Next Article in Journal
Numerical Assessment of Flow Pulsation Effects on Reactant Conversion in Automotive Monolithic Reactors
Previous Article in Journal
Enzymatic Production of Lauroyl and Stearoyl Monoesters of d-Xylose, l-Arabinose, and d-Glucose as Potential Lignocellulosic-Derived Products, and Their Evaluation as Antimicrobial Agents
Previous Article in Special Issue
Improvement of CO2 Photoreduction Efficiency by Process Intensification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Direct Conversion of CO2 into Hydrocarbon Solar Fuels by a Synergistic Photothermal Catalysis

1
School of Materials Science and Engineering, Shanghai Institute of Technology, Shanghai 201418, China
2
School of Chemical and Environment Engineering, Shanghai Institute of Technology, Shanghai 201418, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(6), 612; https://doi.org/10.3390/catal12060612
Submission received: 29 March 2022 / Revised: 24 April 2022 / Accepted: 28 April 2022 / Published: 2 June 2022
(This article belongs to the Special Issue Photocatalytic Reduction of CO2)

Abstract

:
Photothermal coupling catalysis technology has been widely studied in recent years and may be a promising method for CO2 reduction. Photothermal coupling catalysis can improve chemical reaction rates and realize the controllability of reaction pathways and products, even in a relatively moderate reaction condition. It has inestimable value in the current energy and global environmental crisis. This review describes the application of photothermal catalysis in CO2 reduction from different aspects. Firstly, the definition and advantages of photothermal catalysis are briefly described. Then, different photothermal catalytic reductions of CO2 products and catalysts are introduced. Finally, several strategies to improve the activity of photothermal catalytic reduction of CO2 are described and we present our views on the future development and challenges of photothermal coupling. Ultimately, the purpose of this review is to bring more researchers’ attention to this promising technology and promote this technology in solar fuels and chemicals production, to realize the value of the technology and provide a better path for its development.

1. Introduction

With the development of industrialization, energy consumption is also increasing. In the current situation, 85% of primary energy or non-renewable energy resource is still dependent on fossil fuels [1]. The heavy consumption of fossil fuels caused an energy crisis, at the same time, tremendous CO2 growth in the atmosphere has affected the climate system. The annual growth rate of CO2 from 1970 to 2020 is getting faster and CO2 concentration is getting higher [2]. The International Panel on Climate Chang predicted that the global temperature will rise by 1.9 °C in mean value if the concentration of CO2 reaches about 590 ppm by 2100, which will cause serious climate impact [3]. Therefore, climate change’s harm, which threatens both humanity and the environment, urgently requires us to shift away from our current fossil fuel economy toward renewable and carbon-neutral energy technologies [4]. Converting carbon dioxide into valuable chemicals and fuels in the carbon cycle is regarded as the most promising solution to energy and environmental problem [5].
There are at least three ways to reduce CO2 emissions. The first one is to improve the fuel efficiency of existing fossil energy sources or replace fossil fuels with renewables. The second one is to capture CO2 and sequestration. The last is to utilize CO2 as the carbon source to produce value-added chemicals/fuels. Up to now, photocatalytic reduction [6,7,8,9], electrochemical reduction [10,11,12,13], photoelectrochemical reduction [14,15,16,17], thermal chemical reduction [18,19,20,21] and photothermal reduction [22,23,24,25,26] have been proven to be good methods for CO2-conversion. It is well known that CO2-conversion involves multi-electron processes which can lead to different products and selectivity, for example, CO, CH4, CH3OH, HCOOH and so on, single carbon compounds. In addition, multi-carbon compounds formed by the C–C coupling reaction has also been reported [27,28,29]. Therefore, converting CO2 into valuable chemicals and fuels seems to be a crucial path for mitigating global energy shortages and environmental pollution matters.
Compared with electric energy, solar energy has certain advantages, because solar energy is the Earth’s inexhaustible natural energy source, which has the potential to become the ideal renewable source for fuels in the future [30]. Meanwhile, solar fuels have higher energy density and are easier to store and transport [31]. Therefore, researchers use the energy of sunlight to convert CO2 into fuels directly through photocatalyst, but the shortcomings are the limited absorption range of sunlight and the poor utilization rate, limited electron transfer kinetic process, the poor catalytic activity of photocatalyst lead to low conversion rate and poor product selectivity, which impede the large-scale application and remain significant challenges for photocatalytic CO2 reduction [30]. For the above limiting factors, it is necessary to ameliorate a more effective method for CO2 reduction by solar power.
CO2 is an extremely stable molecule with a high bond energy of C=O (750 kJ mol−1) [32]. The implementation of CO2 reduction always requires sufficient energy inputs and appropriate catalysts [33]. Photothermal catalysis can be induced by coupling both the photo and thermal, photo-irradiation can induce photo-generated charge carriers and heating can drive the reaction and accelerate the reaction rate to further enhance the catalytic reaction, which is a promising method for CO2 reduction [34]. Compared with pure-photo or pure-thermal, photothermal remedies the low conversion rate, poor selectivity, slow electrons or hole transfer and fast electron-hole pair recombination of photocatalysis, and substantial heat energy input from thermal catalysis. In other words, photothermal has the characteristics of both photocatalysis and thermal catalysis. Light promotes the generation of electron-hole pairs, while thermal energy provides activation energy which can promote the breaking of old bonds and the formation of new bonds. Meanwhile, thermal energy can increase the speed and kinetic energy of reactant molecules, improving the probability of collisions and energy between molecules. Therefore, photothermal coupling can effectively alleviate the high energy consumption and catalyst deactivation caused by pure thermocatalysis [35]. Our previous works have proved that photothermal coupling can effectively improve the activity of CO2 reduction reactions [36,37,38,39,40]. Therefore, whether from the current research state or long-term development, the photothermal coupling technology to realize the carbon cycle of CO2 is undoubtedly the most potential research and development technology.
Herein, we make this review of photothermal catalysis strategy. The whole review is roughly divided into four parts. The first part will discuss the concept and principle of photothermal catalysis. The second part summarizes some photothermal catalysts for CO2 reduction. Then, main CO2 reduction products are introduced and strategies for controlling products are discussed. Finally, a personal opinion on the developments and application perspectives of photothermal catalytic CO2 reduction into fuels. The purpose of this review is for more researchers to continue studying and improving the photothermal catalytic reduction of CO2 technology. Using photothermal catalysis reduces CO2 technology to achieve carbon neutralization can immediately solve energy shortages and environmental problems.

2. Fundamentals of the Photothermal Catalytic CO2 Reduction Reaction

2.1. Photocatalytic CO2 Reduction

Photocatalytic CO2 reduction is also known as “artificial photosynthesis”. It is similar to photosynthesis in plants. Chlorophyll captures sunlight and uses solar energy to drive H2O and CO2 into O2 and glucose. Photocatalysis is just like the natural photosynthesis process. The overall process of photocatalytic CO2 reduction includes three major steps: (i) absorbing light energy to form photogenerated electron-hole pairs; (ii) photogenerated electron-hole pairs separate and migrate; and (iii) photogenerated electrons or holes react with the surface adsorbed CO2 and H2O at the reactive sites to generate hydrocarbons and oxygen, H+. Firstly, semiconductor materials have an appropriate band structure, which can absorb photons to generate electron-hole pairs by ultraviolet (UV) or Vis, so it is usually used as photocatalysts [41]. Secondly, photogenerated electron-hole pair separation involving electrons will be excited from the valence band (VB) to the conduction band (CB), and there are the same number of holes in the VB; those electrons and holes migrate to the catalyst surface, respectively. During the process of migration, if some electrons and holes are not transferred rapidly, they will recombine or be captured by internal defect traps leading to inactivation [33,42], which is the main factor affecting photocatalytic efficiency. Lastly, surface reaction includes adsorption and activation of CO2 and H2O, CO2 reduction and H2O oxidation and product desorption. Adsorption and activation of the reactants are prerequisites for the succedent redox reaction. Product desorption is also crucial for surface reaction, if untimely desorption of the product occupies the active site, thus terminating the subsequent reaction, and eventually causing the reaction speed to slow down or stop, that is, catalyst deactivation [43]. Therefore, increasing the rapid separation and transfer rates of electron-hole pairs is the main point to improve the efficiency of photocatalytic reduction of CO2 [44]. At the same time, improving the adsorption of reactants, promoting the activation of reactants and accelerating the desorption of products are also essential for CO2 reduction.

2.2. Thermocatalytic CO2 Reduction

Thermocatalytic reduction of CO2 requires an input of heat energy to promote the reaction. For certain chemical reactions, the activation energy is constant, and the higher the temperature, the faster the reaction, which means that for certain chemical reactions, the activation energy (Ea) usually can be driven by thermal energy [33]. Gibbs free energy is a thermal dynamic function for judging the progress of a reaction in chemical thermal dynamics. Thermocatalysis reflects the maximum thermal dynamic tractive force for redox reaction between electrons and holes and reactants [44]. Whether it is thermocatalysis or photocatalysis, the whole process contains the adsorption of reactants, formation of intermediates and desorption of products. The difference is that thermal energy provided by thermocatalysis can activate chemical bonds to facilitate the breaking of old bonds and the formation of new bonds during the reaction process under the activation of appropriate catalysts, which determine the form of the intermediates. Moreover, different intermediates have different reaction paths, which dynamically influence the products and product selectivity [45]. Therefore, intermediates are decisive for the whole reaction. Thermocatalysis also provides more advantages of favorable kinetics for CO2 reduction [46].

2.3. Photothermal Catalytic CO2 Reduction

Photothermal catalysis is one of the typical new catalytic methods and a current focus of photothermal technology research; as a highly efficient catalytic technology integrating photochemistry and thermochemistry, it has received extensive attention in the field of CO2 conversion [47]. Photothermal catalytic CO2 reduction can be understood as the product of the unity between the thermal acceleration of photocatalysis and the photo enhancement of thermocatalysis [44]. The uniqueness of photothermal catalysis is that it is not a simple superposition effect with conventional solar-driven photocatalytic and thermodynamics-driven thermocatalytic activity, but a synergistic effect of the two reactions promoting each other. This synergistic effect between light energy and thermal energy can effectively improve the solar energy harvesting ability, including low-energy visible light and near-infrared light; improve the catalytic activity and product selectivity for the reaction; and promote certain reactions that can only react at high temperature and be carried out at relatively low temperature [48]. In summary, compared with a single photocatalytic system, photothermal catalysis provides a more efficient conversion pathway to convert CO2 into carbon-containing fuels, because the existence of the thermal effect broadens the absorption range of the solar energy spectrum and accelerates the separation and transfer efficiency of photogenerated electron-hole pairs as well as solves the problem of slow reaction speed in photocatalytic process. At the same time, compared with single thermocatalysis, the introduction of light energy can greatly reduce the harsh conditions of high energy consumption and high temperature and high pressure in the thermocatalysis process, and can also reduce material costs and alleviate catalyst deactivation. Due to the different subjects driving the whole reaction, the photothermal catalytic reaction can be divided into three categories: photo-assisted thermocatalysis, thermal-assisted photocatalysis and photothermal co-catalysis. In addition, the heat sources in photothermal catalysis can also be divided into three types: external heating device; the local heating caused by the absorption of photon energy by the catalyst through lattice vibration or electronic oscillation; and plasmon resonance effect. These will be described in detail in the following sections.

2.4. Thermal Classification of Photothermal Catalytic CO2 Reduction

Nowadays, the thermal/heat sources in photothermal catalysis can be divided into three different types. One type is heated from the external heating device;the other is the local heating caused by the absorption of photon energy by the catalyst through lattice vibration or electronic oscillation; and the the third type is heated by plasmon-induced heating.
The external heating device is the most direct and easiest way to heat. Generally, the input of heat energy is directly realized by installing a heating table at the bottom of the reactor or wrapping the reactor with a heating jacket/heating belt and adjusting the temperature according to the actual reaction. In a study on thermal coupled photocatalysis to enhance CO2 reduction activities on Ag-loaded g-C3N4 catalysts [49], the photochemical-thermal CO2 reduction activities of the catalysts were directly evaluated in a stainless steel reactor equipped with a quartz window and a device for controlling the reaction temperature.
A semiconductor absorbs light energy close to or greater than its band gap to generate electron-hole pairs. The exciting high-energy electrons will return from a higher energy state to a lower energy state in two forms. One form is that photogenerated carriers recombine through radiative relaxation, where the process will cause energy loss. The other is that the excited electrons in the form of photons increase the temperature through non-radiative relaxation, while local heating is generated by lattice vibrations. The thermal vibrations of the molecules also cause the molecules to generate heat. In one study, an asymmetric Zn-O-Ge triple atom site was designed to trigger CO2 reduction using light radiation [50]. The results showed that inducing light increases the thermal vibration of the molecule, and the heat generated by the thermal vibration of the molecule promotes the C–C coupling and the evolution of hydrogen species, which greatly speeds up the rate-determining step.
Localized surface plasmon resonance (LSPR) means that when light irradiates on the surface of plasmonic nanomaterials; if the incident photon frequency matches the overall vibration frequency of electrons on the surface of the material, the plasmonic material will have a strong absorption effect on the photon energy, which will cause localized surface plasmon resonance [51,52]. The plasmon resonance phenomenon can obtain heat energy from sunlight. Plasma catalysts can induce localized heating and provide heat carriers to initiate CO2 reduction reactions and/or modulate reaction pathways, leading to high solar-to-fuel conversion efficiency [53,54]. In addition, the photocatalytic performance of photocatalyst can be further increased by enhancing light scattering, hot electron injection, inducing a strong local electric field and heating the surrounding environment to increase the oxidation-reduction reaction rate, mass transport and polarize adsorbed molecules on the surface of photocatalytic materials [55,56]. The localized surface plasmon resonance effect can control the spectral response range of the photocatalytic system by adjusting the composition, morphology and medium environment of nanoparticles [57]. Lian et al. observed a broadened absorption peak in the UV-visible spectrum for Au/CdSe nanorod composite, which suggested a strong orbital hybridization between both Au and CdSe components and transient absorption spectroscopy (TAS) data revealed fast electron transfer efficiency [58]. Ag/TiO2, in a photocatalytic process, converted CO2 and CH4 into C2H4 and CO. The success relies on hot electrons and holes that can be generated by a visible light-induced SPR effect over Ag and UV light-induced photoelectric effect over TiO2. Hot electrons can overcome the Schottky barrier and inject into TiO2 to combine with photogenerated holes; the hot holes on Ag could be captured by CH4. The synergy achieved by the Ag-SPR effect and selective adsorption of CH4 on Ag indicates that ethylene is generated on the surface of Ag [59].

2.5. Classification of Photothermal Catalytic CO2 Reduction

The photo-assisted thermocatalytic system is mainly based on thermocatalysis. Light energy promotes thermocatalysis by increasing the local temperature of the catalyst surface through illumination, rather than through heat conduction. Bai reported on P-Mo2C applied to light-driven thermocatalytic CO2 reduction with H2O. The surface temperature of the catalyst was found to be as high as 140 °C when the catalyst was irradiated by light. Compared with the thermocatalytic activity without light irradiation, the catalytic stability of the P-Mo2C in the light-driven thermocatalytic activity was significantly improved and exhibited good photothermal conversion effect and catalytic activity, which could be attributed to the light-promoted removal of surface hydroxyls that derived from the reaction of CO2 with H2O [60]. m-CN@CsPbBr3 can effectively drive the thermocatalytic reduction of CO2 with H2O. By coupling light into the system, the activity for CO2 to CO reduction is further improved under lower temperatures (150 °C) compared to pure photocatalytic and thermocatalytic activity. This result showed that light excitation can facilitate the thermocatalytic CO2 reduction process, adjust the adsorption of the reaction intermediates of m-CO32− and promote the transformation of b-HCO3 to CO2 on the surface. Finally, CO2 releases CO when it encounters Hads and free electrons. It can be concluded that photo-assisted thermocatalytic activity can effectively improve the stability of the catalyst, alleviating catalyst deactivation, promoting the transformation of reactants or intermediate and reducing the temperature of thermal catalytic reaction [61].
Thermal-assisted photocatalysis is mainly based on photocatalysis; light is the primary driving force for the reaction, while it is helpful to promote a photocatalytic reaction rate rather than drive thermocatalysis. The input of thermal energy can effectively improve the utilization rate of sunlight, promote the excitation and separation of carriers, accelerate the diffusion of reaction molecules and increase the reaction rate [62]. Pt/TiO2 serving as a photothermal catalyst has shown relatively high performance and selectivity for CH4 when using the thermally-assisted photocatalytic reduction of CO2 with H2O. The reason is that thermal energy with the photocatalytic process can enhance the generation of electrons and holes and increase charge carrier transfers to the Pt nanoparticles at the surface of the supports, resulting in the increased formation rates of the products. Moreover, the thermal energy in the photothermal catalytic process assists in the splitting of H2 molecules by the Pt nanoparticles, which subsequently facilitates the reaction between monatomic H· with CO [63].
Photothermal co-catalytic activity is the co-participation of light energy and thermal energy in a catalytic reaction, and the mutual synergy promotes the occurrence of the reaction. In this system, single light energy or thermal energy can drive the reaction, but the photothermal coupling can accelerate the reaction and improve the catalytic activity. Pr3+-doped La1−xPrxMn0.6Ni0.4O3−δ is an efficient artificial photosynthesis catalyst for solar CH3OH; La0.4Pr0.6Mn0.6Ni0.4O3−δ presents the best catalytic activity. Under the pure photocatalytic mode, although the yield of CH3OH was not enough high, a study found it to have good selectivity (80%) for CH3OH. Under the pure thermocatalytic activity, the yield and selectivity for CH3OH were higher than pure photocatalytic, and the product yield and selectivity increased with the increase in temperature. When the temperature reached 300 °C, CH3OH had the highest selectivity (87.8%) and yield (1.85 mmol). However, under photothermal coupling conditions, the yield of methanol was up to 3.97 mmol, which was about two times that of pure heat, and the selectivity reached 98%. The results show that photothermal co-catalytic activity can improve catalytic performance [64].

3. Basic Products of Photothermal Catalytic

Photothermal catalytic CO2 reduction is a complex multistep reaction, which involves multiple electron transfers and reactions. CO2 is a linear molecule with the highest oxidation state of carbon (C4+), which can generate various products in the presence of reducing agents by obtaining different numbers of electrons. Compared with H2 or CH4, H2O may be a good candidate for reducing agents, because of its richness, no toxicity and effectiveness, which can be an H+ source to provide protons [65]. The most possible CO2 reaction reduction in the aqueous medium with potentials versus normal hydrogen electrodes are mentioned below:
CO2 + 2H+ + 2e → CO + H2O E0 = −0.53 V
CO2 + 2H+ + 2e → HCOOH E0 = −0.61 V
CO2 + 4H+ + 4e → HCHO + H2O E0 = −0.48 V
CO2 + 6H+ + 6e → CH3OH + 2H2O E0 = 0.38 V
CO2 + 8H+ + 8e → CH4 + 2H2O E0 = −0.24 V
CO2 + 8H+ + 8e → CH4 + 2H2O  E0 = −0.24 V
2CO2 + 12H+ + 12e → C2H5OH + 3H2O E0 = −0.33 V
2CO2 + 14H+ + 14e → C2H6 + 4H2O E0 = −0.27 V

3.1. CO

Carbon oxide (CO) is both oxidizing and reducing in terms of chemical properties, so it is often chosen as a model for product research. From the point of chemical transformation, CO is preferred because it is an unsaturated metastable molecule, which can be converted into other organic compounds as reactants or primary products, or intermediates under the action of catalysts under different reaction conditions. Therefore CO has a certain application potential and value [66]. Ag@Cr/CaO/CaGa4O7/Ga2O3 photocatalyst has been successfully synthesized and has demonstrated a satisfying formation rate of CO (>835 mol·h−1) and excellent selectivity (95%) by artificial photosynthesis for the conversion of CO2 with H2O [67]. The high formation rate and selectivity for CO are related to the increasing concentration of CO2-related species around the active sites of Ag and the heterojunction formed by Ga2O3/CaGa4O7, which can promote the spatial separation efficiency of photogenerated carriers. Ultrathin two-dimensional Bi4O5Br2 nanosheet (Bi4O5Br2-UN) catalyst has demonstrated an outstanding photocatalytic activity and selectivity for CO2; the formation of CO reaching 63.13 µmol·g−1 and selectivity over 99.5% in the whole catalytic process, which is 2.3 times higher than bulk Bi4O5Br2 (27.56 µmol·g−1); so, changing catalyst structure has significance in improving catalytic activity [68]. In other research, Ag/SrNb2O6 nanorods exhibited a better CO formation rate (287 µmol·g−1) and selectivity (94.1%) in aqueous solutions when NH4HCO3 was used as an additive. From previous experimental results and theoretical analysis, it can be seen that the actual CO2 reactant in the photocatalytic reaction comes from the dissociation of HCO3, and the existence of HCO3 can facilitate the conversion of CO2 in aqueous solutions [69].

3.2. CH4

Methane (CH4) is widely distributed in nature and is one of the main components of natural gas and biogas. It can be used as fuel and raw materials for the production of hydrogen, carbon monoxide, carbon black, acetylene, hydrofluoric acid and formaldehyde. Ulmer reported four methods for the application of photocatalytic CO2 methanation, which were photothermal and plasmon-driven, biophoton-catalytic, heterogeneous photoredox and homogeneous photoredox catalysis (Figure 1a) [70]. It should be noted that H2O and CO2 also have been definitively demonstrated in photothermal or plasmon-driven CO2 methanation. Herein, photothermal or plasmon-driven CO2 methanation will only be discussed. Two possible pathways for CO2 to CH4 were also listed in Figure 1b [71]. For the formaldehyde pathway, carboxyl radical recombines with a hydrogen radical H· to form HCOOH and then HCOOH accepts H· to form a dihydroxymethyl radical, which dehydrates upon attachment of another H· to form HCOH. Two more steps will lead to CH3OH, which is further reduced in another two steps to CH4 [72]. For the carbene pathway, H· attached to the oxygen atom of the CO2 radical, caused an immediate cleavage of the bond between this oxygen atom and the carbon atom to form CO. CO accepts two electrons to form the C atom as with the above step [73]. These radicals can then combine with H· to form CH· radical, CH2· radical, CH3· radical and eventually, CH4. If the CH3· radical combines with the OH· radical, CH3OH will form. There are two important steps in these two methods. The first step is that HCOOH, HCHO and CH3OH are the intermediate steps in the formaldehyde pathway, and CO and C are the intermediate steps in the carbene pathway. The second step is that CH3OH is intermediate in the formaldehyde pathway, but a product in the carbene pathway. Pd-loaded WN-WO3 heterostructure as a direct Z-scheme catalyst broadens the absorption solar spectrum for photothermal assisted conversion CO2-H2O into CH4. The conversion activity reaches 40.6 ± 0.7 mmol·h−1·g−1 for CH4. The main reason for increasing catalytic activity is that the unique carrier transfer mode of Z-scheme catalyst. Under sunlight illumination, WN-WO3 heterostructure can separate and migrate photoinduced carrier efficiency, resulting in more effective electrons involved in the reaction. The local photothermal effect accelerates the shifting of gas molecules and carriers, which improves the CO2 conversion rate [74]. In this study, Ni NPs supported on barium titanate perovskite were prepared for photothermal conversion of CO2. Under the optimum reaction conditions (UV-visible light irradiation), CO2 with H2 was converted into CH4 with yield rates of 103.7 mmol·h−1·g−1 and the selectivity for CH4 is nearly 100%. Mechanism studies indicated that a nonthermal hot-electron-driven pathway dominates reactivity in coexistence with a minor thermal contribution to the photothermal process [75]. The other article reported that strong interaction between metal and semiconductor can alter the reaction pathways of intermediate species during CO2 conversion, leading to dramatic changes in product selectivity. The authors hypothesize that it is reasonable to add hydrogen one by one to CO2 and suggest several possible pathways, which are shown in Figure 2. The hydrogenation on the C atom of CO2 has four possible paths. It can be seen that the initial reaction CO2 → CO2* → HCOO* → H2COO* → H2COOH* was the same for P1, P2, P3 and P4 four reaction paths. The whole reaction path for P1 and P4 are similar, but the difference is that the intermediate state in the previous step before CH4 generation is different. P1 was H3C* → CH4, while P4 was H4COH* → CH4. This was the same for P2 and P3. Meanwhile, P1 and P3, P2 and P4 have different reaction pathways but the similar point is that they have the same intermediate state before producing methane. The paths for P5 and P6 are different from the H3COOH* → CH4 pathway, which are the cross-hydrogenation of O and C atoms. P5 has the second dehydroxylation process [76].

3.3. CH3OH

Methanol (CH3OH) is one of the basic raw materials of many organic chemical products and has important commercial value. Photothermal catalytic CO2 to CH3OH has been a research hotspot in recent years. There are two ways for CH3OH synthesis to occur from CO2: direct hydrogenation and CO2 with H2O into CH3OH [77,78,79,80,81].
2CO2 + 4H2 → 2CH3OH + O2
4CO2 + 2H2O → 4CH3OH + 3O2
Direct hydrogenation using heterogeneous catalysis has been studied extensively and seems to be much easier. CO2 with H2O is one of the most tempting routes. For example, the CoO/Co/TiO2 catalyst used in catalyzing CO2 to CH3OH under photothermal conditions, it has been shown to be as high as 99.9% for selectivity and catalytic activity (39.6 μmol gcat−1·h−1) [82]. From these results, a Z-scheme heterojunction photothermal catalytic system showed high activity for CH3OH formation and a mechanism for photothermal catalytic CO2 over CoO/Co/TiO2 catalyst was proposed (Figure 3a). It proposed that the entire photothermal catalytic reduction of CO2 to CH3OH was at the CB of CoO instead of TiO2 because the potential of the CB of TiO2 (−0.14 eV) is more negative than the standard redox potential E(CO2/CH3OH) = −0.38 eV, so the conversion of CO2 to CH3OH cannot occur. However, TiO2 also had a certain promoting effect on the reaction because one part of the electrons involved in the reaction was provided by CoO and the other part of electrons was provided by TiO2. TiO2 can successfully transfer electrons to the VB of CoO and the surface of Co through the Co0 species. It can indirectly inhibit the recombination of CoO and TiO2 electron-hole pairs and promote the occurrence of CO2 reduction reaction.Meanwhile, it also can promote the thermal reduction of CO2 to CH3OH on Co, thus improving the efficiency of CO2 conversion to CH3OH, based on the good catalytic performances for CH3OH formation originated from the special Z-scheme structure and the high temperature of the reaction. Wu reported a series of hexagonal tungsten bronze M0.33WO3 (M = K, Rb, Cs) photocatalyst showed unexpected photocatalytic CO2 activity directly from air at ambient pressure under full spectrum (UV, visible and NIR lights). After a 4-h reaction, Cs0.33WO3, Rb0.33WO3 and K0.33WO3 exhibited the production of CH3OH to be 17.50, 15.10 and 7.70 μmol·g−1, respectively (Figure 3b). Both Cs0.33WO3 and Rb0.33WO3 exhibited outstanding CH3OH production rates under NIR light irradiation (Figure 3c). After the reaction, the new signals of C−H and O−H functional groups can be clearly seen through the FTIR spectrum, which can prove that the existence of CH3OH really comes from the photocatalytic reduction of carbon dioxide, and the production rate of methanol increases with the increase in temperature (10 °C, 60 °C, 150 °C). Combined with experiments and DFT analysis, it can be concluded that the {010} facets are more conducive to the adsorption of CO2, and the introduction of alkali metals promotes the adsorption capacity of CO2 and inhibits the adsorption of O2. At the same time, the introduction of alkali metals into hexagonal tungsten bronze crystal regularly optimized and reconstructed the electronic structure, particle size, morphology and light capture ability of the M0.33WO3 series, which can provide a large number of free electrons to the crystal. In addition, it also enhanced the transition of polarized electrons, changed the energy band structure of the material and reduced the activation energy of CO2 reaction. Finally, it promotes the conversion of CO2 to CH3OH in the air as a reality [83]. Recently, the Cu/Zn/Al2O3 catalyst achieved higher activity and selectivity for CO2-CH3OH under simultaneous heat (225 °C) and UV-Vis light (350–800 nm) than the thermal catalytic system alone (275 °C); this result showed that light can promote the catalytic reduction of CO2-CH3OH over Cu/Zn/Al2O3 catalyst. Further research found that CO2 initially preferentially adsorbed on ZnO, and ZnO as the main active site promoted the conversion and activation of CO2. The presence of Cu could facilitate the conversion of ZnO-HCOO* to CH3OH because the hydrogenation of HCOO* → H3COO* was mainly driven by Cu. The results further indicate that Cu/ZnO interface adsorption is the key to synthesis of CH3OH. Figure 3d delineates the roles of the ZnO surface, Cu surface and ZnO-Cu interface in the Cu/ZnO/Al2O3 catalyst and the mechanism for the photo-driven CO2 hydrogenation process [84].

3.4. C2+

Although photothermal catalytic CO2 reduction to hydrocarbons has been successful, the obtained products mainly focused on C1 chemicals with low added value. The synthesis of higher value-added C2+ hydrocarbons products via a photothermal route is also important. Up to now, there have been many researchers who have successfully obtained C1-C3 hydrocarbons from CO2 hydrogenation and CO2 and H2O. Zhang and colleagues reported a dual-function Au-Ru/TiO2 catalyst for the simultaneous decomposition of H2O to H2 and CO2 hydrogenation, which can convert CO2 and H2O to fuels by the photothermal coupling method (Figure 4a). In this paper, the thermocatalytic CO2 hydrogenation, photothermal catalytic CO2 hydrogenation, photothermal catalytic H2O splitting to H2 and photothermal catalytic reduction of CO2 and H2O were studied. Au/TiO2 showed a high CO selectivity and production rate in thermocatalytic CO2 hydrogenation. Ru/TiO2 showed a high formation rate of CH4 and C2H6. The same as Au-Ru/TiO2, but its activity was not as good as Ru/TiO2 due to the partial blocking effect of Au. However, the production rates of CH4 and C2H6 increased with the increase in temperature. Au-Ru/TiO2 revealed much higher activity under photothermal catalytic conversion of CO2 + H2. The results indicated that light illumination can improve thermal catalytic activity. In addition, Au-Ru/TiO2 can also decompose H2O to H2 under similar conditions as CO2 hydrogenation, but the difference is that the production rate of H2 is greatly affected by temperature. The optimal thermal temperature can achieve the maximum rate of H2 production, and the reaction rate of CO2 hydrogenation is higher than that of H2O to H2. Finally, combined with the above experiments, the photothermal coupling catalytic reduction of CO2 and H2O over Au-Ru/TiO2 catalyst was studied directly. The results showed that the efficient thermocatalytic CO2 hydrogenation can be realized on Au-Ru/TiO2, so H2 generated from H2O by photocatalysis can directly react with CO2 to generate alkanes. The whole reaction process can greatly accelerate the migration of photogenerated electrons and avoid the recombination of photogenerated electron-hole pairs, and ultimately improve the collaborative reduction of CO2 and H2O by photocatalytic and thermocatalytic activity [85]. At present, most of the catalysts to generate multi-carbons need the support of noble metals or through the combination of metal elements and various oxide catalysts or the construction of a complex structure of the catalyst to achieve this. Our group synthesized a proton conductor BZCY532, under the photothermal coupling conditions successfully converted CO2 and H2O into CH4, C2H6, and C3H8 [39]. Although the yield of C2H6 and C3H8 were not high and the catalytic mechanism was not very clear, the realization of multi-carbon conversion of CO2 through cheap materials and simple structure is worth further study in terms of economy and cost. Deng introduced WO3−x, which contains oxygen vacancies and exhibits an excellent photothermal conversion efficiency [86]. WO3−x produced 5.30 and 0.93 μmol·g−1 C2H4 and C2H6, respectively, after 4 h. The results showed that oxygen defects can expand the light absorption range, improve the carrier separation efficiency and enhance the H2O activation and hydrogenation of adsorbed CO. Under UV-Vis-IR irradiation, solar-energy-driven photothermal catalytic C–C coupling occurs from CH2/CH3 intermediates (Figure 4b). Table 1 summarizes the photothermal catalytic product.

4. Photothermal Catalysts

The photothermal catalysts are the key to achieving high-efficiency photothermal catalytic activity and selectivity. An efficient photothermal catalyst should possess the capacity of both catalytic properties: (1) harvesting the full sunlight spectrum, especially for visible light and even for infrared light; (2) promoting the effective separation and migration of electron-hole pairs; and (3) reducing the activation energy (Ea) for chemical reaction progresses. So far, there have been many different types of photothermal catalysis, but all photothermal catalysts can be systematically classified into three categories, plasmon-mediated (Pt, Au, Ag and Cu) catalysts, not-plasmon-mediated and other common catalysts.

4.1. Metal Oxides Photothermal Catalyst

Metal oxides as catalysts have been widely studied in photocatalytic and photothermal catalytic processes for CO2 due to their facile preparation, low cost and tunability, such as TiO2, ZnO, NiO, Fe2O3, V2O3, MnO2, WO3 and MoO3. The catalytic principle of metal oxide catalyst and its influence on CORR can be referred to in the literature [101,102,103]. The lattice structure, electronic structure, surface properties, volume defects and metal-oxygen bond strength of metal oxides directly affect the catalytic performance [104]. The poor catalytic activity of single metal oxides is due to the low efficiency of photogenerated charge separation and transfer and the lack of active sites on the surface for adsorption and activation of reactant molecules [101]. Therefore, the catalytic activity and product selectivity for the catalysts can be changed by modifying metal oxide catalysts, including introducing oxygen vacancies [105], doping [106], loading noble metal [107], coupling with other metal oxides [108] to adjust the material lattice structures, electronic structures, lattice defects and the diffusion. In our former work, a series of mesoporous WO3 (m-WO3−x) with different oxygen vacancies were synthesized after hydrogenation treatment at different temperatures (350 °C, 450 °C and 550 °C). Under the same photothermal coupling conditions (300 W Xe lamp with a UV-light filter (λ > 420 nm), 250 °C, 25 Kpa), the CH4 evolution rate reached 25.77 μmol·g−1 by m-WO3-H550, which is 36 and 22 times higher than com-WO3 (1.17 μmol·g−1) and m-WO3 (1.17 μmol·g−1), so the good photothermal catalytic activity is closely related to the formation and concentration of oxygen vacancies (Figure 5). A mechanism for the catalytic reduction of CO2 by hydrogenated m-WO3−x was proposed; a high oxygen vacancy concentration of m-WO3−x can reduce CO2 to C, which is deposited on the catalyst surface and reacts with H to generate CH4, but a low oxygen vacancy concentration can reduce CO2 to CO, which is dispersed in the system and reacts with H2 to generate CH3OH. Therefore, the formation of oxygen vacancy can effectively improve the photothermal coupling efficiency and promote the reduction of CO2 to generate hydrocarbons [37]. Li reported a variety of W-doped TiO2 (W doping concentrations ranging from 2 to 10%) for photothermal catalytic CO2 reduction (150 W UV lamp, 393 K); the results showed that doping can effectively enhance catalytic activities. However, appropriate doping is beneficial to improve the catalytic activity, because doping too low is not conducive to the formation of oxygen vacancies and thus lacking active sites required for CO2 reduction; doping too high may cause the undoped ions to react with the anions on the catalyst to generate other substances, thus inhibiting the occurrence of CO2 reduction [109]. In this research, when the W doping concentration reached 4%, it exhibited 3.5 times higher catalytic activity than pure TiO2, which further proved that doping concentration influences the catalytic activity of the catalyst. Although the photothermal catalytic activity of single metal oxide is not high, the catalytic activity can be improved by loading, doping or multi-catalyst combination. Table 2 summarizes the metal oxide photothermal catalysts and products.

4.2. Plasmonic-Metal Photothermal Catalyst

In plasma catalysis, the free electrons of the metal catalyst resonate with the electric field formed by the incident light to generate collective oscillation, thus promoting the migration of electrons in the metal catalyst to the catalyst surface to participate in the catalytic reaction [122]. Surface plasmon resonance (SPR) is widely used in CO2 reduction. Based on noble metals, the SPR effect has a great improvement in catalytic activity. Noble metals such as Au, Ag, Cu and Pd nanoparticles are promising for harvesting photon energy for chemical reactions due to their extraordinary and tailorable LSPR properties [48,123]. Metal nanostructured catalysts absorb light energy through in-band or inter-band transitions to generate abundant energetic charge carriers (hot carriers or electron-hole pairs). In one study, electron-electron or electron-photon caused local heating compared with traditional thermal catalytic heating, which offers a relatively mild condition and consumes less energy [122]. It has also been shown that the hot carriers can help the activation of chemical bonds and the transformation of intermediate species through excited transient electrons, which can influence the product’s selectivity and expedite redox reactions by whittling certain chemical bonds [124]. Plasmon resonance as an attractive alternative to conventional heating reaction with full-spectrum absorption has attracted attention in the study on photothermal catalytic reduction of CO2. With certain illumination, light can cooperate with the thermal and non-thermal mechanisms for conventional thermal catalysts or specific plasma photocatalysts to improve the overall chemical reaction rate and product selectivity at relatively low temperatures. To understand the catalytic properties for the plasma resonance effect of noble metal in driving catalytic reaction, Rh/TiO2 photocatalyst was used to demonstrate and illustrate the application of non-thermal effects in plasmon to improve CO2 conversion into CH4. It was found that Rh/TiO2 enhanced the process of CO2 methanation due to the plasma resonance effect, and non-thermal CH4 production rate and apparent quantum efficiency are significantly correlated with top-surface temperature. Under medium and low temperatures, the synergistic effects of light and heat can accelerate the generation of CH4, but when the temperature rises to 350 °C, begins to act as the dominant reaction to affect the occurrence of light-induced reaction because light enhances the occurrence of the reverse reaction of CO2 reduction reaction [125]. Zhao et al. [76] used plasmonic photocatalysis (Au, Ag and Pd/3D porous ZnO nanosheets). To effectively promote chemical polarization and activation of inert molecules, which was designed specifically to expose the polar {001} facet, based on the physical prototype of field-field coupling, to benefit chemical polarization and activation of the inert molecule (Figure 6). The characterization results indicated that the M/m-ZnO-x NSs array catalysts show a good photocatalytic performance of the CO2-to-hydrocarbon (Figure 7) with CO, CH4 and C2H6 as hydrocarbon products. However, Au/m-ZnO-4.6 photocatalyst generated a large number of C2H6 besides CH4 and CO, Ag/m-ZnO-x and Pd/m-ZnO-x photocatalysts which predominantly generated CO and CH4, respectively. The thermal dynamic analysis with DFT calculations found that the loading of plasmonic metal cocatalysts can reduce the thermal dynamic barrier and alter the molecular pathways; consequently, the product selectivity and the polarization of the adsorbent surface caused the O=C=O bond to be bent and negatively charging, while the electrophilic attack by the dehydroxylation from COOH* and H2COOH* led to the formation of CO and CH4. The enhancement of polarization results from the field-field coupling showed that plasmonic metals in the photothermal catalytic reduction of CO2 to carbon fuel also have a good application. Table 3 summarizes plasmonic-metal photothermal catalysts and products.

4.3. Perovskite Photothermal Catalyst

Perovskite oxides (ABO3) are widely studied for various photocatalytic applications owing to their high stability, structural tunability and excellent photocatalytic activity [138,139,140]. The composition of elements at A and B sites will directly affect band structure and photophysical properties for materials [141]. Figure 8 shows the corresponding structures and compositions of these different classes of perovskites, as well as the various elements commonly used to date to construct stable perovskite-type catalysts. The most notable advantage of perovskite-type catalysts in photocatalysis is their flexibility in A and B-site elements’ composition and structure. The different perovskite-type catalysts exhibit significant distinction in reduction activity and product selectivity for CO2, which has been demonstrated in many studies [142,143,144]. In our previous study, double perovskites LaSrCoFeO6−δ (LSCF) and LaSrCoFeO6−δ (3DOM-LSCF) catalysts with three-dimensionally ordered macro-porous structure showed good catalytic activity and product selectivity for CO2 and H2O into CH4 under the photothermal coupling condition. 3DOM-LSCF catalyst showed the production of CH4 reached (557.88 μmol·g−1), which was 1.6 times higher than LSCF (351.32 μmol·g−1) at the same condition. The excellent catalytic performance of 3DOM-LSCF because of 3DOM structure with highly porous structure, high surface areas, small crystallite size, mixed ions, mixed-valence, self-formed heterostructures and self-formed oxygen vacancies can improve CO2 and H2O absorption and reaction on the catalyst surface [36]. The other research on LaCoxFe1−xO3 [38] and LaNixFe1−xO3 [40] perovskite catalysts was also investigated. The results showed that catalytic activity varied with the amount of Co and Ni-doping. When Co = 0.6 and Ni = 0.6, the CH4 and CH3OH reached 437.28 μmol·g−1, 13.75 μmol·g−1 and 471.39·μmol·g−1, 15.50 μmol·g−1 under photothermal coupling with CO2 and H2O, which were 3.2, 4.0 and 3.5, 4.1 times than pure LaFeO3, respectively. It was found that doping-Co and Ni in LaFeO3 can effectively change the positions of VB and CB of semiconductor materials, the gap width decreases and the valence band and conduction band are more prone to oxidizing reaction, promoting the generation of CH4 and CH3OH (0 ≤ x ≤1). Except for the perovskite oxide photocatalysts, halide perovskites (ABX3, X = Cl, Br and I) are of interest for photocatalysis, especially for CO2 reduction into hydrocarbons, which have been studied extensively. As with perovskite oxides, halide perovskite has a strong light absorption capacity and its band gap value, light absorption range, catalytic activity and product selectivity can be adjusted by changing the element composition and structure of halide [145,146]. The toxicity issue, poor stability and inadequate active sites of pristine metal halide perovskite limit the application in photocatalytic reactions, which raise an intense recent research effort to develop various engineering strategies for improving their performance (e.g., selectivity, activity, stability, recyclability and environmental compatibility) [147]. Fe(II)-doped CsPbBr3 perovskite nanocrystals have shown enhanced catalytic activity, and also predominantly leading to the evolution of CH4 instead of CO. The reason for this finding was that Fe(II)-doped CsPbBr3 nanocrystals, the adsorption energy for CH4, became more positive, indicating active CH4 molecules once formed were faster desorbed from the catalyst surfaces and as a better photoconductor to show a superior photo-response compared to undoped CsPbBr3 nanocrystals [148]. In general, whether it be perovskite oxide or halide perovskite, improving the activity of photothermal catalytic reduction of CO2 through structural modification is worth further studying. Table 4 summarizes perovskite photothermal catalysts and products.

4.4. Other Photothermal Catalysts

Apart from these three important photothermal catalysts, other photothermal catalysts with strong light and thermal absorption capacity, high catalytic activity and product selectivities, such as g-C3N4, graphene and carbon nanotubes/nanowires also show the potential for photothermal catalytic reduction CO2 into fuel [160,161,162,163]. Cui and Liu et al. [99] reported that photothermal catalytic reduction of CO2 to produce CH3CH2OH using Au-Cu alloy nanoparticles loaded with ultra-thin porous g-C3N4 nanosheets. The positive charge on the surface of Au is conducive to enhancing the adsorption of CO2 molecules on its surface, while the negative charge transferred from the surface of Au to Cu in the alloy, which makes the Cu surface enriched with a negative charge and promotes intermediate species (CO2 and *CO) formation on the surface. The close connection and strong interaction between g-C3N4 and alloy are favorable for the conduction of photogenic carriers. In the process of reaction, increasing the reaction temperature can accelerate the thermal movement of molecules and promote the synergistic effect of photocatalysis and thermocatalysis, which was favorable for the C–C coupling of *CO polymerization and increased the generation of ethanol. The ethanol yield of Au-Cu (1.0%)/g-C3N4 catalyst reached 0.89 mmol·g−1·h−1 under the condition of heating to 120 °C, which was 4.2 and 7.6 times higher than single photocatalysis and thermocatalysis, respectively. TiO2-G (30g) composite showed CH4 and CO production rates of 26.7 μmol·g−1·h−1 and 5.2 μmol·g−1·h−1, which were 5.1 and 2.8 times higher than pure TiO2. The reason was that graphene acted as storage in TiO2-based photocatalyst that can efficiently collect and accept photogenerated electrons from TiO2, which can sharply inhibit the recombination and enhance the separation efficiency of carriers. Upon light illumination, graphene can transform the absorbed light energy into heat energy to induce a local photothermal effect. The increase in graphene surface temperature is beneficial to improving the conversion of CO2 and also establishes a balance with TiO2 light absorption [112].

5. Challenges and Strategies to Boost CO2 Photothermal Conversion

5.1. Enlarge and Enhance Light Absorption and Photothermal Conversion

The wavelength range of sunlight is mainly between 250 and 2500 nm, in which ultraviolet only accounts for 5%, visible light accounts for 43% and near-infrared light accounts for 52% of the entire solar spectrum [164]. The first step for any photothermal catalysis conversion process is light absorption. Most of the traditional semiconductor catalysis has a wide bandgap, such as TiO2, which can only absorb ultraviolet light to generate electron-hole pairs to achieve catalytic ability [92,165]. Therefore, broadening the spectral response range aims to improve the utilization rate of solar energy and it is one of the main ways to enhance the catalytic performance of semiconductors. In general, regulating the band structure of semiconductors is a common method to broaden the spectral response range. It is well known that surface defects (such as the existence of oxygen vacancies) can effectively enhance light absorption and the LSPR effect leads to plasmonic excitations being observed across the visible, near-infrared and mid-infrared regions of the electromagnetic spectrum that can enhance light absorption [166,167]. Oxygen vacancies cannot only serve as electron capture centers but also enhance the ability of light capture to improve the catalytic reaction activity. A catalyst with oxygen vacancy confined in PO4 oxoanion-doped Bi2WO6 atom layers (Vo-PO4-BWO) has been proposed. Combined with theoretical and experimental results, it can be concluded that the light absorption of the Vo-PO4-BWO atom layer obviously shifted to visible light and effectively narrowed the band gap. The conversion of CO2 to CH3OH reached 157 μmol·g−1 h−1. Doping of PO4 oxoanion promoted the increase of carriers, and the existence of oxygen vacancy caused the emergence of a new defect level, which enabled electrons to be easily excited by sunlight and quickly transferred to the surface of the catalyst. This work was an effective approach for efficient solar energy conversion using spatially positioned material designs to expand catalyst light absorption and enhance charge transport at the atomic level and demonstrates the importance of well-positioned material designs for CO2 reduction [168]. Loading Cu, Ag, Au, Pt and doping non-metallic elements such as S and N seem to be promising techniques to effectively enhance the light response range of semiconductor catalytic [169,170]. Khalid et al. [171] achieved metal ions (Ag and Cu) loaded and N doped TiO2 nanometer photocatalyst. In this study, the light absorption range and catalytic activity of TiO2, N/TiO2, Ag-N/TiO2 and Cu-N/TiO2 were studied. The results showed that N-doping would generate an intermediate energy level between the bandgap of TiO2, Ag and Cu loading would promote the interaction between metal and non-metal, leading to a significant decrease in the energy band, thus improving the visible light absorption of the catalyst. In terms of catalytic activity, Ag-N/TiO2 had the highest catalytic activity, because Ag loading not only enhances the absorption of visible light of the catalyst, but also acts as the separation center of electron trap and electron-hole pair to improve the activity of photocatalytic reduction of CO2 and H2O.

5.2. Improving Carrier Separation and Migration Efficiency

In the process of photothermal catalytic CO2 reduction, photogenerated electron-hole pair separation and migration are the rate-limiting step in the whole reaction, which determines photothermal catalytic CO2 reduction activity. The rapid and effective separation of electron-hole pairs can generate more active electrons CO2 reduction and the rapid migration can avoid the recombination of photogenerated electrons and holes, thus the number of effective electrons increases and the catalytic activity improves. To facilitate the separation and migration of photogenerated electrons and holes, the modification of catalysts has been extensively studied, such as combining with wide bandgap and narrow bandgap semiconductors, noble metal loading to build heterojunction, doping to change the bandgap value and introducing level defects in the bulk of the catalyst or near the CB, forming shell structure and so on.

5.2.1. Building Heterojunction

To enhance the low activity of semiconductor catalysts, many strategies have been studied. A well-defined heterojunction structure can accelerate the separation of photogenerated e/h+ pairs and suppress the recombination of e/h+ pairs, thereby considerably promoting the activity during the photocatalytic reduction of CO2 into solar fuels. There are three types of heterojunctions: p-n heterojunctions [172], no heterojunction and Z-scheme heterojunction [173]. A p-n heterojunction can be formed when n-type semiconductors couple with an appropriate p-type semiconductor with matching electronic band structures; non-p-n heterojunction semiconductor A (SA) and semiconductor B (SB) are tightly bonded to form a heterojunction, e can transfer from the CB of SA to the CB of SB, whereas h+ transfers from the VB of SB to the VB of SA under light irradiation, because of the staggered alignment of the energy levels; the Z-scheme heterojunction system is typically made up of two semiconductors with the staggered alignment of band structures, but the reduction and oxidation reaction occur on CB and VB of the two semiconductors and the whole reaction cannot be maintained independently [174]. Co3O4@CdIn2S4 p-n heterojunction has been shown to exhibit a high CO2 convert ratio with visible light. In this study, the high CO2 reduction performance of Co3O4@CdIn2S4 p-n heterojunction photocatalyst was due to the strong coupling interface formed by the composition of hybrid materials. There was an internal electric field between the interfaces. When visible light is irradiated, the electron-hole pairs generated by both Co3O4 and CdIn2S4 and rapidly transferred from CB of Co3O4 to CB of CdIn2S4, and the hole rapidly transferred from VB of CdIn2S4 to VB of Co3O4 under the action of the internal electric field, which can effectively avoid the recombination of carriers. In addition, the Co3O4@CdIn2S4 composite catalyst had more catalytic active sites to promote the adsorption of CO2. The main reason for promoting solar-driven CO2 is to generate carbon fuel. This study also provides enlightenment for the further rational design of heterojunction photocatalysts [175]. In another study, BiPO4-BiOBrxI1−x p-n heterojunction catalysts were successfully exploited for photocatalytic reduction of CO2 to CH4 and CO, because the formation of the p-n junction can reduce the recombination of photogenerated electrons and holes, improve the separation efficiency and enhance the absorption of visible light. Loading (5%) of BiPO4 in BiPO4-BiOBr0.75I0.25 showed the highest CO (24.9umol·g−1) and CH4 (9.4umol·g−1) yield rates, which showed the optimal amount loading result higher conversion efficiency. This report demonstrated that the construction and design of BiPO4-BiOBrxI1−x p-n heterojunction provides a novel approach for efficient catalytic reduction of CO2 [176]. TiO2/ SrTiO3 heterojunction structure films were fabricated and applied for CO2 photoreduction. TiO2/SrTiO3 reflected a good catalytic activity of CO2 to CH4. The reason is attributed to the low recombination of charge carriers, the fast separation of photogenerated e-h+ pairs, large BET surface, strong photo conversion ability and adsorption capacity for CO2 of TiO2/SrTiO3 heterojunction catalyst, which promote the efficiency of CO2 and H2O conversion into CH4 and CO [177]. A direct Z-scheme Sn-In2O3/In2S3 heterogeneous photocatalyst effectively enhanced photocatalytic CO2 reduction activity and expanded the range of light absorption. There are few defect traps in the heterojunction, which could promote the separation and transmission of e and h+ pairs without sacrificing the redox capacity of the carriers due to the good lattice match [178]. All these results indicate that construct heterojunction structure photocatalysts can rapidly inhibit the recombination and accelerate the separation and migration of carriers to strengthen the activity of CO2 photocatalytic reduction.

5.2.2. Doping

Regulating semiconductor bandgap is one of the effective methods to improve photothermal catalytic activity. Doping is the simplest and most effective way to change the band structure of semiconductor-based photothermal catalysts. After doping, impurity levels are formed in the energy bandgap. Generally, the introduction of impurity level can shorten the bandgap, inducing the change of CB position and light with lower energy can also excite electrons, which broaden the range of light response. Tuning the CB can obtain appropriate reduction potentials of electrons and control the reaction from a thermal dynamic point of view. Doping results in lattice defects, which are favorable for the formation of more active centers. Moreover, doped ions can be used as capture centers to separate electron-hole pairs. Wei et al. synthesized a series of LaNixCo1−xO3 perovskites. Under photothermal reaction conditions, Ni doping can effectively improve the catalytic performance of CO2 plus H2O reduction to CH4/CH3OH compared with LaCoO3. LaNi0.4Co0.6O3 showed the highest catalytic activity; the accumulated yield of CH4 and CH3OH reached 678.57 μmol·g−1 and 20.83 μmol·g−1, which were 3.4 and 3.8 times higher than LaCoO3 under the same condition. The reason was that the introduction of Ni can change the positions of the CB and VB, reduce the bandgap value and improve the absorption capacity of visible light. In addition, the introduced Ni also formed additional oxygen vacancies to promote the adsorption of reactants, leading to the distortion of CO2 and H2O molecules, accelerating the decomposition of molecules, and can capture electrons to reduce the recombination rate of carriers and promote the occurrence of carbon dioxide reduction reaction as the active site of the reaction [151]. In another research [39], proton conductor BaZr0.5Ce0.3Y0.2CoO3 (BZCY 532) as a new photothermal catalyst can convert CO2 into CH4 C2H6 and C3H8 under photothermal coupling. Introducing Ni and Co effectively improves the photocatalytic efficiency, especially for Co-doping, because doping elements can change the original ion space of the catalyst and the structure of the catalyst is deformed, which could promote the absorption of light, increase the surface oxygen vacancies, produce more reactive sites to promote the CO2 absorption and activation and stabilize the adsorption of intermediate ·CH3 to promote C–C and CCC coupling.

5.2.3. Core-Shell Structure

Core-shell structure has attracted the attention of researchers because of its unique structural characteristics, which integrate the properties for inner and outer materials and complement each other. Its high specific surface area and special shape are favorable for application in photothermal catalysis. The core usually exhibits different properties from the shell and can chemically contact the shell in three dimensions, leading to enhanced properties such as catalysis and interfacial charge transfer. Moreover, the shell with a large bandgap may improve the photostability of the core catalyst with a narrow bandgap [179]. Based on structure and morphology, core-shell structures were classified into the core-shell, yolk-shell/hollow structures and sandwiched core-shell structures, as shown in Figure 9. Au@TiO2-Au catalysts with core-shell morphology were found to be superior for CO2 conversion. The thickest TiO2 semiconductor structure is the reaction. Plasmonic field effects or a “hot electron transfer” from the gold core may affect the reactivity of the small gold nanoparticles on the outer TiO2 shell. The efficiency of CO2 converted into CH4 and CO has been shown to improve because of the synergistic effect between photogenerated electrons, hot electrons and gold core, titanium shell and Au NPs exposed on the outer three structure features [180]. In another study, Au@TiO2 yolk-shell hollow spheres were created for CO2 reduction under UV-Vis light irradiation, and were found to not only generate CH4 but also produce C2H6. Experimental and calculated analysis elucidated that Au NPS plays an important role in improving photocatalytic efficiency and promoting multi-electron reactions to achieve C–C coupling. The LSPR of Au NPs contributes to the generation of electron hole pairs on the TiO2 shell, and collective oscillation caused by LSPR can cause local electromagnetic field and photocurrent, which can greatly promote the generation and separation of electron-hole pairs and increase the number of carriers. The closer the TiO2 shell was to the local electromagnetic field of Au NPs, the more favorable it was for the accumulation of ⋅CH3 radicals and dimerization ⋅CH3 radicals to form C2H6. The farther from Au NPs, the more beneficial it was for the formation of CH4 [181].

5.3. Photothermal Reactor

To improve the activity of photothermal catalytic reduction of CO2 into valuable chemicals or fuels, one method, described above, is to modify catalysts based on different aspects to improve the reactivity; another way is to design a photothermal catalytic reactor to improve the photothermal catalytic conversion efficiency [182]. Up to now, there are many types of reactors designed for the application of photothermal catalytic CO2 reduction. A novel multi-channel mini-reactor with 1.17–2 mm channel internal diameter was reported (Figure 10a). Taking Ni on the alumina (25% Ni) catalyst as an example to study the difference in product yield and selectivity in catalytic CO2 hydrogenation between the multi-channel micro-reactor and the traditional fixed-bed reaction at different temperatures (Figure 10b,c). It was observed that the multi-channel mini-reactor had higher CO2 conversion and CH4 selectivity than the fixed-bed reactor at the same reaction condition. The reason is that the multi-channel microreactor has a very high specific surface, which not only is conducive to the uniform distribution of the catalyst in the reactor but also promote the reaction gas and catalyst contact to be more adequate to improve the thermal conversion efficiency and eliminate the formation of hot spots [183]. Another study reported a “HI-light” (glass waveguide-based ‘‘shell-and-tube’’ photoreactor platform) was applied to photothermal catalytic CO2 reduction (Figure 10d). The reactor design is characterized by flexibility in both diameter and length. The reaction activity of In2O3−x (OH) Y nanocrystals as a photothermal catalyst for CO2 hydrogenation to CO on the HI-light platform was studied from the reaction temperature, light intensity and residence time of reactants in the reactor. Through the experimental results, the photothermal coupling (UV light, 300 °C) catalytic CO2-CO production rate was 20 times higher than the cubic In2O3−x(OH)Y catalyst, which has reported the highest catalytic activity under similar conditions, and the CO2 conversion and CO production rate increased with the increase in reaction temperature under HI-light platform reactor [184]. This study further proved that reactor design can be a useful approach to improving the catalytic activity of catalysts.
The photothermal coupling catalytic reaction consists of photoexcitation and thermal promotion. There are various ways where the temperature can be manipulated inside the photothermal reactor, either by changing the distance of the lamp with the reactor, or thermal energy will generally be produced by the irradiation of light on the catalyst, which will convert solar energy to heat energy by absorbing the incident photons or by using an external heater. The first way is that light provides heat. The main factors affecting heat are the category of light, the light intensity and the height of the light source. In the research on the Co nanostructure catalyst, by adjusting the distance of a lamp from the reactor, the temperature elevated to 158 °C and found that Co catalysts are efficiently heated due to their full absorption of UV-Vis-NIR light [185]. Li studied the light-driven temperature and CO2 conversion rates of the Ni/Y2O3 nanosheets with (Ni/Y2O3 + S) and without (Ni/Y2O3) the selective light absorber-assisted photothermal system, respectively, under different intensities of sunlight irradiation [186]. The result showed that with the light intensity increased, both temperature and CO2 conversion rate increased (Figure 11a,b). Time-dependent surface temperature profiles of different catalyst films upon laser illumination were researched. The results showed that the surface temperature of the catalyst increased with the irradiation time, and CO selectivity increased with reaction temperature [24]. Figure 11c shows the surface temperature and product yield of the catalyst under different light sources. It is clear that the surface temperature of the catalyst is the highest under VU-VIS-IR light and the production of the products over WO3−x is higher than under UV-Vis and IR irradiation. This suggested that the catalytic activity is not only related to the reaction temperature but also indicates that different light sources will produce different thermal effects in photothermal catalysis [86].
The other study found that catalysts can absorb light and convert solar energy into heat energy, the increase in local temperature can help the whole reaction occur more easily and promote CO2 conversion rapidly. Generally, the main reason for the temperature rise is the LSPR effects of plasmonic metals or photon vibrations of non-plasmonic metals, which were introduced in the former. The last way is to configure heating devices outside the reactor when the catalyst is unable to efficiently generate heat by absorbing solar energy or the heat generated is not high enough to drive the chemical reaction. External heating devices are used to ensure that the reactor can be maintained at a fixed temperature to provide steady heat energy for the progression of reaction over the catalyst. Usually, the external heating device is a heating plate, electrical furnace, heating sleeve or heating belt [40,187,188].

6. Conclusions and Outlook

In conclusion, photothermal catalysis technology is a combination of photocatalysis and thermocatalysis, which is an improvement on traditional catalytic technology. Carbon dioxide is an abundant source of C that can be converted into high-value-added chemicals or fuels through chemical methods. Low activity and selectivity are the main factors that limit wide use of single catalytic technology. The photothermal catalysis of CO2 reduction can effectively promote the synergistic effect of light energy and heat energy, which can make the system have higher catalytic activity and selectivity to target products even under mild conditions. Until now, the photothermal catalytic CO2 reduction has made great progress, both in the selection of photothermal catalysts and in the study of the photothermal catalytic reaction mechanisms. In this paper, different types of photothermal catalysts, photothermal catalytic products and ways to improve the catalytic activity of photothermal catalysts are described. While photothermal catalysis opens new avenues for the widespread use of renewable energy sources such as solar energy to reduce greenhouse gases in the atmosphere while generating useful green energy, fuels and chemicals, it is of inestimable value in both the energy and environmental fields. The reduction and effective utilization of carbon dioxide, regardless of whether in the current state or long-term development and the realization of the carbon cycle of CO2 through photothermal coupling is undoubtedly the most promising research and development technology. However, much of the research and exploration are still in the most basic state and cannot be industrialized in practical application. Therefore, further research in this field still faces many challenges. Some typical issues are listed below:
(1)
Catalysts: Although various types of photothermal catalysts have been studied so far, poor light absorption conversion ability, low activity and low stability are still the biggest problems. Photothermal catalysts with high catalytic activity supported by noble metals have been extensively studied, but their wide application is limited by the high cost. It still recommends using low-cost and highly active materials. At the same time, the stability of photothermal catalysts also needs to be constantly improved, if a true sense of industrialization is to be achieved, low-cost, effective light absorption, high activity, high selectivity, and strong catalyst stability are essential.
(2)
Reactants: There are several applications for photothermal catalytic reduction of CO2 into hydrocarbon. Reducing agents are mainly H2, CH4 and H2O, which can also be divided into CO2 dry reforming, CO2 hydrogenation reaction and artificial photosynthesis. In hydrogenation and dry reforming, the advantage of reducing agents is that target products such as carbon CO, CH4 and CH3OH are easy to obtain. However, the reducing agent is water in artificial light synthesis. In the process of photothermal catalytic CO2 reduction, at first, H2O needs to be hydrolyzed in the valence band to provide a hydrogen source and then CO2 reduction reaction occurs in the conduction band, so the reaction difficulty is increased. H2O is inexhaustible clean energy and a good candidate for photocatalytic or photothermal catalytic CO2 reduction. If the reaction activity is greatly improved in the photothermal catalytic reduction of H2O with CO2, it will accelerate the realization of clean energy in the real sense.
(3)
Products: At present, photothermal catalytic reduction of CO2 has a high selectivity for low value-added single-carbon products, such as CO or CH4, but high value-added multi-carbon products are more valuable, such as alcohols or C2+ hydrocarbons. If more value-added products can be obtained by catalyst design or adjusting reaction selectivity, then carbon cycling using CO2 will be more valuable.
(4)
Mechanism: There are few studies on the mechanism for the photothermal catalytic reduction of CO2. In the intermediate process of reaction, the process and intermediate that is promoted by light and heat and the roles played in each intermediate process are all worthy of further studying.
(5)
Reactor: By considering the structure of the reactor, the materials, temperature resistance, strength, volume, type of light source, external heating configuration and the operation mode of the reaction, the reactor is designed to be easy to operate and effectively promote the reaction, which can effectively improve the efficiency of photothermal catalytic reduction of CO2 to hydrocarbons.
Photothermal catalysis opens up a new way for the widespread use of renewable energy sources such as solar energy to reduce greenhouse gases in the atmosphere while generating useful green energy, fuels and chemicals, which have inestimable value in both the energy and environmental fields. The reduction and effective utilization of CO2, regardless of whether in the current state or long-term development and the realization of the carbon cycle of CO2 through photothermal coupling is undoubtedly the most promising research and development technology.

Author Contributions

J.T.: framework, drafting and writing of the review, graphical representation; R.H.: literature review and summary, Q.G., Z.Z. and N.S.: drafting, providing input for refining the paper and English editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Key R&D Program of China (Nos. 2017YF. B1103500 and 2017YFB1103502) and Shanghai Technical Service Center for Advanced Ceramics Structure Design and Precision Manufacturing (NO. 20DZ2294000).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Calderone, L. What is the Future of Non-renewable Resources. AltEnergyMag, 23 July 2019. [Google Scholar]
  2. Kuo, G. Carbon Dioxide Emissions and Carbon Footprint. MAHB, 24 July 2018. [Google Scholar]
  3. Solomon, S. Climate Change 2007-The Physical Science Basis: Working Group I Contribution to the Fourth Assessment Report of the IPCC; Cambridge University Press: Cambridge, UK, 2007; Volume 4. [Google Scholar]
  4. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon Capture and Storage (CCS): The Way Forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef] [Green Version]
  5. Bagger, A.; Christensen, O.; Ivaništšev, V.; Rossmeisl, J. Catalytic CO2/CO Reduction: Gas, Aqueous, and Aprotic Phases. ACS Catal. 2022, 12, 2561–2568. [Google Scholar] [CrossRef]
  6. Chai, X.; Huang, H.-H.; Liu, H.; Ke, Z.; Yong, W.-W.; Zhang, M.-T.; Cheng, Y.-S.; Wei, X.-W.; Zhang, L.; Yuan, G. Highly Efficient and Selective Photocatalytic CO2 to CO Conversion in Aqueous Solution. Chem. Commun. 2020, 56, 3851–3854. [Google Scholar] [CrossRef] [PubMed]
  7. Tahir, M.; Amin, N.S. Indium-Doped TiO2 Nanoparticles for Photocatalytic CO2 Reduction with H2O Vapors to CH4. Appl. Catal. B Environ. 2015, 162, 98–109. [Google Scholar] [CrossRef]
  8. Phongamwong, T.; Chareonpanich, M.; Limtrakul, J. Role of Chlorophyll in Spirulina on Photocatalytic Activity of CO2 Reduction under Visible Light over Modified N-Doped TiO2 Photocatalysts. Appl. Catal. B Environ. 2015, 168, 114–124. [Google Scholar] [CrossRef]
  9. Hamdy, M.S.; Amrollahi, R.; Sinev, I.; Mei, B.; Mul, G. Strategies to Design Efficient Silica-Supported Photocatalysts for Reduction of CO2. J. Am. Chem. Soc. 2014, 136, 594–597. [Google Scholar] [CrossRef]
  10. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  11. Yin, Z.; Palmore, G.T.R.; Sun, S. Electrochemical Reduction of CO2 Catalyzed by Metal Nanocatalysts. Trends Chem. 2019, 1, 739–750. [Google Scholar] [CrossRef]
  12. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective Electrochemical Reduction of Carbon Dioxide to Ethylene and Ethanol on Copper(I) Oxide Catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  13. He, M.; Li, C.; Zhang, H.; Chang, X.; Chen, J.G.; Goddard, W.A.; Cheng, M.; Xu, B.; Lu, Q. Oxygen Induced Promotion of Electrochemical Reduction of CO2 via Co-Electrolysis. Nat. Commun. 2020, 11, 3844. [Google Scholar] [CrossRef]
  14. Kumar, B.; Llorente, M.; Froehlich, J.; Dang, T.; Sathrum, A.; Kubiak, C.P. Photochemical and Photoelectrochemical Reduction of CO2. Annu. Rev. Phys. Chem. 2012, 63, 541–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Guo, L.; Cao, J.; Zhang, J.; Hao, Y.; Bi, K. Photoelectrochemical CO2 Reduction by Cu2O/Cu2S Hybrid Catalyst Immobilized in TiO2 Nanocavity Arrays. J. Mater. Sci. 2019, 54, 10379–10388. [Google Scholar] [CrossRef]
  16. Ikeda, S.; Tanaka, Y.; Kawaguchi, T.; Fujikawa, S.; Harada, T.; Takayama, T.; Iwase, A.; Kudo, A. Photoelectrochemical Reduction of CO2 to CO Using a CuGaS2 Thin-Film Photocathode Prepared by a Spray Pyrolysis Method. Chem. Lett. 2018, 47, 1424–1427. [Google Scholar] [CrossRef] [Green Version]
  17. Pawar, A.U.; Kim, C.W.; Nguyen-Le, M.-T.; Kang, Y.S. General Review on the Components and Parameters of Photoelectrochemical System for CO2 Reduction with in Situ Analysis. ACS Sustain. Chem. Eng. 2019, 7, 7431–7455. [Google Scholar] [CrossRef]
  18. Welte, M.; Barhoumi, R.; Zbinden, A.; Scheffe, J.R.; Steinfeld, A. Experimental Demonstration of the Thermochemical Reduction of Ceria in a Solar Aerosol Reactor. Ind. Eng. Chem. Res. 2016, 55, 10618–10625. [Google Scholar] [CrossRef] [PubMed]
  19. Wu, X.-Y.; Ghoniem, A.F. Hydrogen-Assisted Carbon Dioxide Thermochemical Reduction on La0.9Ca0.1FeO3−δ Membranes: A Kinetics Study. ChemSusChem 2018, 11, 483–493. [Google Scholar] [CrossRef]
  20. Guene Lougou, B.; Shuai, Y.; Zhang, H.; Ahouannou, C.; Zhao, J.; Kounouhewa, B.B.; Tan, H. Thermochemical CO2 Reduction over NiFe2O4@alumina Filled Reactor Heated by High-Flux Solar Simulator. Energy 2020, 197, 117267. [Google Scholar] [CrossRef]
  21. Hare, B.J.; Maiti, D.; Ramani, S.; Ramos, A.E.; Bhethanabotla, V.R.; Kuhn, J.N. Thermochemical Conversion of Carbon Dioxide by Reverse Water-Gas Shift Chemical Looping Using Supported Perovskite Oxides. Catal. Today 2019, 323, 225–232. [Google Scholar] [CrossRef]
  22. Wang, L.; Liu, X.; Dang, Y.; Xie, H.; Zhao, Q.; Ye, L. Enhanced Solar Induced Photothermal Synergistic Catalytic CO2 Conversion by Photothermal Material Decorated TiO2. Solid State Sci. 2019, 89, 67–73. [Google Scholar] [CrossRef]
  23. Zhao, Z.; Doronkin, D.E.; Ye, Y.; Grunwaldt, J.-D.; Huang, Z.; Zhou, Y. Visible Light-Enhanced Photothermal CO2 Hydrogenation over Pt/Al2O3 Catalyst. Chin. J. Catal. 2020, 41, 286–293. [Google Scholar] [CrossRef] [Green Version]
  24. Cai, M.-J.; Li, C.-R.; He, L. Enhancing Photothermal CO2 Catalysis by Thermal Insulating Substrates. Rare Met. 2020, 39, 881–886. [Google Scholar] [CrossRef]
  25. Xu, C.; Huang, W.; Li, Z.; Deng, B.; Zhang, Y.; Ni, M.; Cen, K. Photothermal Coupling Factor Achieving CO2 Reduction Based on Palladium-Nanoparticle-Loaded TiO2. ACS Catal. 2018, 8, 6582–6593. [Google Scholar] [CrossRef]
  26. Li, D.; Huang, Y.; Li, S.; Wang, C.; Li, Y.; Zhang, X.; Liu, Y. Thermal Coupled Photoconductivity as a Tool to Understand the Photothermal Catalytic Reduction of CO2. Chin. J. Catal. 2020, 41, 154–160. [Google Scholar] [CrossRef]
  27. Xie, C.; Chen, C.; Yu, Y.; Su, J.; Li, Y.; Somorjai, G.A.; Yang, P. Tandem Catalysis for CO2 Hydrogenation to C2–C4 Hydrocarbons. Nano Lett. 2017, 17, 3798–3802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Urbain, F.; Smirnov, V.; Becker, J.-P.; Lambertz, A.; Yang, F.; Ziegler, J.; Kaiser, B.; Jaegermann, W.; Rau, U.; Finger, F. Multijunction Si Photocathodes with Tunable Photovoltages from 2.0 V to 2.8 V for Light Induced Water Splitting. Energy Environ. Sci. 2016, 9, 145–154. [Google Scholar] [CrossRef] [Green Version]
  29. Kim, W.; Seok, T.; Choi, W. Nafion Layer-Enhanced Photosynthetic Conversion of CO2 into Hydrocarbons on TiO2 Nanoparticles. Energy Environ. Sci. 2012, 5, 6066–6070. [Google Scholar] [CrossRef] [Green Version]
  30. Li, K.; An, X.; Park, K.H.; Khraisheh, M.; Tang, J. A Critical Review of CO2 Photoconversion: Catalysts and Reactors. Catal. Today 2014, 224, 3–12. [Google Scholar] [CrossRef] [Green Version]
  31. Osterloh, F.E. Photocatalysis versus Photosynthesis: A Sensitivity Analysis of Devices for Solar Energy Conversion and Chemical Transformations. ACS Energy Lett. 2017, 2, 445–453. [Google Scholar] [CrossRef]
  32. He, M.; Sun, Y.; Han, B. Green Carbon Science: Scientific Basis for Integrating Carbon Resource Processing, Utilization, and Recycling. Angew. Chem. Int. Ed. 2013, 52, 9620–9633. [Google Scholar] [CrossRef]
  33. Wang, Z.; Song, H.; Liu, H.; Ye, J. Coupling of Solar Energy and Thermal Energy for Carbon Dioxide Reduction: Status and Prospects. Angew. Chem. Int. Ed. 2020, 59, 8016–8035. [Google Scholar] [CrossRef]
  34. Jia, J.; Wang, H.; Lu, Z.; O’Brien, P.G.; Ghoussoub, M.; Duchesne, P.; Zheng, Z.; Li, P.; Qiao, Q.; Wang, L.; et al. Photothermal Catalyst Engineering: Hydrogenation of Gaseous CO2 with High Activity and Tailored Selectivity. Adv. Sci. 2017, 4, 1700252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ghoussoub, M.; Xia, M.; Duchesne, P.N.; Segal, D.; Ozin, G. Principles of Photothermal Gas-Phase Heterogeneous CO2 Catalysis. Energy Environ. Sci. 2019, 12, 1122–1142. [Google Scholar] [CrossRef]
  36. Ha, M.N.; Lu, G.; Liu, Z.; Wang, L.; Zhao, Z. 3DOM-LaSrCoFeO6−δ as a Highly Active Catalyst for the Thermal and Photothermal Reduction of CO2 with H2O to CH4. J. Mater. Chem. A 2016, 4, 13155–13165. [Google Scholar] [CrossRef]
  37. Wang, L.; Wang, Y.; Cheng, Y.; Liu, Z.; Guo, Q.; Ha, M.N.; Zhao, Z. Hydrogen-Treated Mesoporous WO3 as a Reducing Agent of CO2 to Fuels (CH4 and CH3OH) with Enhanced Photothermal Catalytic Performance. J. Mater. Chem. A 2016, 4, 5314–5322. [Google Scholar] [CrossRef]
  38. Xu, L.; Ha, M.N.; Guo, Q.; Wang, L.; Ren, Y.; Sha, N.; Zhao, Z. Photothermal Catalytic Activity of Combustion Synthesized LaCoxFe1−xO3 (0 ≤ x ≤ 1) Perovskite for CO2 Reduction with H2O to CH4 and CH3OH. RSC Adv. 2017, 7, 45949–45959. [Google Scholar] [CrossRef] [Green Version]
  39. Tian, J.; Ren, Y.; Liu, L.; Guo, Q.; Sha, N.; Zhao, Z. Photothermal Catalysis for CO2 Convert into C1–C3 Hydrocarbons by Proton Conductor BZCY532. Mater. Res. Express 2020, 7, 085504. [Google Scholar] [CrossRef]
  40. Zheng, D.; Wei, G.; Xu, L.; Guo, Q.; Hu, J.; Sha, N.; Zhao, Z. LaNixFe1−xO3 (0 ≤ x ≤ 1) as Photothermal Catalysts for Hydrocarbon Fuels Production from CO2 and H2O. J. Photochem. Photobiol. Chem. 2019, 377, 182–189. [Google Scholar] [CrossRef]
  41. Zhang, F.; Li, Y.-H.; Qi, M.-Y.; Yamada, Y.M.A.; Anpo, M.; Tang, Z.-R.; Xu, Y.-J. Photothermal Catalytic CO2 Reduction over Nanomaterials. Chem Catal. 2021, 1, 272–297. [Google Scholar] [CrossRef]
  42. Chang, X.; Wang, T.; Gong, J. CO2 Photo-Reduction: Insights into CO2 Activation and Reaction on Surfaces of Photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  43. Du, C.; Wang, X.; Chen, W.; Feng, S.; Wen, J.; Wu, Y.A. CO2 Transformation to Multicarbon Products by Photocatalysis and Electrocatalysis. Mater. Today Adv. 2020, 6, 100071. [Google Scholar] [CrossRef]
  44. Tang, S.; Sun, J.; Hong, H.; Liu, Q. Solar Fuel from Photo-Thermal Catalytic Reactions with Spectrum-Selectivity: A Review. Front. Energy 2017, 11, 437–451. [Google Scholar] [CrossRef]
  45. Fu, J.; Jiang, K.; Qiu, X.; Yu, J.; Liu, M. Product Selectivity of Photocatalytic CO2 Reduction Reactions. Mater. Today 2020, 32, 222–243. [Google Scholar] [CrossRef]
  46. Das, S.; Pérez-Ramírez, J.; Gong, J.; Dewangan, N.; Hidajat, K.; Gates, B.C.; Kawi, S. Core–Shell Structured Catalysts for Thermocatalytic, Photocatalytic, and Electrocatalytic Conversion of CO2. Chem. Soc. Rev. 2020, 49, 2937–3004. [Google Scholar] [CrossRef] [PubMed]
  47. Ma, R.; Sun, J.; Li, D.H.; Wei, J.J. Review of Synergistic Photo-Thermo-Catalysis: Mechanisms, Materials and Applications. Int. J. Hydrogen Energy 2020, 45, 30288–30324. [Google Scholar] [CrossRef]
  48. Kho, E.T. A Review on Photo-Thermal Catalytic Conversion of Carbon Dioxide. Green Energy 2017, 2, 204–217. [Google Scholar] [CrossRef]
  49. Han, X.; Li, M.; Ma, Y.; Li, Y.; Ma, H.; Wang, C. Thermal Coupled Photocatalysis to Enhance CO2 Reduction Activities on Ag Loaded G-C3N4 Catalysts. Surf. Interfaces 2021, 23, 101006. [Google Scholar] [CrossRef]
  50. Zhu, J.; Shao, W.; Li, X.; Jiao, X.; Zhu, J.; Sun, Y.; Xie, Y. Asymmetric Triple-Atom Sites Confined in Ternary Oxide Enabling Selective CO2 Photothermal Reduction to Acetate. J. Am. Chem. Soc. 2021, 143, 18233–18241. [Google Scholar] [CrossRef]
  51. Petryayeva, E.; Krull, U.J. Localized Surface Plasmon Resonance: Nanostructures, Bioassays and Biosensing—A Review. Anal. Chim. Acta 2011, 706, 8–24. [Google Scholar] [CrossRef]
  52. Lu, C.; Li, X.; Wu, Q.; Li, J.; Wen, L.; Dai, Y.; Huang, B.; Li, B.; Lou, Z. Constructing Surface Plasmon Resonance on Bi2WO6 to Boost High-Selective CO2 Reduction for Methane. ACS Nano 2021, 15, 3529–3539. [Google Scholar] [CrossRef]
  53. Jian, Z.; Hong, L.; Wang, H. Photothermal Catalysis for CO2 Conversion. Chin. Chem. Lett. 2022, in press. [Google Scholar]
  54. Zou, J.; Si, Z.; Cao, Y.; Ran, R.; Wu, X.; Weng, D. Localized Surface Plasmon Resonance Assisted Photothermal Catalysis of CO and Toluene Oxidation over Pd–CeO2 Catalyst under Visible Light Irradiation. J. Phys. Chem. C 2016, 120, 29116–29125. [Google Scholar] [CrossRef]
  55. Song, C.; Wang, Z.; Yin, Z.; Xiao, D.; Ma, D. Principles and Applications of Photothermal Catalysis. Chem. Catal. 2022, 2, 52–83. [Google Scholar] [CrossRef]
  56. Liu, H.; Dao, T.D.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Light Assisted CO2 Reduction with Methane over Group VIII Metals: Universality of Metal Localized Surface Plasmon Resonance in Reactant Activation. Appl. Catal. B Environ. 2017, 209, 183–189. [Google Scholar] [CrossRef]
  57. Ning, S.; Xu, H.; Qi, Y.; Song, L.; Zhang, Q.; Ouyang, S.; Ye, J. Microstructure Induced Thermodynamic and Kinetic Modulation to Enhance CO2 Photothermal Reduction: A Case of Atomic-Scale Dispersed Co–N Species Anchored Co@C Hybrid. ACS Catal. 2020, 10, 4726–4736. [Google Scholar] [CrossRef]
  58. Wu, K.; Chen, J.; McBride, J.R.; Lian, T. Efficient Hot-Electron Transfer by a Plasmon-Induced Interfacial Charge-Transfer Transition. Science 2015, 349, 632–635. [Google Scholar] [CrossRef] [Green Version]
  59. Li, N.; Jiang, R.; Li, Y.; Zhou, J.; Ma, Q.; Shen, S.; Liu, M. Plasma-Assisted Photocatalysis of CH4 and CO2 into Ethylene. ACS Sustain. Chem. Eng. 2019, 7, 11455–11463. [Google Scholar] [CrossRef]
  60. Bai, Y.; Zhao, J.; Feng, S.; Liang, X.; Wang, C. Light-Driven Thermocatalytic CO2 Reduction over Surface-Passivated β-Mo2C Nanowires: Enhanced Catalytic Stability by Light. Chem. Commun. 2019, 55, 4651–4654. [Google Scholar] [CrossRef]
  61. Bian, H.; Li, D.; Wang, S.; Yan, J.; Liu, S. 2D-C3N4 Encapsulated Perovskite Nanocrystals for Efficient Photo-Assisted Thermocatalytic CO2 Reduction. Chem. Sci. 2022, 13, 1335–1341. [Google Scholar] [CrossRef]
  62. Sun, M.; Zhao, B.; Chen, F.; Liu, C.; Lu, S.; Yu, Y.; Zhang, B. Thermally-Assisted Photocatalytic CO2 Reduction to Fuels. Chem. Eng. J. 2021, 408, 127280. [Google Scholar] [CrossRef]
  63. Yu, F.; Wang, C.; Ma, H.; Song, M.; Li, D.; Li, Y.; Li, S.; Zhang, X.; Liu, Y. Revisiting Pt/TiO2 Photocatalysts for Thermally Assisted Photocatalytic Reduction of CO2. Nanoscale 2020, 12, 7000–7010. [Google Scholar] [CrossRef]
  64. Han, R.; Chen, L.; Xing, B.; Guo, Q.; Tian, J.; Sha, N.; Zhao, Z. Pr3+-Doped La1−xPrxMn0.6Ni0.4O3−δ as Efficient Artificial Photosynthesis Catalysts for Solar Methanol. Catal. Commun. 2022, 165, 106440. [Google Scholar] [CrossRef]
  65. Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-the-Art Accomplishment, Challenges, and Prospects. Adv. Mater. 2014, 26, 4607–4626. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, L.; Liang, D.; Wang, Y.; Li, D.; Zhang, J.; Wu, L.; Feng, M.; Yi, F.; Xu, L.; Lei, L.; et al. Caged Circular SiRNAs for Photomodulation of Gene Expression in Cells and Mice. Chem. Sci. 2018, 9, 44–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Pang, R.; Teramura, K.; Morishita, M.; Asakura, H.; Hosokawa, S.; Tanaka, T. Enhanced CO Evolution for Photocatalytic Conversion of CO2 by H2O over Ca Modified Ga2O3. Commun. Chem. 2020, 3, 137. [Google Scholar] [CrossRef]
  68. Bai, Y.; Yang, P.; Wang, L.; Yang, B.; Xie, H.; Zhou, Y.; Ye, L. Ultrathin Bi4O5Br2 Nanosheets for Selective Photocatalytic CO2 Conversion into CO. Chem. Eng. J. 2019, 360, 473–482. [Google Scholar] [CrossRef]
  69. Pang, R.; Teramura, K.; Asakura, H.; Hosokawa, S.; Tanaka, T. Role of Bicarbonate Ions in Aqueous Solution as a Carbon Source for Photocatalytic Conversion of CO 2 into CO. ACS Appl. Energy Mater. 2019, 2, 5397–5405. [Google Scholar] [CrossRef]
  70. Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A. Fundamentals and Applications of Photocatalytic CO2 Methanation. Nat. Commun. 2019, 10, 3169. [Google Scholar] [CrossRef] [Green Version]
  71. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  72. Subrahmanyam, M. A Screening for the Photo Reduction of Carbon Dioxide Supported on Metal Oxide Catalysts for C1–C3 Selectivity. Appl. Catal. B Environ. 1999, 23, 169–174. [Google Scholar] [CrossRef]
  73. Tan, S.S.; Zou, L.; Hu, E. Kinetic Modelling for Photosynthesis of Hydrogen and Methane through Catalytic Reduction of Carbon Dioxide with Water Vapour. Catal. Today 2008, 131, 125–129. [Google Scholar] [CrossRef]
  74. Liu, D.; Xu, Y.; Sun, M.; Huang, Y.; Yu, Y.; Zhang, B. Photothermally Assisted Photocatalytic Conversion of CO2–H2O into Fuels over a WN–WO3 Z-Scheme Heterostructure. J. Mater. Chem. A 2020, 8, 1077–1083. [Google Scholar] [CrossRef]
  75. Mateo, D.; Morlanes, N.; Maity, P.; Shterk, G.; Mohammed, O.F.; Gascon, J. Efficient Visible-Light Driven Photothermal Conversion of CO2 to Methane by Nickel Nanoparticles Supported on Barium Titanate. Adv. Funct. Mater. 2021, 31, 2008244. [Google Scholar] [CrossRef]
  76. Zhao, J.; Liu, B.; Meng, L.; He, S.; Yuan, R.; Hou, Y.; Ding, Z.; Lin, H.; Zhang, Z.; Wang, X.; et al. Plasmonic Control of Solar-Driven CO2 Conversion at the Metal/ZnO Interfaces. Appl. Catal. B Environ. 2019, 256, 117823. [Google Scholar] [CrossRef]
  77. Wang, Y.; Godin, R.; Durrant, J.R.; Tang, J. Efficient Hole Trapping in Carbon Dot/Oxygen-Modified Carbon Nitride Heterojunction Photocatalysts for Enhanced Methanol Production from CO2 under Neutral Conditions. Angew. Chem. Int. Ed. 2021, 60, 20811–20816. [Google Scholar] [CrossRef]
  78. Ye, J.; Liu, C.; Mei, D.; Ge, Q. Methanol Synthesis from CO2 Hydrogenation over a Pd4/In2O3 Model Catalyst: A Combined DFT and Kinetic Study. J. Catal. 2014, 317, 44–53. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, X.; Wang, Y.; Gao, M.; Shen, J.; Pu, X.; Zhang, Z.; Lin, H.; Wang, X. BiVO4/Bi4Ti3O12 Heterojunction Enabling Efficient Photocatalytic Reduction of CO2 with H2O to CH3OH and CO. Appl. Catal. B Environ. 2020, 270, 118876. [Google Scholar] [CrossRef]
  80. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic Reduction of CO2 by H2 for Synthesis of CO, Methanol and Hydrocarbons: Challenges and Opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  81. Yamashita, H.; Fujii, Y.; Ichihashi, Y.; Zhang, S.G.; Ikeue, K.; Park, D.R.; Koyano, K.; Tatsumi, T.; Anpo, M. Selective Formation of CH3OH in the Photocatalytic Reduction of CO2 with H2O on Titanium Oxides Highly Dispersed within Zeolites and Mesoporous Molecular Sieves. Catal. Today 1998, 45, 221–227. [Google Scholar] [CrossRef]
  82. He, Z.-H.; Jiang, C.-S.; Wang, K.; Wang, Z.-Y.; Li, N.; Wang, W.-T.; Liu, Z.-T. Photothermal CO2 Hydrogenation to Methanol over a CoO/Co/TiO2 Catalyst in Aqueous Media under Atmospheric Pressure. Catal. Today 2020, 356, 579–588. [Google Scholar] [CrossRef]
  83. Wu, X.; Li, Y.; Zhang, G.; Chen, H.; Li, J.; Wang, K.; Pan, Y.; Zhao, Y.; Sun, Y.; Xie, Y. Photocatalytic CO2 Conversion of M0.33WO3 Directly from the Air with High Selectivity: Insight into Full Spectrum-Induced Reaction Mechanism. J. Am. Chem. Soc. 2019, 141, 5267–5274. [Google Scholar] [CrossRef]
  84. Xie, B.; Wong, R.J.; Tan, T.H.; Higham, M.; Gibson, E.K.; Decarolis, D.; Callison, J.; Aguey-Zinsou, K.-F.; Bowker, M.; Catlow, C.R.A.; et al. Synergistic Ultraviolet and Visible Light Photo-Activation Enables Intensified Low-Temperature Methanol Synthesis over Copper/Zinc Oxide/Alumina. Nat. Commun. 2020, 11, 1615. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, L.; Kong, G.; Meng, Y.; Tian, J.; Zhang, L.; Wan, S.; Lin, J.; Wang, Y. Direct Coupling of Thermo- and Photocatalysis for Conversion of CO2–H2O into Fuels. ChemSusChem 2017, 10, 4709–4714. [Google Scholar] [CrossRef] [PubMed]
  86. Deng, Y.; Li, J.; Zhang, R.; Han, C.; Chen, Y.; Zhou, Y.; Liu, W.; Wong, P.K.; Ye, L. Solar-Energy-Driven Photothermal Catalytic C–C Coupling from CO2 Reduction over WO3−x. Chin. J. Catal. 2022, 43, 1230–1237. [Google Scholar] [CrossRef]
  87. Li, Y.; Wen, M.; Wang, Y.; Tian, G.; Wang, C.; Zhao, J. Plasmonic Hot Electrons from Oxygen Vacancies for Infrared Light-Driven Catalytic CO2 Reduction on Bi2O3−x. Angew. Chem. Int. Ed. 2021, 60, 910–916. [Google Scholar] [CrossRef]
  88. Wang, Y.; Zhou, Q.; Zhu, Y.; Xu, D. High Efficiency Reduction of CO2 to CO and CH4 via Photothermal Synergistic Catalysis of Lead-Free Perovskite Cs3Sb2I9. Appl. Catal. B Environ. 2021, 294, 120236. [Google Scholar] [CrossRef]
  89. Wang, R.; Zou, Y.; Hong, S.; Xu, M.; Ling, L. High-performance Pt0.01Fe0.05-g-C3N4 Catalyst for Photothermal Catalytic CO2 Reduction. Acta Chim. Sin. 2021, 79, 932. [Google Scholar] [CrossRef]
  90. Li, Y.; Wang, C.; Song, M.; Li, D.; Zhang, X.; Liu, Y. TiO2−x/CoOx Photocatalyst Sparkles in Photothermocatalytic Reduction of CO2 with H2O Steam. Appl. Catal. B Environ. 2019, 243, 760–770. [Google Scholar] [CrossRef]
  91. Novoa-Cid, M.; Baldovi, H.G. Study of the Photothermal Catalytic Mechanism of CO2 Reduction to CH4 by Ruthenium Nanoparticles Supported on Titanate Nanotubes. Nanomaterials 2020, 10, 2212. [Google Scholar] [CrossRef]
  92. Li, D.; Chen, Y.; Abanades, S.; Zhang, Z. Enhanced Activity of TiO2 by Concentrating Light for Photoreduction of CO2 with H2O to CH4. Catal. Commun. 2018, 113, 6–9. [Google Scholar] [CrossRef]
  93. Yu, S.; Wilson, A.J.; Heo, J.; Jain, P.K. Plasmonic Control of Multi-Electron Transfer and C–C Coupling in Visible-Light-Driven CO2 Reduction on Au Nanoparticles. Nano Lett. 2018, 18, 2189–2194. [Google Scholar] [CrossRef]
  94. Sorcar, S.; Thompson, J.; Hwang, Y.; Park, Y.H.; Majima, T.; Grimes, C.A.; Durrant, J.R.; In, S.-I. High-Rate Solar-Light Photoconversion of CO2 to Fuel: Controllable Transformation from C1 to C2 Products. Energy Environ. Sci. 2018, 11, 3183–3193. [Google Scholar] [CrossRef]
  95. Yu, S.; Jain, P.K. Plasmonic Photosynthesis of C1–C3 Hydrocarbons from Carbon Dioxide Assisted by an Ionic Liquid. Nat. Commun. 2019, 10, 2022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Park, H.; Ou, H.-H.; Colussi, A.J.; Hoffmann, M.R. Artificial Photosynthesis of C1–C3 Hydrocarbons from Water and CO2 on Titanate Nanotubes Decorated with Nanoparticle Elemental Copper and CdS Quantum Dots. J. Phys. Chem. A 2015, 119, 4658–4666. [Google Scholar] [CrossRef] [PubMed]
  97. Ijaz, S.; Ehsan, M.F.; Ashiq, M.N.; He, T. Synthesis of a Bi2S3 /CeO2 Nanocatalyst and Its Visible Light-Driven Conversion of CO2 into CH3OH and CH4. Catal. Sci. Technol. 2015, 5, 5208–5215. [Google Scholar] [CrossRef]
  98. Ning, W.; Wang, T.; Chen, H.; Yang, X.; Jin, Y. The Effect of Fe2O3 Crystal Phases on CO2 Hydrogenation. PLoS ONE 2017, 12, e0182955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin Porous G-C3N4 Nanosheets Modified with AuCu Alloy Nanoparticles and C-C Coupling Photothermal Catalytic Reduction of CO to Ethanol. Appl. Catal. B Environ. 2020, 266, 118618. [Google Scholar] [CrossRef]
  100. Li, P.; Liu, L.; An, W.; Wang, H.; Cui, W. Efficient Photothermal Catalytic CO2 Reduction to CH3CH2OH over Cu2O/g-C3N4 Assisted by Ionic Liquids. Appl. Surf. Sci. 2021, 565, 150448. [Google Scholar] [CrossRef]
  101. Védrine, J.C. Importance, Features and Uses of Metal Oxide Catalysts in Heterogeneous Catalysis. Chin. J. Catal. 2019, 40, 1627–1636. [Google Scholar] [CrossRef]
  102. Khdary, N.H.; Alayyar, A.S.; Alsarhan, L.M.; Alshihri, S.; Mokhtar, M. Metal Oxides as Catalyst/Supporter for CO2 Capture and Conversion, Review. Catalysts 2022, 12, 300. [Google Scholar] [CrossRef]
  103. Chen, X.; Chen, Y.; Song, C.; Ji, P.; Wang, N.; Wang, W.; Cui, L. Recent Advances in Supported Metal Catalysts and Oxide Catalysts for the Reverse Water-Gas Shift Reaction. Front. Chem. 2020, 8, 709. [Google Scholar] [CrossRef]
  104. Liu, Y. Recent Advances in Anion-Doped Metal Oxides for Catalytic Applications. J. Mater. Chem. A 2019, 7, 7280–7300. [Google Scholar] [CrossRef]
  105. Chen, S.; Wang, H.; Kang, Z.; Jin, S.; Zhang, X.; Zheng, X.; Qi, Z.; Zhu, J.; Pan, B.; Xie, Y. Oxygen Vacancy Associated Single-Electron Transfer for Photofixation of CO2 to Long-Chain Chemicals. Nat. Commun. 2019, 10, 788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Jia, Z.; Ning, S.; Tong, Y.; Chen, X.; Hu, H.; Liu, L.; Ye, J.; Wang, D. Selective Photothermal Reduction of CO2 to CO over Ni-Nanoparticle/N-Doped CeO2 Nanocomposite Catalysts. ACS Appl. Nano Mater. 2021, 4, 10485–10494. [Google Scholar] [CrossRef]
  107. Li, Z.; Zhang, L.; Huang, W.; Xu, C.; Zhang, Y. Photothermal Catalysis for Selective CO2 Reduction on the Modified Anatase TiO2 (101) Surface. ACS Appl. Energy Mater. 2021, 4, 7702–7709. [Google Scholar] [CrossRef]
  108. Huo, S.; Weng, Z.; Wu, Z.; Zhong, Y.; Wu, Y.; Fang, J.; Wang, H. Coupled Metal/Oxide Catalysts with Tunable Product Selectivity for Electrocatalytic CO 2 Reduction. ACS Appl. Mater. Interfaces 2017, 9, 28519–28526. [Google Scholar] [CrossRef]
  109. Li, Y.; Walsh, A.G.; Li, D.; Do, D.; Ma, H.; Wang, C.; Zhang, P.; Zhang, X. W-Doped TiO2 for Photothermocatalytic CO2 Reduction. Nanoscale 2020, 12, 17245–17252. [Google Scholar] [CrossRef]
  110. Wang, L.; Dong, Y.; Yan, T.; Hu, Z.; Jelle, A.A.; Meira, D.M.; Duchesne, P.N.; Loh, J.Y.Y.; Qiu, C.; Storey, E.E.; et al. Black Indium Oxide a Photothermal CO2 Hydrogenation Catalyst. Nat. Commun. 2020, 11, 2432. [Google Scholar] [CrossRef]
  111. Hu, B.; Guo, Q.; Wang, K.; Wang, X. Enhanced Photocatalytic Activity of Porous In2O3 for Reduction of CO2 with H2O. J. Mater. Sci. Mater. Electron. 2019, 30, 7950–7962. [Google Scholar] [CrossRef]
  112. Xu, M.; Hu, X.; Wang, S.; Yu, J.; Zhu, D.; Wang, J. Photothermal Effect Promoting CO2 Conversion over Composite Photocatalyst with High Graphene Content. J. Catal. 2019, 377, 652–661. [Google Scholar] [CrossRef]
  113. Wu, D.; Deng, K.; Hu, B.; Lu, Q.; Liu, G.; Hong, X. Plasmon-Assisted Photothermal Catalysis of Low-Pressure CO2 Hydrogenation to Methanol over Pd/ZnO Catalyst. ChemCatChem 2019, 11, 1598–1601. [Google Scholar] [CrossRef]
  114. Mateo, D.; Albero, J.; García, H. Graphene Supported NiO/Ni Nanoparticles as Efficient Photocatalyst for Gas Phase CO2 Reduction with Hydrogen. Appl. Catal. B Environ. 2018, 224, 563–571. [Google Scholar] [CrossRef]
  115. Tan, X.; Wu, S.; Li, Y.; Zhang, Q.; Hu, Q.; Wu, J.; Zhang, A.; Zhang, Y. Highly Efficient Photothermocatalytic CO2 Reduction in Ni/Mg-Doped Al2O3 with High Fuel Production Rate, Large Light-to-Fuel Efficiency, and Good Durability. Energy Environ. Mater. 2021, 2, 12193. [Google Scholar]
  116. Zhang, Z.; Gao, Z.; Liu, H.; Abanades, S.; Lu, H. High Photothermally Active Fe2O3 Film for CO2 Photoreduction with H2O Driven by Solar Light. ACS Appl. Energy Mater. 2019, 2, 8376–8380. [Google Scholar] [CrossRef]
  117. Zhao, J.; Yang, Q.; Shi, R.; Waterhouse, G.I.N.; Zhang, X.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. FeO–CeO2 Nanocomposites: An Efficient and Highly Selective Catalyst System for Photothermal CO2 Reduction to CO. NPG Asia Mater. 2020, 12, 5. [Google Scholar] [CrossRef] [Green Version]
  118. Vega-Mendoza, M.S.; Lievano-Hipolito, E.; Torres-Martínez, L.M. Design and Fabrication of Photocatalytic Coatings with α-β-Bi2O3 and Recycled-Fly Ash for Environmental Remediation and Solar Fuel Generation. Ceram. Int. 2021, 47, 26907–26918. [Google Scholar] [CrossRef]
  119. Wei, L.; Huang, G.; Zhang, Y. Dependence of the Intrinsic Phase Structure of Bi2O3 Catalysts on Photocatalytic CO2 Reduction. Catal. Sci. Technol. 2021, 11, 2021–2025. [Google Scholar] [CrossRef]
  120. Li, J.; Ye, Y.; Ye, L.; Su, F.; Ma, Z.; Huang, J.; Xie, H.; Doronkin, D.E.; Zimina, A.; Grunwaldt, J.-D.; et al. Sunlight Induced Photo Thermal Synergistic Catalytic CO2 Conversion via Localized Surface Plasmon Resonance of MoO3−x. J. Mater. Chem. A 2019, 7, 2821–2830. [Google Scholar] [CrossRef] [Green Version]
  121. Oliveira, J.A.; Torres, J.A.; Gonçalves, R.V.; Ribeiro, C.; Nogueira, F.G.E.; Ruotolo, L.A.M. Photocatalytic CO2 Reduction over Nb2O5/Basic Bismuth Nitrate Nanocomposites. Mater. Res. Bull. 2021, 133, 111073. [Google Scholar] [CrossRef]
  122. Novello, P.; Varanasi, C.V.; Liu, J. Effects of Light on Catalytic Activities and Lifetime of Plasmonic Au Catalysts in the CO Oxidation Reaction. ACS Catal. 2019, 9, 578–586. [Google Scholar] [CrossRef]
  123. Li, B.; Gu, T.; Ming, T.; Wang, J.; Wang, P.; Wang, J.; Yu, J.C. (Gold Core)@(Ceria Shell) Nanostructures for Plasmon-Enhanced Catalytic Reactions under Visible Light. ACS Nano 2014, 8, 8152–8162. [Google Scholar] [CrossRef]
  124. Zhang, X.; Li, X.; Reish, M.E.; Zhang, D.; Su, N.Q.; Gutiérrez, Y.; Moreno, F.; Yang, W.; Everitt, H.O.; Liu, J. Plasmon-Enhanced Catalysis: Distinguishing Thermal and Nonthermal Effects. Nano Lett. 2018, 18, 1714–1723. [Google Scholar] [CrossRef] [PubMed]
  125. Li, X.; Everitt, H.O.; Liu, J. Confirming Nonthermal Plasmonic Effects Enhance CO2 Methanation on Rh/TiO2 Catalysts. Nano Res. 2019, 12, 1906–1911. [Google Scholar] [CrossRef]
  126. Zhang, H.; Itoi, T.; Konishi, T.; Izumi, Y. Dual Photocatalytic Roles of Light: Charge Separation at the Band Gap and Heat via Localized Surface Plasmon Resonance To Convert CO2 into CO over Silver–Zirconium Oxide. J. Am. Chem. Soc. 2019, 141, 6292–6301. [Google Scholar] [CrossRef] [PubMed]
  127. Zhang, W.; Wang, L.; Wang, K.; Khan, M.U.; Wang, M.; Li, H.; Zeng, J. Integration of Photothermal Effect and Heat Insulation to Efficiently Reduce Reaction Temperature of CO2 Hydrogenation. Small 2017, 13, 1602583. [Google Scholar] [CrossRef]
  128. Wang, C.; Ranasingha, O.; Natesakhawat, S.; Ohodnicki, P.R.; Andio, M.; Lewis, J.P.; Matranga, C. Visible Light Plasmonic Heating of Au–ZnO for the Catalytic Reduction of CO2. Nanoscale 2013, 5, 6968. [Google Scholar] [CrossRef] [PubMed]
  129. Wang, C.; Fang, S.; Xie, S.; Zheng, Y.; Hu, Y.H. Thermo-Photo Catalytic CO2 Hydrogenation over Ru/TiO2. J. Mater. Chem. A 2020, 8, 7390–7394. [Google Scholar] [CrossRef]
  130. Ravindranathan Thampi, K.; John, K.; Michael, G. Methanation and photo-methanation of carbon dioxide at room temperature and atmospheric pressure. Nature 1987, 327, 506–508. [Google Scholar] [CrossRef]
  131. O’Brien, P.G.; Sandhel, A.; Wood, T.E.; Jelle, A.A.; Hoch, L.B.; Perovic, D.D.; Mims, C.A.; Ozin, G.A. Photomethanation of Gaseous CO2 over Ru/Silicon Nanowire Catalysts with Visible and Near-Infrared Photons. Adv. Sci. 2014, 1, 1400001. [Google Scholar] [CrossRef]
  132. Gu, M.; Liu, D.; Ding, T.; Liu, X.; Chen, T.; Shen, X.; Yao, T. Plasmon-Assisted Photocatalytic CO2 Reduction on Au Decorated ZrO2 Catalysts. Dalton Trans. 2021, 50, 6076–6082. [Google Scholar] [CrossRef]
  133. Humayun, M.; Ullah, H.; Shu, L.; Ao, X.; Tahir, A.A.; Wang, C.; Luo, W. Plasmon Assisted Highly Efficient Visible Light Catalytic CO2 Reduction Over the Noble Metal Decorated Sr-Incorporated g-C3N4. Nano-Micro Lett. 2021, 13, 209. [Google Scholar] [CrossRef]
  134. Dilla, M.; Pougin, A.; Strunk, J. Evaluation of the Plasmonic Effect of Au and Ag on Ti-Based Photocatalysts in the Reduction of CO2 to CH4. J. Energy Chem. 2017, 26, 277–283. [Google Scholar] [CrossRef]
  135. Jia, J.; O’Brien, P.G.; He, L.; Qiao, Q.; Fei, T.; Reyes, L.M.; Burrow, T.E.; Dong, Y.; Liao, K.; Varela, M.; et al. Visible and Near-Infrared Photothermal Catalyzed Hydrogenation of Gaseous CO2 over Nanostructured Pd@Nb2 O5. Adv. Sci. 2016, 3, 1600189. [Google Scholar] [CrossRef] [PubMed]
  136. Ge, H.; Kuwahara, Y.; Kusu, K.; Yamashita, H. Plasmon-Induced Catalytic CO2 Hydrogenation by a Nano-Sheet Pt/HxMoO3−y Hybrid with Abundant Surface Oxygen Vacancies. J. Mater. Chem. A 2021, 9, 13898–13907. [Google Scholar] [CrossRef]
  137. Grote, R.; Habets, R.; Rohlfs, J.; Sastre, F.; Meulendijks, N.; Xu, M.; Verheijen, M.A.; Elen, K.; Hardy, A.; Van Bael, M.K.; et al. Collective Photothermal Effect of Al2O3-supported Spheroidal Plasmonic Ru Nanoparticle Catalysts in the Sunlight-powered Sabatier Reaction. ChemCatChem 2020, 12, 5618–5622. [Google Scholar] [CrossRef]
  138. Kumar, A.; Kumar, A.; Krishnan, V. Perovskite Oxide Based Materials for Energy and Environment-Oriented Photocatalysis. ACS Catal. 2020, 10, 10253–10315. [Google Scholar] [CrossRef]
  139. Hwang, J.; Rao, R.R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in Catalysis and Electrocatalysis. Science 2017, 358, 751–756. [Google Scholar] [CrossRef] [Green Version]
  140. Voorhoeve, R.J.H.; Johnson, D.W.; Pemeika, J.P.; Gallagher, P.K. Perovskite Oxides: Meterials Science in Catalysis. Science 1977, 195, 827–833. [Google Scholar] [CrossRef]
  141. Tanaka, H.; Misono, M. Advances in Designing Perovskite Catalysts. Curr. Opin. Solid State Mater. Sci. 2001, 5, 381–387. [Google Scholar] [CrossRef]
  142. Mateo, D.; Albero, J.; García, H. Titanium-Perovskite-Supported RuO2 Nanoparticles for Photocatalytic CO2 Methanation. Joule 2019, 3, 1949–1962. [Google Scholar] [CrossRef]
  143. Zangeneh, N.P.; Sharifnia, S.; Karamian, E. Modification of Photocatalytic Property of BaTiO3 Perovskite Structure by Fe2O3 Nanoparticles for CO2 Reduction in Gas Phase. Environ. Sci. Pollut. Res. 2020, 27, 5912–5921. [Google Scholar] [CrossRef]
  144. Qin, J.; Lin, L.; Wang, X. A Perovskite Oxide LaCoO3 Cocatalyst for Efficient Photocatalytic Reduction of CO2 with Visible Light. Chem. Commun. 2018, 54, 2272–2275. [Google Scholar] [CrossRef] [PubMed]
  145. Lu, C.; Itanze, D.S.; Aragon, A.G.; Ma, X.; Li, H.; Ucer, K.B.; Hewitt, C.; Carroll, D.L.; Williams, R.T.; Qiu, Y.; et al. Synthesis of Lead-Free Cs3Sb2Br9 Perovskite Alternative Nanocrystals with Enhanced Photocatalytic CO2 Reduction Activity. Nanoscale 2020, 12, 2987–2991. [Google Scholar] [CrossRef] [PubMed]
  146. Shyamal, S.; Dutta, S.K.; Das, T.; Sen, S.; Chakraborty, S.; Pradhan, N. Facets and Defects in Perovskite Nanocrystals for Photocatalytic CO2 Reduction. J. Phys. Chem. Lett. 2020, 11, 3608–3614. [Google Scholar] [CrossRef] [PubMed]
  147. Han, C.; Zhu, X.; Martin, J.S.; Lin, Y.; Spears, S.; Yan, Y. Recent Progress in Engineering Metal Halide Perovskites for Efficient Visible-Light-Driven Photocatalysis. ChemSusChem 2020, 13, 4005–4025. [Google Scholar] [CrossRef]
  148. Shyamal, S.; Dutta, S.K.; Pradhan, N. Doping Iron in CsPbBr3 Perovskite Nanocrystals for Efficient and Product Selective CO2 Reduction. J. Phys. Chem. Lett. 2019, 10, 7965–7969. [Google Scholar] [CrossRef]
  149. Xu, Y.-F.; Lee, M.; Jun, Y.; Ozin, G.A. Perovskite, the Chameleon CO2 Photocatalyst. Cell Rep. Phys. Sci. 2021, 2, 100300. [Google Scholar] [CrossRef]
  150. Shi, R.; Waterhouse, G.I.N.; Zhang, T. Recent Progress in Photocatalytic CO2 Reduction Over Perovskite Oxides. Sol. RRL 2017, 1, 1700126. [Google Scholar] [CrossRef]
  151. Wei, G.; Zheng, D.; Xu, L.; Guo, Q.; Hu, J.; Sha, N.; Zhao, Z. Photothermal Catalytic Activity and Mechanism of LaNixCo1−xO3 (0 ≤ x ≤ 1) Perovskites for CO2 Reduction to CH4 and CH3OH with H2O. Mater. Res. Express 2019, 6, 086221. [Google Scholar] [CrossRef]
  152. Zhang, X.; Zhang, L.; Deng, B.; Jin, J.; Xu, C.; Zhang, Y. Visible Light-Responding Perovskite Oxide Catalysts for Photo-Thermochemical CO2 Reduction. Catal. Commun. 2020, 138, 105955. [Google Scholar] [CrossRef]
  153. Xie, K.; Umezawa, N.; Zhang, N.; Reunchan, P.; Zhang, Y.; Ye, J. Self-Doped SrTiO3−δ Photocatalyst with Enhanced Activity for Artificial Photosynthesis under Visible Light. Energy Environ. Sci. 2011, 4, 4211. [Google Scholar] [CrossRef]
  154. Kim, D.H.; Park, J.L.; Park, E.J.; Kim, Y.D.; Uhm, S. Dopant Effect of Barium Zirconate-Based Perovskite-Type Catalysts for the Intermediate-Temperature Reverse Water Gas Shift Reaction. ACS Catal. 2014, 4, 3117–3122. [Google Scholar] [CrossRef]
  155. Fresno, F.; Jana, P.; Reñones, P.; Coronado, J.M.; Serrano, D.P.; de la Peña O’Shea, V.A. CO2 Reduction over NaNbO3 and NaTaO3 Perovskite Photocatalysts. Photochem. Photobiol. Sci. 2017, 16, 17–23. [Google Scholar] [CrossRef] [PubMed]
  156. Jeong, Y.D.; You, I.; Byeong, S.K.; Sun, M.P.; Misook, K. Preparation of basalt fiber@perovskite PbTiO3 core–shell composites and their effects on CH4 production from CO2 photoreduction. Ceram. Int. 2016, 42, 5942–5951. [Google Scholar]
  157. Chen, X.; Wang, J.; Huang, C.; Zhang, S.; Zhang, H.; Li, Z.; Zou, Z. Barium Zirconate: A New Photocatalyst for Converting CO2 into Hydrocarbons under UV Irradiation. Catal. Sci. Technol. 2015, 5, 1758–1763. [Google Scholar] [CrossRef]
  158. Zhou, H.; Guo, J.; Li, P.; Fan, T.; Zhang, D.; Ye, J. Leaf-Architectured 3D Hierarchical Artificial Photosynthetic System of Perovskite Titanates Towards CO2 Photoreduction Into Hydrocarbon Fuels. Sci. Rep. 2013, 3, 1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Wang, J.; Huang, C.; Chen, X.; Zhang, H.; Li, Z.; Zou, Z. Photocatalytic CO2 Reduction of BaCeO3 with 4f Configuration Electrons. Appl. Surf. Sci. 2015, 358, 463–467. [Google Scholar] [CrossRef]
  160. Kuriki, R.; Sekizawa, K.; Ishitani, O.; Maeda, K. Visible-Light-Driven CO2 Reduction with Carbon Nitride: Enhancing the Activity of Ruthenium Catalysts. Angew. Chem. Int. Ed. 2015, 54, 2406–2409. [Google Scholar] [CrossRef]
  161. Su, X.; Yang, X.-F.; Huang, Y.; Liu, B.; Zhang, T. Single-Atom Catalysis toward Efficient CO2 Conversion to CO and Formate Products. Acc. Chem. Res. 2019, 52, 656–664. [Google Scholar] [CrossRef]
  162. Wang, I.-W.; Kutteri, D.A.; Gao, B.; Tian, H.; Hu, J. Methane Pyrolysis for Carbon Nanotubes and COx-Free H2 over Transition-Metal Catalysts. Energy Fuels 2019, 33, 197–205. [Google Scholar] [CrossRef]
  163. Cheng, Y.; Zhao, S.; Johannessen, B.; Veder, J.; Saunders, M.; Rowles, M.R.; Cheng, M.; Liu, C.; Chisholm, M.F.; Marco, R.; et al. Atomically Dispersed Transition Metals on Carbon Nanotubes with Ultrahigh Loading for Selective Electrochemical Carbon Dioxide Reduction. Adv. Mater. 2018, 30, 1706287. [Google Scholar] [CrossRef]
  164. Levinson, R.; Berdahl, P.; Akbari, H. Solar Spectral Optical Properties of Pigments—Part I: Model for Deriving Scattering and Absorption Coefficients from Transmittance and Reflectance Measurements. Sol. Energy Mater. Sol. Cells. 2005, 89, 319–349. [Google Scholar] [CrossRef]
  165. Abdullah, H.; Khan, M.M.R.; Ong, H.R.; Yaakob, Z. Modified TiO2 Photocatalyst for CO2 Photocatalytic Reduction: An Overview. J. CO2 Util. 2017, 22, 15–32. [Google Scholar] [CrossRef]
  166. Mateo, D.; Morlanes, N.; Maity, P.; Shterk, G.; Mohammed, O.F.; Gascon, J. Photothermal Catalysis: Efficient Visible-Light Driven Photothermal Conversion of CO2 to Methane by Nickel Nanoparticles Supported on Barium Titanate. Adv. Funct. Mater. 2021, 31, 2170053. [Google Scholar] [CrossRef]
  167. Qi, Y.; Song, L.; Ouyang, S.; Liang, X.; Ning, S.; Zhang, Q.; Ye, J. Photoinduced Defect Engineering: Enhanced Photothermal Catalytic Performance of 2D Black In2O3−x Nanosheets with Bifunctional Oxygen Vacancies. Adv. Mater. 2020, 32, 1903915. [Google Scholar] [CrossRef] [PubMed]
  168. Hou, J.; Cao, S.; Wu, Y.; Liang, F.; Sun, Y.; Lin, Z.; Sun, L. Simultaneously Efficient Light Absorption and Charge Transport of Phosphate and Oxygen-Vacancy Confined in Bismuth Tungstate Atomic Layers Triggering Robust Solar CO2 Reduction. Nano Energy 2017, 32, 359–366. [Google Scholar] [CrossRef] [Green Version]
  169. Tian, Y.; Tatsuma, T. Mechanisms and Applications of Plasmon-Induced Charge Separation at TiO2 Films Loaded with Gold Nanoparticles. J. Am. Chem. Soc. 2005, 127, 7632–7637. [Google Scholar] [CrossRef]
  170. Zhang, Y.; Shi, J.; Huang, Z.; Guan, X.; Zong, S.; Cheng, C.; Zheng, B.; Guo, L. Synchronous Construction of CoS2 In-Situ Loading and S Doping for g-C3N4: Enhanced Photocatalytic H2-Evolution Activity and Mechanism Insight. Chem. Eng. J. 2020, 401, 126135. [Google Scholar] [CrossRef]
  171. Khalid, N.R.; Ahmed, E.; Niaz, N.A.; Nabi, G.; Ahmad, M.; Tahir, M.B.; Rafique, M.; Rizwan, M.; Khan, Y. Highly Visible Light Responsive Metal Loaded N/TiO2 Nanoparticles for Photocatalytic Conversion of CO2 into Methane. Ceram. Int. 2017, 43, 6771–6777. [Google Scholar] [CrossRef]
  172. Karamian, E.; Sharifnia, S. Enhanced Visible Light Photocatalytic Activity of BiFeO3-ZnO p-n Heterojunction for CO2 Reduction. Mater. Sci. Eng. B 2018, 238, 142–148. [Google Scholar] [CrossRef]
  173. Mu, Y.-F.; Zhang, C.; Zhang, M.-R.; Zhang, W.; Zhang, M.; Lu, T.-B. Direct Z-Scheme Heterojunction of Ligand-Free FAPbBr3/α-Fe2O3 for Boosting Photocatalysis of CO2 Reduction Coupled with Water Oxidation. ACS Appl. Mater. Interfaces 2021, 13, 22314–22322. [Google Scholar] [CrossRef]
  174. Wei, L.; Yu, C.; Zhang, Q.; Liu, H.; Wang, Y. TiO2-Based Heterojunction Photocatalysts for Photocatalytic Reduction of CO2 into Solar Fuels. J. Mater. Chem. A 2018, 6, 22411–22436. [Google Scholar] [CrossRef]
  175. Huang, L.; Li, B.; Su, B.; Xiong, Z.; Zhang, C.; Hou, Y.; Ding, Z.; Wang, S. Fabrication of Hierarchical Co3O4@CdIn2S4 p–n Heterojunction Photocatalysts for Improved CO2 Reduction with Visible Light. J. Mater. Chem. A 2020, 8, 7177–7183. [Google Scholar] [CrossRef]
  176. Yin, H.Y.; Zheng, Y.F.; Song, X.C. Synthesis and Enhanced Visible Light Photocatalytic CO2 Reduction of BiPO4–BiOBrxI1−xp–n Heterojunctions with Adjustable Energy Band. RSC Adv. 2019, 9, 11005–11012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Bi, Y.; Zong, L.; Li, C.; Li, Q.; Yang, J. Photoreduction of CO2 on TiO2/SrTiO3 Heterojunction Network Film. Nanoscale Res. Lett. 2015, 10, 345. [Google Scholar] [CrossRef] [Green Version]
  178. Ma, Y.; Zhang, Z.; Jiang, X.; Sun, R.; Xie, M.; Han, W. Direct Z-Scheme Sn-In2O3/In2S3 Heterojunction Nanostructures for Enhanced Photocatalytic CO2 Reduction Activity. J. Mater. Chem. C 2021, 9, 3987–3997. [Google Scholar] [CrossRef]
  179. Ijaz, S.; Ehsan, M.F.; Ashiq, M.N.; Karamat, N.; He, T. Preparation of CdS@CeO2 Core/Shell Composite for Photocatalytic Reduction of CO2 under Visible-Light Irradiation. Appl. Surf. Sci. 2016, 390, 550–559. [Google Scholar] [CrossRef]
  180. Pougin, A.; Dodekatos, G.; Dilla, M.; Tüysüz, H.; Strunk, J. Au@TiO2 Core-Shell Composites for the Photocatalytic Reduction of CO2. Chem. Eur. J. 2018, 24, 12416–12425. [Google Scholar] [CrossRef]
  181. Tu, W.; Zhou, Y.; Li, H.; Li, P.; Zou, Z. Au@TiO2 Yolk–Shell Hollow Spheres for Plasmon-Induced Photocatalytic Reduction of CO2 to Solar Fuel via a Local Electromagnetic Field. Nanoscale 2015, 7, 14232–14236. [Google Scholar] [CrossRef]
  182. Fan, W.K.; Tahir, M. Recent Developments in Photothermal Reactors with Understanding on the Role of Light/Heat for CO2 Hydrogenation to Fuels: A Review. Chem. Eng. J. 2022, 427, 131617. [Google Scholar] [CrossRef]
  183. Pérez, S.; Aragón, J.J.; Peciña, I.; Garcia-Suarez, E.J. Enhanced CO2 Methanation by New Microstructured Reactor Concept and Design. Top. Catal. 2019, 62, 518–523. [Google Scholar] [CrossRef] [Green Version]
  184. Cao, X.E.; Kaminer, Y.; Hong, T.; Schein, P.; Liu, T.; Hanrath, T.; Erickson, D. HI-Light: A Glass-Waveguide-Based “Shell-and-Tube” Photothermal Reactor Platform for Converting CO2 to Fuels. iScience 2020, 23, 101856. [Google Scholar] [CrossRef] [PubMed]
  185. Wu, C.; Xing, X.; Yang, G.; Tong, T.; Wang, Z.M.; Bao, J. Understanding the Generation of Long-Chain Hydrocarbons from CO2 and Water Using Cobalt Nanostructures and Light. J. Catal. 2020, 390, 206–212. [Google Scholar] [CrossRef]
  186. Li, Y.; Hao, J.; Song, H.; Zhang, F.; Bai, X.; Meng, X.; Zhang, H.; Wang, S.; Hu, Y.; Ye, J. Selective Light Absorber-Assisted Single Nickel Atom Catalysts for Ambient Sunlight-Driven CO2 Methanation. Nat. Commun. 2019, 10, 2359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Kim, C.; Hyeon, S.; Lee, J.; Kim, W.D.; Lee, D.C.; Kim, J.; Lee, H. Energy-Efficient CO2 Hydrogenation with Fast Response Using Photoexcitation of CO2 Adsorbed on Metal Catalysts. Nat. Commun. 2018, 9, 3027. [Google Scholar] [CrossRef] [Green Version]
  188. Bhatta, S.; Nagassou, D.; Mohsenian, S.; Trelles, J.P. Photo-Thermochemical Decomposition of Carbon-Dioxide in a Direct Solar Receiver-Reactor. Sol. Energy 2019, 178, 201–214. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of the solar methanation process and several different methods [70] Reprinted with permission from Ref. [70]. Copyright @ 2019, Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A.; (b) the possible pathways for the reduction of CO2 into CH4 [71]. Reprinted with permission from Ref. [71]. Copyright @ 2013, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 1. (a) Schematic diagram of the solar methanation process and several different methods [70] Reprinted with permission from Ref. [70]. Copyright @ 2019, Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A.; (b) the possible pathways for the reduction of CO2 into CH4 [71]. Reprinted with permission from Ref. [71]. Copyright @ 2013, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 12 00612 g001
Figure 2. Six possible pathways for CO2 convert into CH4. (a) the hydrogenation on the C atom; (b) the cross-hydrogenation on the O and C atoms [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Figure 2. Six possible pathways for CO2 convert into CH4. (a) the hydrogenation on the C atom; (b) the cross-hydrogenation on the O and C atoms [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Catalysts 12 00612 g002
Figure 3. (a) Mechanism for photothermal CO2 to methanol over a Z-scheme CoO/Co/TiO2 [82]. Reprinted with permission from Ref. [82]. Copyright @ 2020, Elsevier B.V.; (b) CO2 reduction activities of M0.33WO3 series, Rb0.33WO3.165 and W18O49 as well as WO3 under full-spectrum light irradiation; (c) CO2 reduction activities of the samples under NIR light irradiation [83]; Reprinted with permission from Ref. [83]. Copyright @ 2019, American Chemical Society; (d) the light-aided CO2 hydrogenation reaction over the Cu/ZnO/Al2O3 system [84]. Reprinted with permission from Ref. [84]. Copyright @ 2020, Crown.
Figure 3. (a) Mechanism for photothermal CO2 to methanol over a Z-scheme CoO/Co/TiO2 [82]. Reprinted with permission from Ref. [82]. Copyright @ 2020, Elsevier B.V.; (b) CO2 reduction activities of M0.33WO3 series, Rb0.33WO3.165 and W18O49 as well as WO3 under full-spectrum light irradiation; (c) CO2 reduction activities of the samples under NIR light irradiation [83]; Reprinted with permission from Ref. [83]. Copyright @ 2019, American Chemical Society; (d) the light-aided CO2 hydrogenation reaction over the Cu/ZnO/Al2O3 system [84]. Reprinted with permission from Ref. [84]. Copyright @ 2020, Crown.
Catalysts 12 00612 g003
Figure 4. (a) Direct thermal-photocatalytic reduction of CO2-H2O over Au-Ru/TiO2a) [85]; Reprinted with permission from Ref. [85]. Copyright @ 2017, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim; (b) the mechanism for photothermal photothermal catalytic CO2 reduction over WO3−x [86] Reprinted with permission from Ref. [86]. Copyright @ 2022, Elsevier B.V.
Figure 4. (a) Direct thermal-photocatalytic reduction of CO2-H2O over Au-Ru/TiO2a) [85]; Reprinted with permission from Ref. [85]. Copyright @ 2017, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim; (b) the mechanism for photothermal photothermal catalytic CO2 reduction over WO3−x [86] Reprinted with permission from Ref. [86]. Copyright @ 2022, Elsevier B.V.
Catalysts 12 00612 g004
Figure 5. (a) Evolution rate of CH4 over com-WO3, m-WO3 and H2-treated m-WO3 catalysts under Vis-light only; (b) heating only; (c) photothermal coupling; (d) photothermal coupling catalytic CO2 reduction over H2-treated m-WO3 catalyst with H2O [37]. Reprinted with permission from Ref. [37]. Copyright @ 2016, Copyright Royal Society of Chemistry.
Figure 5. (a) Evolution rate of CH4 over com-WO3, m-WO3 and H2-treated m-WO3 catalysts under Vis-light only; (b) heating only; (c) photothermal coupling; (d) photothermal coupling catalytic CO2 reduction over H2-treated m-WO3 catalyst with H2O [37]. Reprinted with permission from Ref. [37]. Copyright @ 2016, Copyright Royal Society of Chemistry.
Catalysts 12 00612 g005
Figure 6. (a) Schematic preparation procedure for plasmonic metal/m-ZnO-x NSs arrays; (b) crystal structure and charge distribution of ZnO NSs along the {001} facet; (c) TEM image of Au/m-ZnO-4.6 sample; (d) HRTEM image of Au/m-ZnO-4.6 sample; (e) HRTEM image of Ag/m-ZnO-1.7 sample; (f) HRTEM image of Pd/m-ZnO-1.9 sample [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Figure 6. (a) Schematic preparation procedure for plasmonic metal/m-ZnO-x NSs arrays; (b) crystal structure and charge distribution of ZnO NSs along the {001} facet; (c) TEM image of Au/m-ZnO-4.6 sample; (d) HRTEM image of Au/m-ZnO-4.6 sample; (e) HRTEM image of Ag/m-ZnO-1.7 sample; (f) HRTEM image of Pd/m-ZnO-1.9 sample [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Catalysts 12 00612 g006
Figure 7. Solar photocatalytic conversion of CO2 to hydrocarbons with water vapor. (a) Hydrocarbon products function as light irradiation time using the Au/m-ZnO-4.6 array; (b) hydrocarbon production rates function as Au loading; (c) hydrocarbon production rates function as Ag loading; (d) hydrocarbon production rates function as Pd loading [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Figure 7. Solar photocatalytic conversion of CO2 to hydrocarbons with water vapor. (a) Hydrocarbon products function as light irradiation time using the Au/m-ZnO-4.6 array; (b) hydrocarbon production rates function as Au loading; (c) hydrocarbon production rates function as Ag loading; (d) hydrocarbon production rates function as Pd loading [76]. Reprinted with permission from Ref. [76]. Copyright @ 2019, Elsevier B.V.
Catalysts 12 00612 g007
Figure 8. (a) the structures and compositions of these different classes of perovskites [149]. Reprinted with permission from Ref. [149]. Copyright @ 1969, Elsevier; (b) various elements commonly used to date to construct stable perovskite-type catalysts for CO2 reduction (including co-catalysts) [150]. Reprinted with permission from Ref. [150]. Copyright @ 2017, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 8. (a) the structures and compositions of these different classes of perovskites [149]. Reprinted with permission from Ref. [149]. Copyright @ 1969, Elsevier; (b) various elements commonly used to date to construct stable perovskite-type catalysts for CO2 reduction (including co-catalysts) [150]. Reprinted with permission from Ref. [150]. Copyright @ 2017, Wiler-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 12 00612 g008
Figure 9. Schematic diagram of different forms of core-shell structure [46]. Reprinted with permission from Ref. [46]. Copyright @ 1920, Copyright Royal Society of Chemisty.
Figure 9. Schematic diagram of different forms of core-shell structure [46]. Reprinted with permission from Ref. [46]. Copyright @ 1920, Copyright Royal Society of Chemisty.
Catalysts 12 00612 g009
Figure 10. (a) Developed multi-channel reactor using additive technologies by Tecnalia R&I (16; 68 and 388 multi-channel reactors developed); (b) reaction yield; (c) CH4 selectivity versus temperature in fixed-bed and multi-channel mini-reactors [183] Reprinted with permission from Ref. [183]. Copyright @ 2019, Copyright Pérez, S.; Aragón, J.J.; Peciña, I.; Garcia-Suarez, E.J.; (d) waveguide-based ‘‘shell-and-tube’’ reactor [184]. Reprinted with permission from Ref. [184]. Copyright @ 2020, Copyright Cao, X.E.; Kaminer, Y.; Hong, T.; Schein, P.; Liu, T.; Hanrath, T.; Erickson, D.
Figure 10. (a) Developed multi-channel reactor using additive technologies by Tecnalia R&I (16; 68 and 388 multi-channel reactors developed); (b) reaction yield; (c) CH4 selectivity versus temperature in fixed-bed and multi-channel mini-reactors [183] Reprinted with permission from Ref. [183]. Copyright @ 2019, Copyright Pérez, S.; Aragón, J.J.; Peciña, I.; Garcia-Suarez, E.J.; (d) waveguide-based ‘‘shell-and-tube’’ reactor [184]. Reprinted with permission from Ref. [184]. Copyright @ 2020, Copyright Cao, X.E.; Kaminer, Y.; Hong, T.; Schein, P.; Liu, T.; Hanrath, T.; Erickson, D.
Catalysts 12 00612 g010
Figure 11. (a) The light-driven temperature; (b) the light-driven CO2 conversion rates of Ni/Y2O3 and Ni/Y2O3 + S nanosheets under different intensities of sunlight irradiation [186] Reprinted with permission from Ref. [186]. Copyright @ 2019,Yaguang Li et al.; (c) the hydrocarbon fuel production rates over WO3−x under UV-Vis, IR and UV-Vis-IR irradiation, and 73, 106 and 177 °C [86]. Reprinted with permission from Ref. [86]. Copyright @ 2022, Elsevier B.V.
Figure 11. (a) The light-driven temperature; (b) the light-driven CO2 conversion rates of Ni/Y2O3 and Ni/Y2O3 + S nanosheets under different intensities of sunlight irradiation [186] Reprinted with permission from Ref. [186]. Copyright @ 2019,Yaguang Li et al.; (c) the hydrocarbon fuel production rates over WO3−x under UV-Vis, IR and UV-Vis-IR irradiation, and 73, 106 and 177 °C [86]. Reprinted with permission from Ref. [86]. Copyright @ 2022, Elsevier B.V.
Catalysts 12 00612 g011
Table 1. Photothermal catalytic products.
Table 1. Photothermal catalytic products.
CatalystReactantReaction ConditionsProductsYield/Distribution/SelectivityRef
Bi2O3−xCO2 and H2420 nm LED
light
200 °C
CO16.15 μmol·g−1·h−1[87]
Pd-TiO2CO2Hg lamp 500 W
500 °C
CO11.05 μmol·g−1·h−1[25]
Cs3Sb2I9CO2 and
H2O
200 mW cm−2 Xenon lamp Visible light
235 °C
CO
CH4 (less)
95.7 μmol·g−1·h−1[88]
Pt0.01Fe0.05-g-C3N4CO2 and H2300 W Xe lamp 360 °CCO7.36 mmol·h−1·gcat−1[89]
Oxygen vacancies TiO2CO2 and
H2
UV light
120 °C
CO22.5 ppm[26]
TiO2−x/CoOxCO2 and
H2O steam
150 W UV lamp
120 °C
CH4
CO
16.464 μmol·g−1·h−1
16.275 μmol·g−1·h−1
[90]
Ru/TiNT nanotubesCO2 and
H2
150 mW/cm2
simulated
sunlight
210 °C
CH412.4 mmol·gcat−1·h−1[91]
TiO2 nanotubeCO2 and
H2O
300 W Xenon lampCH420.67 μmol·g−1·h−1[92]
Au NP colloid in 10% (v/v) aqueous isopropyl alcohol solutionCO2 and
H2O
Visible light
50 °C
CH4
C2H6
6.8 NP−1
5.6 NP−1
[93]
Pt-sensitized
graphene-wrapped blue-colored titania
CO2 and
H2O
Solar spectrum (AM 1.5 G) illuminationCH4
C2H6
259 μmol·g−1
77 μmol·g−1
[94]
Au NPsCO2 and
H2O (ionic liquid medium 5%)
UV-Vis extinctionCH4
C2H4
C2H2
C3H6
C3H8
4.53 NP−1·h−1
1.10 NP−1·h−1
0.99 NP−1·h−1
0.93 NP−1·h−1
0.56 NP−1·h−1
[95]
ex-Ti-oxide/Y-zeolite (1.1 wt% as
TiO2)
CO2 with H2OUV irradiation
53 °C
CH4 CH3OH7.25 μmol·g−1-TiO2·h−1
4.89 μmol·g−1-TiO2·h−1
[81]
Pt-loaded ex-Ti-oxide/Y-zeolite12.25 μmol·g−1-TiO2·h−1
1.38 μmol·g−1-TiO2·h−1
Nf/Pd-TiO2CO2 with H2OUV irradiatedCH4
C2H6
11.5 μmol
10.2 μmol
[29]
CdS/(Cu-NaxH2-xTi3O7)CO2 with H2O450 W Xe lamp
25 °C
CH4
C2H6
C3H8
C2H4
C3H6
28 μL·g−1·h−1
17 μL·g−1·h−1
10 μL·g−1·h−1
2 (×10) μL·g−1·h−1
8 (×10) μL·g−1·h−1
[96]
CoO/Co/TiO2CO2 and
H2O
300 W Xe lamp
120 °C
CH3OH39.6 μmol·gcat−1·h−1[82]
BiVO4/Bi4Ti3O12CO2 and
H2O
300W Xe lamp
298 K
CH3OH
CO
16. 6μmol·g−1·h−1
13.29 μmol·g−1·h−1
[79]
Bi2S3/CeO2CO2 and
H2O
Visible light
15 °C
CH3OH
CH4
1346.8 μmol·g−1
15.84 μmol·g−1
[97]
CeO2-Pt@mSiO2-CoCO2 and
H2 (H2/CO2 = 3)
250 °CCH4
C2-C4
40%
60%
[27]
α-Fe2O3 impregnated with ZnCO2 and
H2
230 °CCO
CH4
C2-C4
C5+
38.1 mol %
13.1 mol %
17.7 C mol %
31.1 mol %
[98]
γ-Fe2O3 impregnated with K27.0 mol %
11.0 mol %
21.1 mol %
40.9 mol %
α-Fe2O3 impregnated with Cu23.4 mol %
10.1 mol %
19.5 mol %
47.0 mol %
BaZr0.5Ce0.3Y0.2O3CO2 and
H2O
UV light
350 °C
CH4
C2H6
C3H8
39.13 umol·g−1
8.64 umol·g−1
3.22 umol·g−1
[39]
Ba(Zr0.5Ce0.3Y0.2)0.05Co0.05O3266.08 umol·g−1
122.69 umol·g−1
21.57 umol·g−1
Ba(Zr0.5Ce0.3Y0.2)0.05Ni0.05O3219.13 umol·g−1
82.95 umol·g−1
12.56 umol·g−1
AuCu/g-C3N4CO2 and
H2O
300 W Xenon lamp (λ > 420 nm) Vis
120°C
CH3CH2OH0.89 mmol·g−1·h−1[99]
Cu2O/g-C3N4CO2 and
H2O (with ionic liquids)
300 W Xenon lamp (λ > 420 nm) Vis
100°C
CH3CH2OH0.71mmol⋅g−1⋅h−1[100]
Table 2. Metal oxide photothermal catalysts for CO2 reduction and products.
Table 2. Metal oxide photothermal catalysts for CO2 reduction and products.
PromoterCatalystReaction ConditionsProductsRef
In2O3In2O3−x/ In2O3Stainless steel batch reactor
300 W Xe lamp (Local temperature 261 °C)
CO (1874.62 μmol·h−1·m−2)
CH3OH (1.48 μmol h−1·m−2)
[110]
1/1-In2O3Visible light irradiation
200 °C
H2 (5.32 µmol·gcat·h−1)
CO (8.25 µmol·gcat−1·h−1)
CH4 (27.19 µmol·gcat·h)
[111]
TiO2CoO/Co/TiO2300 W Xe lamp
120 °C
CH3OH (39.6 µmol·gcat·h)[82]
TiO2-graphene300 W Xe lampCH4 (26.7 μmol·g−1·h−1)
CO (5.2 μmol·g−1·h−1)
[112]
2% CuS/TiO2a 300 W xenon lamp
138 °C
CO (25.97 μmol·g−1·h−1)[22]
Pt/TiO2−xXe lamp (150 W)
120 °C
CH4 (0.3412 μmol·h−1)[63]
ZnOPd/ZnOLight irradiation
250 °C
CH3OH (3.8 mmol·g−1·h−1)
CO (3.6 mmol·g−1·h−1)
[113]
NiONiO/Ni-GUV-Vis light from a 300 W Xe lamp
200 °C
CH4 (642 µmol·gNi−1·h−1)[114]
Al2O3Ni/Mg-Al2O3500 W Xe lamp
450 °C
H2 (69.71 mmol·min−1·g−1)
CO (74.57 mmol·min−1·g−1)
[115]
Fe2O3Fe2O3 filmSolar light
500 °C
CH4 (1470.7 µmol·gcat−1)
C2H4 (736.2 µmol·gcat−1)
C2H6 (277.2 µmol·gcat−1)
[116]
FeO-CeO2FeCe-300Xe lamp
419 °C
CO (19.61 mmol·h−1·gcat−1)[117]
Bi2O3α/β-Bi2O3Visible lightHCOOH (1932 μmol·g−1)
CH3OH (6 μmol·g−1)
[118]
γ-Bi2O3Xe lamp 300 mW cm−2CO (48.10 μmol·h−1·g−1)[119]
MoO3MoO3−xUV-Vis-IR
160 °C
CH4 (2.08 μmol·g−1·h−1)
CO (10.3 μmol·g−1·h−1)
[120]
Nb2O5Pd@Nb2O5300 W Xe lamp
470 °C (local temperature)
CH4 (0.11 μmol·gPd−1·h−1)
CO (0.75 mol·gPd−1·h−1)
[34]
BBN/Nb2O5CO (2.8 μmol·g−1·h−1)
C2H4 (0.1 μmol·g−1·h−1)
[121]
Table 3. Plasmonic-metal photothermal catalysts for CO2 reduction and products.
Table 3. Plasmonic-metal photothermal catalysts for CO2 reduction and products.
CatalystReactantsReaction ConditionsProductsYieldRef
Ag (5.0wt%) ZrO2CO2 and H2UV−visible light
119 °C
CO0.57 μmol·h−1·gcat−1[126]
CO2 and H2O0.0031 μmol·h−1·gcat−1
Au-Ru/TiO2CO2 and H2OHg lamp
85 °C
CH427.1 μmol·g−1·h−1[85]
Au&Pt@ZIFCO2 and H2Xe lamp
150 °C
CH3OH9.1 mmol[127]
Au-ZnOCO2 and H2Visible light
600 °C
CO4.22 mmol·g−1·h−1[128]
Ru/TiO2CO2 and H21.5 G sunlight
300 °C
CH469.49 mmol·gcat−1·h−1[129]
Ru/RuOx/TiO2CO2 and H2Simulated sunlight 80 mW cm−2
46 °C
CH449 mmol·gcat−1·h−1[130]
Ru/Si nanowireCO2 and H2Xe lamp
simulated sunlight (14.5 suns) irradiation
150 °C
CH40.8 mmol·g−1·h−1[131]
Au/ZrO2CO2 and H2OVisible light
Room temperature
CO
CH4
25.6 μmol·g−1·h−1
5.1 μmol·g−1·h−1
[132]
Pt/0.15Sr-C3N4CO2 and H2OVisible lightCH4
CO
48.55 μmol·h−1·g−1
74.54 μmol·h−1·g−1
[133]
Au/TiO2CO2 and H2OUV and visible light
278K
CH436 ppm[134]
Ag/TiO2CH411 ppm
Pd@Nb2O5CO2 and H2300 W Xe lamp at 25 kW m−2
160 °C
CO4.9 mmol·gcat−1·h−1[135]
Pt/HxMoO3-y(Sheet)CO2 and H2Visible light
200 °C
CO120 mmol[136]
Ru/Al2O3CO2 and H26.2 suns
220 °C
CH45.09 mol gRu−1·h−1[137]
Table 4. Perovskite photothermal catalyst for CO2 reduction and products.
Table 4. Perovskite photothermal catalyst for CO2 reduction and products.
CatalystReactantsReaction ConditionsProductsYield or SelectivityRef
LaCo0.6Fe0.4O3CO2 and H2OVisible light
350 °C
CH4
CH3OH
437.28 μmol·g−1
13.75 μmol·g−1
[38]
LaNi0.6Fe0.4O3CO2 and H2OVisible light
350 °C
CH4
CH3OH
471.39 μmol·g−1
15.50 μmol·g−1
[40]
LaSrCoFeO6−δCO2 and H2OVisible light
350 °C
CH4351.32 μmol·g−1 [36]
3DOM-LaSrCoFeO6−δ557.88 μmol·g−1
LaNi0.4Co0.6O3CO2 and H2OVisible light
350 °C
CH4
CH3OH
678.5 μmol·g−1
20.83 μmol·g−1
[151]
RuO2/SrTiO3CO2 and H2UV-Vis Xe lamp
150 °C
CH414.6 mmol·h−1·g−1[142]
SrTiO3CO2Hg lamp 500 W
500 °C
CO1.04 μmol·g−1·h−1[152]
1.5Au-SrTiO37.28 μmol·g−1·h−1
self-doped SrTiO3CO2 and H2OVisible lightCH40.25 mmol m−2cata h−1[153]
BaZr0.8Y0.16Zn0.04O3CO2 and H2600 °CCO0.97[154]
NaNbO3CO2 and H2OUV fluorescent lamps
50 °C
CO
H2
CH4
CH3OH
75.50 μmol·gcat−1
26.35 μmol·gcat−1
1.65 μmol·gcat−1
3.36 μmol·gcat−1
[155]
NaTaO3CO
H2
CH4
CH3OH
74.51 μmol·gcat−1
16.45 μmol·gcat−1
2.35 μmol·gcat−1
3.26 μmol·gcat−1
BF@PbTiO3CO2 and H2OHg lamp UV
303 K
CH4290 μmol·gcat−1·L−1[156]
Ag-BaZrO3CO2 and H2O300W Xe lamp
UV light
CH40.57 μmol·g−1·h−1[157]
Au-CaTiO3CO2 and H2OUV lightCH40.029 μmol·g−1·h−1[158]
Ag-BaCeO3CO2 and H2O300-W Xe lamp
UV light
CH40.55 μmol·g−1·h−1[159]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tian, J.; Han, R.; Guo, Q.; Zhao, Z.; Sha, N. Direct Conversion of CO2 into Hydrocarbon Solar Fuels by a Synergistic Photothermal Catalysis. Catalysts 2022, 12, 612. https://doi.org/10.3390/catal12060612

AMA Style

Tian J, Han R, Guo Q, Zhao Z, Sha N. Direct Conversion of CO2 into Hydrocarbon Solar Fuels by a Synergistic Photothermal Catalysis. Catalysts. 2022; 12(6):612. https://doi.org/10.3390/catal12060612

Chicago/Turabian Style

Tian, Jindan, Ru Han, Qiangsheng Guo, Zhe Zhao, and Na Sha. 2022. "Direct Conversion of CO2 into Hydrocarbon Solar Fuels by a Synergistic Photothermal Catalysis" Catalysts 12, no. 6: 612. https://doi.org/10.3390/catal12060612

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