn@UiO-bpy-Cu); (c) Active sites of MOFs; (d,e) STEM-HAADF images of ultrasmall Cu/ZnOx and CuZn@UiO-bpy catalyst (f) Product selectivity. Reproduced from [63] with permission.
Figure 9. (a) The preparation of CuZn@UiO-bpy catalyst; (b) The precatalyst structure (Zn@UiO-bpy-Cu); (c) Active sites of MOFs; (d,e) STEM-HAADF images of ultrasmall Cu/ZnOx and CuZn@UiO-bpy catalyst (f) Product selectivity. Reproduced from [63] with permission.
Catalysts 12 00094 g009
Figure 10. (a) The preparation of g-C3N4/CuO@MIL-125(Ti) composite photocatalyst; (b) CO2 reduction mechanism; (c) Photocatalytic stability and product selectivity; (dg) The morphological analysis of g-C3N4/1% CuO@MIL-125(Ti) catalyst with (SEM/TEM/HR-TEM) images. Reproduced from [64] with permission.
Figure 10. (a) The preparation of g-C3N4/CuO@MIL-125(Ti) composite photocatalyst; (b) CO2 reduction mechanism; (c) Photocatalytic stability and product selectivity; (dg) The morphological analysis of g-C3N4/1% CuO@MIL-125(Ti) catalyst with (SEM/TEM/HR-TEM) images. Reproduced from [64] with permission.
Catalysts 12 00094 g010
Figure 11. (a) Synthesis of rGO@CuZnO@Fe3O4 catalyst; (b) FE-SEM image; (c) low-resolution TEM(LR-TEM); (d) HR-TEM of rGO@CuZnO@Fe3O4 catalyst; (e) Yield of methanol; (f) Recyclability; (g) Methanol yield using Cu wt% in ZnO. Reproduced from [75] with permission.
Figure 11. (a) Synthesis of rGO@CuZnO@Fe3O4 catalyst; (b) FE-SEM image; (c) low-resolution TEM(LR-TEM); (d) HR-TEM of rGO@CuZnO@Fe3O4 catalyst; (e) Yield of methanol; (f) Recyclability; (g) Methanol yield using Cu wt% in ZnO. Reproduced from [75] with permission.
Catalysts 12 00094 g011
Figure 12. (a) The preparation of (Vv-rich and Vv-poor) atomic layers o-BiVO4, (b) Defects characterization positron lifetime spectrum, diagram for trapped positrons; (c) Photocatalytic methanol production (irradiation with visible light). Reproduced from [76] with permission.
Figure 12. (a) The preparation of (Vv-rich and Vv-poor) atomic layers o-BiVO4, (b) Defects characterization positron lifetime spectrum, diagram for trapped positrons; (c) Photocatalytic methanol production (irradiation with visible light). Reproduced from [76] with permission.
Catalysts 12 00094 g012
Figure 13. (a) Schematic representation for preparation of ultrathin Ti3C2/Bi2WO6 nanosheets; (bd) TEM, enlarged image, elemental mapping; (e) FESEM; (f) Transient photocurrent spectra of the prepared catalyst; (g) Photocatalytic activity; (h) Energy profile diagram of Ti3C2 and Bi2WO6 and process electron transfer. Reproduced from [77] with permission.
Figure 13. (a) Schematic representation for preparation of ultrathin Ti3C2/Bi2WO6 nanosheets; (bd) TEM, enlarged image, elemental mapping; (e) FESEM; (f) Transient photocurrent spectra of the prepared catalyst; (g) Photocatalytic activity; (h) Energy profile diagram of Ti3C2 and Bi2WO6 and process electron transfer. Reproduced from [77] with permission.
Catalysts 12 00094 g013
Figure 14. Characterization of In2O3−x(OH)y photocatalyst: (ac) SEM/TEM and HR-TEM images; (d) Stability of catalytic performance; (e) The mechanism for CO2 hydrogenation into CH3OH. Reproduced from [81] with permission.
Figure 14. Characterization of In2O3−x(OH)y photocatalyst: (ac) SEM/TEM and HR-TEM images; (d) Stability of catalytic performance; (e) The mechanism for CO2 hydrogenation into CH3OH. Reproduced from [81] with permission.
Catalysts 12 00094 g014
Figure 15. (a,b) Characterization (SEM/TEM images) of the Ag2CrO4/g-C3N4/GO composite photocatalyst; (c) CH3OH formation of methanol over different samples of CN, under induced sunlight irradiation (1 h); (d) CH3OH and CH4 formation over various catalysts; (e) The Z-scheme mechanism for reduction of CO2. Reproduced from [86] with permission.
Figure 15. (a,b) Characterization (SEM/TEM images) of the Ag2CrO4/g-C3N4/GO composite photocatalyst; (c) CH3OH formation of methanol over different samples of CN, under induced sunlight irradiation (1 h); (d) CH3OH and CH4 formation over various catalysts; (e) The Z-scheme mechanism for reduction of CO2. Reproduced from [86] with permission.
Catalysts 12 00094 g015
Figure 16. (a) Preparation of g-C3N4/CoPc-COOH photocatalyst; (b) HR-TEM images of catalyst; (c) Yield of methanol. Reproduced from [87] with permission.
Figure 16. (a) Preparation of g-C3N4/CoPc-COOH photocatalyst; (b) HR-TEM images of catalyst; (c) Yield of methanol. Reproduced from [87] with permission.
Catalysts 12 00094 g016
Figure 17. (a) Pattern of p-C3N4; (b) Sketch for p-CNCS; (c,d) The HR-TEM images of p-CNCS; (e) Methanol generation selectivity employing with various size of particles; (f) Catalytic performance of as-prepared catalysts. Reproduced from [88] with permission.
Figure 17. (a) Pattern of p-C3N4; (b) Sketch for p-CNCS; (c,d) The HR-TEM images of p-CNCS; (e) Methanol generation selectivity employing with various size of particles; (f) Catalytic performance of as-prepared catalysts. Reproduced from [88] with permission.
Catalysts 12 00094 g017
Figure 18. (a) FESEM image of ZnV2O6/pCN (100%) nanosheets; (b,c) TEM and HR-TEM images of ZnV2O6/pCN (100%) nanosheets; (d) yield of CH3OH over various photocatalysts; (e) yield of CH3OH for UV and visible light against illumination time. Reproduced from [89] with permission.
Figure 18. (a) FESEM image of ZnV2O6/pCN (100%) nanosheets; (b,c) TEM and HR-TEM images of ZnV2O6/pCN (100%) nanosheets; (d) yield of CH3OH over various photocatalysts; (e) yield of CH3OH for UV and visible light against illumination time. Reproduced from [89] with permission.
Catalysts 12 00094 g018
Figure 19. The characterization of (TiO2 and 2.0 CNT-TiO2) catalysts (ad). (e) Structures of CNT-TiO2 composite; (f) Sketch for CNT-TiO2 orbitals; (g) Time-dependent profiles for production of methanol; (h) Correlation of methanol production under UVA and visible light; (i) Schematic sketch exhibit formation of product. Reproduced from [90] with permission.
Figure 19. The characterization of (TiO2 and 2.0 CNT-TiO2) catalysts (ad). (e) Structures of CNT-TiO2 composite; (f) Sketch for CNT-TiO2 orbitals; (g) Time-dependent profiles for production of methanol; (h) Correlation of methanol production under UVA and visible light; (i) Schematic sketch exhibit formation of product. Reproduced from [90] with permission.
Catalysts 12 00094 g019
Figure 20. (a) HR-TEM images of CN; (b) HR-TEM images of mCD/CN nanocomposite; (c) The fundamental insights for high methanol selectivity; (d) Yield of methanol; (e) Schematic diagram for photoreduction of CO2 over the mCD/CN nanocomposites. Reproduced from [38] with permission.
Figure 20. (a) HR-TEM images of CN; (b) HR-TEM images of mCD/CN nanocomposite; (c) The fundamental insights for high methanol selectivity; (d) Yield of methanol; (e) Schematic diagram for photoreduction of CO2 over the mCD/CN nanocomposites. Reproduced from [38] with permission.
Catalysts 12 00094 g020
Figure 21. Characterization of the catalysts: (ad) TEM images of K2Ti4O9, H2Ti4O9, TiO2/Ti4O9 and TiO2/Ti4O9/Cu2O photocatalysts; (e) HR-TEM image of TiO2/Ti4O9/Cu2O photocatalyst; (f) Photocatalytic methanol production over various photocatalysts at 70 °C under artificial solar light. Reproduced from [93] with permission.
Figure 21. Characterization of the catalysts: (ad) TEM images of K2Ti4O9, H2Ti4O9, TiO2/Ti4O9 and TiO2/Ti4O9/Cu2O photocatalysts; (e) HR-TEM image of TiO2/Ti4O9/Cu2O photocatalyst; (f) Photocatalytic methanol production over various photocatalysts at 70 °C under artificial solar light. Reproduced from [93] with permission.
Catalysts 12 00094 g021
Figure 22. (a) The mechanism for the preparation of carbon@TiO2 composite hollow structure; (be) FESEM and TEM images of samples T60 and T120; (f) STEM image of T60 and comparable elemental mapping images of C, O and Ti; (g) The photocatalytic activity of carbon@TiO2 composite samples; (h) Photo-excitation process of the carbon@TiO2 composite. Reproduced from [94] with permission.
Figure 22. (a) The mechanism for the preparation of carbon@TiO2 composite hollow structure; (be) FESEM and TEM images of samples T60 and T120; (f) STEM image of T60 and comparable elemental mapping images of C, O and Ti; (g) The photocatalytic activity of carbon@TiO2 composite samples; (h) Photo-excitation process of the carbon@TiO2 composite. Reproduced from [94] with permission.
Catalysts 12 00094 g022
Figure 23. (a) Schematic representation for the preparation of NiO-TiO2/ACF photocatalyst; SEM images of (b) activated carbon fibers (ACF) and (c) ACF-TiO2; (d) SEM images of NiO-TiO2/ACF photocatalyst; (e) Methanol selectivity test for CO2 conversion to CH3OH and CO; (f) The mechanism for photocatalytic CO2 reduction into methanol by using NiO-TiO2/ACF photocatalyst. Reproduced from [96] with permission.
Figure 23. (a) Schematic representation for the preparation of NiO-TiO2/ACF photocatalyst; SEM images of (b) activated carbon fibers (ACF) and (c) ACF-TiO2; (d) SEM images of NiO-TiO2/ACF photocatalyst; (e) Methanol selectivity test for CO2 conversion to CH3OH and CO; (f) The mechanism for photocatalytic CO2 reduction into methanol by using NiO-TiO2/ACF photocatalyst. Reproduced from [96] with permission.
Catalysts 12 00094 g023
Figure 24. (a) Schematic representation for the synthesis of TiO2/Ni(OH)2 composite; (bd) FESEM, TEM and HR-TEM images of TiO2/Ni(OH)2 (15 wt% of Ni denoted as sample (TN15) composite nanofibers; (e) Photocatalytic activity of as-prepared catalysts; (f) Time-dependent conversion of CO2 to products over TN15 (TiO2/Ni(OH)2;) photocatalyst; (g) Schematic demonstration of transfer and separation of photogenerated charge carriers with TiO2/Ni(OH)2 nanofibre. Reproduced from [97] with permission.
Figure 24. (a) Schematic representation for the synthesis of TiO2/Ni(OH)2 composite; (bd) FESEM, TEM and HR-TEM images of TiO2/Ni(OH)2 (15 wt% of Ni denoted as sample (TN15) composite nanofibers; (e) Photocatalytic activity of as-prepared catalysts; (f) Time-dependent conversion of CO2 to products over TN15 (TiO2/Ni(OH)2;) photocatalyst; (g) Schematic demonstration of transfer and separation of photogenerated charge carriers with TiO2/Ni(OH)2 nanofibre. Reproduced from [97] with permission.
Catalysts 12 00094 g024
Figure 25. (a) HR-TEM image of Cu/ZnO catalyst; (b) Methanol production; (c) N2 sorption isotherms; (d) The mechanism of methanol preparation. Reproduced from [109] with permission.
Figure 25. (a) HR-TEM image of Cu/ZnO catalyst; (b) Methanol production; (c) N2 sorption isotherms; (d) The mechanism of methanol preparation. Reproduced from [109] with permission.
Catalysts 12 00094 g025
Table 1. Reported MOF-based photocatalysts for reduction of CO2 to fuels.
Table 1. Reported MOF-based photocatalysts for reduction of CO2 to fuels.
PhotocatalystLight SourceMethodProductsYield (µmol h−1 g−1)Proton Donor/
Reaction Reagent
YearRef.
Cu-porphyrin MOFsVisible CH3OH (262.6) TEOA2013[34]
Cu (II) ZIFVisibleHydrothermalCH3OH (1712.7)H2O2013[65]
Zn2GeO4/ZIF-8UVHydrothermalCH3OH (0.22)H2O2013[35]
Cd0.2Zn0.8S@UiO-66-NH2VisibleSolvothermal methodCH3OH (6.8)H2O2017[55]
g-C3N4/ZIF-8300 W Xe lampIn situ heterogeneous deposition methodCH3OH (0.75)H2O2017[60]
Cu/ZnOx MOFs-SolvothermalCH3OH (2.59)H22017[63]
Cu-TiO2 ZIF-8UVRapid thermal deposition (RTD)CH3OH, COTEOA2017[61]
TiO2 ZIF-8UV-VisibleLayer-by-layer processCH3OH, CH3CH2OHNa2SO42018[62]
Zn-Ni bimetallic MOF Calcination/ThermalCH3OH 2018[66]
2Cu-ZIF-8N2VisibleHydrothermalCH3OH (35.82)Na2SO32018[67]
Cu-porphyrin-Ti-MCM-48VisibleImpregnationCH3OH (297)Na2SO32018[68]
O-ZnO/rGO/UiO-66-NH2 SolvothermalCH3OH (34.85), HCOOH (6.41)TEOA2019[58]
rGO-NH2-MIL-125(Ti)Visible CH3OHTEOA2020[33]
g-C3N4/CuO@MIL-125(Ti)Visible lightComplexation-oxidation
method
CH3OH, CO, CH3CH2OH, CH3CHO 2020[64]
Aux@ZIF-67Visible CH3OH (2.5), CH3CH2OH (0.5) 2020[69]
CuSA@UiO-66 NH2Visible CH3OH (5.33), CH3CH2OH (4.22)TEOA2020[59]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shinde, G.Y.; Mote, A.S.; Gawande, M.B. Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol. Catalysts 2022, 12, 94. https://doi.org/10.3390/catal12010094

AMA Style

Shinde GY, Mote AS, Gawande MB. Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol. Catalysts. 2022; 12(1):94. https://doi.org/10.3390/catal12010094

Chicago/Turabian Style

Shinde, Gajanan Y., Abhishek S. Mote, and Manoj B. Gawande. 2022. "Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol" Catalysts 12, no. 1: 94. https://doi.org/10.3390/catal12010094

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