Next Article in Journal
Investigation on the Role of Pd, Pt, Rh in Methane Abatement for Heavy Duty Applications
Next Article in Special Issue
Ni-Based Catalyst for Carbon Dioxide Methanation: A Review on Performance and Progress
Previous Article in Journal
Catalytic Hydrodeoxygenation of Guaiacol to Cyclohexanol over Bimetallic NiMo-MOF-Derived Catalysts
Previous Article in Special Issue
Research Progress and Reaction Mechanism of CO2 Methanation over Ni-Based Catalysts at Low Temperature: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress in Semiconductor Materials with Application in the Photocatalytic Reduction of CO2

Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control, Collaborative Innovation Center of Atmospheric Environment and Equipment Technology, School of Environmental Science and Engineering, Nanjing University of Information Science & Technology, Nanjing 210044, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(4), 372; https://doi.org/10.3390/catal12040372
Submission received: 8 March 2022 / Revised: 21 March 2022 / Accepted: 22 March 2022 / Published: 24 March 2022

Abstract

:
The large-scale burning of non-renewable fossil fuels leads to the gradual increase of the CO2 concentration in the atmosphere, which is associated with negative impacts on the environment. The consequent need to reduce the emission of CO2 resulting from fossil fuel combustion has led to a serious energy crisis. Research reports indicate that the photocatalytic reduction of CO2 is one of the most effective methods to control CO2 pollution. Therefore, the development of novel high-efficiency semiconductor materials has become an important research field. Semiconductor materials need to have a structure with abundant catalytic sites, among other conditions, which is of great significance for the practical application of highly active catalysts for CO2 reduction. This review systematically describes various types of semiconductor materials, as well as adjustments to the physical, chemical and electronic characteristics of semiconductor catalysts to improve the performance of photocatalytic reduction of CO2. The principle of photocatalytic CO2 reduction is also provided in this review. The reaction types and conditions of photocatalytic CO2 reduction are further discussed. We believe that this review will provide a good basis and reference point for future design and development in this field.

1. Introduction

Over the past hundred years, the establishment and development of industrial civilization has promoted a profound transformation of human social productivity and created huge social wealth for mankind. In the process of industrial development, the annual increase in the consumption of fossil fuels, an important energy source for industrial development, has led to the degradation of the natural carbon cycle as well as a surge in atmospheric CO2 concentration [1,2,3,4]. Especially in recent decades, due to the rapid population increase and the rapid development of industry, the CO2 produced by human respiration and fossil fuel combustion has far exceeded historical levels [5]. In addition, human development has led to ecological destruction, weakening the ability of Nature to absorb CO2 and resulting in increasing levels of CO2 in the atmosphere.
In recent years, studies on CO2 fixation and transformation have gradually developed. Adsorption, storage, reduction and conversion methods can reduce atmospheric CO2 concentration [6]. However, these methods achieve little, and they do not solve the fundamental problem. CO2 is a potential carbon resource, and its utilization is a promising way to reduce CO2 to fuel. Therefore, strategies for CO2 conversion have been developed. These reduction pathways include electrochemical reduction, photochemical reduction, and biotransformation [7,8]. Photochemical reduction is the most promising method for CO2 conversion. It converts Nature’s nearly unlimited solar energy into usable chemical energy, achieving the conversion and utilization of CO2 without consuming fossil energy. At the same time, solar energy is a clean and renewable resource.
In view of this, the conversion of CO2 to hydrocarbons is considered a forward-looking solution because it can solve both the energy problem and the environmental crisis [9,10,11,12,13]. In this regard, artificial photocatalytic reduction is one of the most promising solutions, and the use of this method to photocatalytically reduce CO2 into usable fuels has attracted special attention worldwide [14]. Various semiconductors, including TiO2 [15,16,17], MOFs [18,19,20], BiVO4 [21,22,23] and metal halide perovskites (MHPs) [24,25,26], have been used as photoreduction catalysts for CO2. Although these semiconductor photocatalysts have a wide band gap and can provide enough negative electrons for the photocatalytic reduction of CO2, most of them only respond to ultraviolet light (UV), which severely limits their use of light. In addition, these semiconductor photocatalysts usually have low photocarrier separation and migration efficiency [27,28]. These problems greatly reduce the efficiency of photocatalytic CO2 reduction. Take the most commonly used TiO2 photocatalyst as an example. TiO2 photocatalysts have low electron mobility and poor adsorption activation capacity for CO2 reactants [29]. Therefore, much research has been devoted to various modifications to improve the efficiency of photocatalytic CO2 reduction. The transition metals cobalt and nickel are often used in the study of photocatalytic CO2 reduction due to their rich crustal content, high stability, low cost and high catalytic activity. Cobalt and nickel metal ions can also effectively avoid the generation of high-energy intermediates and promote the multi-electron reduction of CO2, so they are favorable components in photocatalytic CO2 reduction processes [30,31,32]. It has been reported that defect engineering can change the band structure and optical properties of semiconductor photocatalysts, thus improving the light absorption capacity of semiconductors and improving the carrier separation efficiency [33,34,35]. As one of the most efficient solutions, surface modification can optimize the surface atomic states, surface structures and charge properties of semiconductor photocatalysts, and is therefore a promising photocatalytic reaction technology [36,37,38]. Vu et al. [36] reviewed different oxide catalysts for CO2 reduction together with various structural engineering strategies (e.g., nanostructure engineering and heterostructure engineering). Each strategy was found to have its own advantages and disadvantages, so further adjustment is needed to further improve the performance of photocatalysts. For example, photocatalysts are modified with precious metals. These precious metals can play the role of modifier and provide more catalytic active sites for the photocatalytic reduction of CO2. Although these precious-metal modifiers can significantly improve the photoreduction performance of CO2, their rarity and high cost greatly affect their practical application [39,40,41,42]. Therefore, it is necessary to develop non-noble metal or no-metal modified photocatalysts. Given the challenges and limitations of current solutions for photocatalytic CO2 reduction, it is of great significance to design novel photocatalysts to improve the photoreduction of CO2.
However, existing reviews on the photocatalytic reduction of CO2 by semiconductor materials are not comprehensive, and the recent research progress has not been sufficiently summarized. Therefore, this review summarizes the progress of research on semiconductor catalysts which have recently demonstrated excellent performance. In this paper, we review the effects of different types of semiconductor catalysts on the photocatalytic reduction of CO2. In addition, we summarize the mechanisms of photocatalytic CO2 reduction on these semiconductor catalyst. Finally, we discuss future development trends of semiconductor catalysts.

2. Semiconductor Materials for the Photocatalytic Reduction of CO2

The photocatalytic reduction of CO2 is a surface/interface reaction. As discussed above, it is important to find and use raw materials which are environmentally friendly and effective as catalysts. In 1979, Inoue et al. [43] used semiconductor materials such as CdS, TiO2 and WO3 to reduce CO2 to CO, CH4 and other chemical. Figure 1 shows the band structures of different semiconductor photocatalysts. Currently known semiconductor photocatalyst materials for the photocatalytic reduction of CO2 include metal oxides (Co3O4, BiVO4, TiO2), metal chalcogenides (CdS, MoS2, SnS2), nitrides (g-C3N4), bismuth halide oxides (BiOCl, BiOBr) and bimetallic hydroxides (NiAl-LDH), among others. Although the photocatalysts mentioned above have been proven able to reduce CO2, they still have some disadvantages. For example, bismuth halide compounds have strong oxidation ability, but the bottom position of their conduction band is comparatively positive, and their reduction performance is not very good [44,45]. Some photocatalysts have unstable chemical properties and are prone to photocorrosion during photocatalytic reaction, the products of which are toxic to the environment.

2.1. TiO2 Photocatalyst

As an n-type semiconductor material, TiO2 is regarded as an ideal semiconductor material for environmental pollution control due to its good chemical stability, non-toxicity and environmental friendliness. It has great potential application value in the field of environment and energy [47]. In 1972, Fujishima and Honda reported [14] that TiO2 could decompose water molecules into hydrogen and oxygen under ultraviolet light, bringing extensive attention to TiO2 as a photocatalyst material. TiO2 has been widely used in the photocatalytic decomposition of decomposition of hydrogen in aquatic products [48,49,50], the degradation of pollutants [51,52,53], the reduction of CO2 [54,55,56,57] etc., and a series of research achievements have been made. TiO2 photocatalyst materials are also very common in daily life; they are widely used in solar cells, coatings, cosmetics, antibacterial materials and air purifiers, as examples.
TiO2 can be found in the form of three crystal polymorphs: anatase, rutile and brookite [58]. The three crystals have a twisted octahedral structure of six oxygen atoms around a titanium atom. Anatase and rutile forms have tetragonal crystal structures. The different electronic structures of the three crystal types of TiO2 lead to great differences in their photocatalytic performance. Tang et al. [59] studied the effect of different crystal types on the photocatalytic performance of TiO2. It was found that the degradation rate of pollutants was nearly 100% when anatase or mixes with anatase and rutile forms was used as photocatalyst. The degradation rate was less than 15% when pure rutile TiO2 was used as photocatalyst. Jin et al. [60] prepared PbO-decorated TiO2 composites by a one-pot method with highly photoactive CO2 conversion, as shown in Figure 2. The heterojunction formed by the catalyst could effectively inhibit the recombination of photogenerated charge, while the PbO could improve the adsorption of CO2 on the catalyst. Therefore, the photocatalytic activity of the heterojunction complex for CO2 reduction was significantly improved.
Morphology is also an important factor affecting the photocatalytic performance of TiO2. The specific surface areas, active sites, charge transfer rates and exposed crystal surfaces of TiO2 catalysts with different morphologies are significantly different, leading to great differences in their performance. Cao et al. [61] prepared a TiO2 photocatalyst with nanorods and nanorod-hierarchical nanostructures. The catalyst had better photocatalytic performance for CO2 reduction than commercial TiO2 (P25). The high catalytic activity was mainly attributed to the improved charge transfer performance, specific surface area and light absorption performance of the catalyst lent by the nanorod-hierarchical nanostructures. Tan et al. [62] synthesized an Ag/Pd bimetal supported on a N-doped TiO2 nanosheet for CO2 reduction. Due to the modification of the Ag/Pd bimetal and the N doping, the absorption of visible light in the TiO2 nanosheets was improved. This system also provided abundant surface defects and oxygen vacancies for the TiO2 nanosheets, causing the catalyst to exhibit high performance for the photocatalytic reduction of CO2 to CH4. Kar et al. [63] prepared a TiO2 nanotube photocatalyst which showed a highly efficient photocatalytic reduction of CO2 to CH4. The high photocatalytic activity was mainly attributed to the enhancement of visible light absorption by the nanotube structure. In addition to the crystal size and morphology of TiO2 catalyst, the intensity of the external light source also has an effect on the photocatalytic activity of TiO2—generally, as the light intensity increases, more photogenerated electrons are generated by the catalyst’s excitation, and the photocatalytic reaction is promoted as a result. Yang et al. [64] found that the photocatalytic degradation rate of TiO2 to paracetamol attenuated with the decrease of light intensity.

2.2. Co/Ni-Based Catalysts

The transition metals cobalt and nickel, which have various redox states, have been widely used in the study of photocatalytic CO2 reduction due to their rich crustal content, low cost, high catalytic activity for CO2 and strong adsorption capacity [65,66,67]. In 2015, Wang et al. [68] prepared Co-ZIF-9 by a solvothermal method for the study of photocatalytic CO2 reduction. The results showed that the three-dimensional MOFs structure of Co-ZIF-9 was beneficial for CO2 enrichment. The Co2+ ions and imidazole groups played a synergistic role in the photocatalytic reduction of CO2. The Co2+ ion was beneficial for electron transport, and the imidazole group was beneficial for the activation of CO2 molecules. Wang et al. [69,70] also studied the photocatalytic CO2 reduction performance of cobalt-based spinel oxides. MnCo2O4 microspheres and ZnCo2O4 nanorods were prepared by solvothermal calcination and hydrothermal methods, respectively. The results showed that both catalysts exhibited excellent catalytic stability, which confirmed the possibility of Co-based spinel oxides for photocatalytic CO2 reduction. Zhang et al. [71] studied the performance of six Co-MOFs with different coordination environments applied to photocatalytic CO2 reduction, as shown in Figure 3. In pure CO2 atmosphere, the MAF-X27l-OH material had high CO selectivity (98.2%). When the relative pressure of CO2 dropped to 0.1 atm, the conversion rate of MAF-X27l-OH remained at about 80%, while the CO conversion rate of Co-MOF materials without this ligand decreased significantly. These results suggest that MAF-X27l-OH has excellent photocatalytic CO2 reduction performance. These conclusions provide a theoretical basis for future studies on the photocatalytic CO2 reduction of Co-based nanocatalysts.
In addition to Co-based nanomaterials, many Ni-based nanomaterials have also been applied in the study of photocatalytic CO2 reduction. Niu et al. [72] synthesized a spongy nickel-organic heterogeneous photocatalyst (Ni(TPA/TEG), which could effectively adsorb CO2. This novel Ni-based photocatalyst significantly inhibited the production of H2, CH4 and CH3OH during the photocatalytic reduction of CO2. Thus, efficient CO2 conversion was achieved, and the selectivity of CO was close to 100%. Yu et al. [73] obtained a covalent organic framework bearing single Ni sites (Ni-TpBpy) for the selective reduction of CO2 to CO. Figure 4 shows the photocatalytic CO2 reduction performance of Ni-TpBpy under visible light. The results showed that Ni-TpBpy effectively promoted the formation of CO in the reaction medium, and the selectivity of CO was 96% after 5 h of reaction. More importantly, the catalyst maintained 76% CO selectivity in a low CO2 atmosphere. Ni-TpBpy showed good photocatalytic performance for the selective reduction of CO2, which was mainly attributed to the synergistic effect of single Ni catalytic site and TpBpy supporter.

2.3. Metal Halide Perovskites (MHPs)

In recent years, metal halide perovskite (MHP) materials have become attractive in the field of optoelectronics and energy conversion [74]. Compared with traditional semiconductor nanocrystals, these materials have high extinction coefficient, narrow band emission, long carrier diffusion length and high defect tolerance. In addition, the diversity of perovskite structures also allows the band gap to be adjusted in order to enhance light capture [75]. The crystal structure of MHPs is similar to that of oxide perovskite. The chemical formula is ABX3, where A is a monovalent cation, B is a divalent metal cation (the most common are Pb2+ and Sn2+) and X is a halogen ion. MHP nanocrystals exhibit a halogen-rich surface structure. Perovskites can be classified as inorganic or organic-inorganic hybrid halogenated perovskites according to the type of cations in their chemical structure. The emergence of MHPs with unique photoelectric characteristics brings new opportunities for efficient photocatalytic CO2 reduction.
Since the reduction potential of MHPs for CO2 reduction usually changes with the nanocrystal size, the catalytic activity of MHP nanocrystals is also affected by their size. Sun et al. [76] first synthesized CsPbBr3 quantum dots of different sizes to study the effect of quantum dot size on CO2 reduction. It was found that CsPbBr3 with a diameter of 8.5 nm had the longest carrier lifetime, more negative band bottom potential and the highest catalytic activity. As shown in Figure 5, Xu et al. [77] designed CsPbBr3/GO composites by combining graphene oxide (GO) with CsPbBr3 and used them for photocatalytic CO2 reduction. The CsPbBr3/GO composites exhibited higher CO2 reduction activity than pure CsPbBr3 nanocrystals. It was found that the improved CO2 reduction performance was mainly attributed to the existence of efficient charge transfer in CsPbBr3/GO composites, and charge injection graphene oxide effectively promoted the charge injection from MHPs to GO. Other studies have shown that the catalytic activity of MHPs can be improved by coupling various modifiers with MHPs nanocrystals. Man et al. [78] used NHx-rich porous g-C3N4 nanosheets (PCNs) to stabilize CsPbBr3 nanocrystals. The formation of a N–Br bond leads to close contact with g-C3N4 nanocrystals and CsPbBr3 nanocrystals. In order to suppress the serious charge recombination in MHPs, Jiang et al. [79] designed a novel Z-scheme alpha-Fe2O3/amine-RGO/CsPbBr3 catalyst for high-efficiency CO2 reduction. The construction of the Z-scheme heterojunction promoted charge separation and retained strong reducing electrons in CsPbBr3 as well as strong oxidation holes in α-Fe2O3. Finally, it promoted the activity of artificial photosynthesis.

2.4. Metal-Organic Frameworks (MOFs)

Metal-organic framework materials (MOFs) are organic-inorganic hybrid materials with intramolecular pores formed by self-assembly between inorganic metal ions or clusters and organic ligands through coordination bonds. MOFs’ large specific surface area, high porosity and tunable structure make them energy storage materials with strong development potential. The structure of MOFs can be controlled by changing the central metal atoms and the interaction between different organic ligands. Transition metals such as Fe, Co and Ni are often selected as central metal sources. As the raw materials of MOFs, such transition metals are abundant, widely distributed and easily available on Earth, which to some extent makes the raw materials cost of MOFs low. Therefore, MOFs are widely used in catalytic energy conversion and other applications [80]. In photocatalytic processes, MOFs can be used as photocatalysts or modifiers to promote the photocatalytic reaction.
MOFs are a relatively new type of frame material, and their uniformly dispersed metal nodes are conducive to the adsorption and activation of gas molecules. The combination of semiconductors and MOFs to form inorganic-organic nanocomposites can not only satisfy the absorption of photons by semiconductor materials to generate carriers, but also enable the materials to adsorb and activate highly stable CO2 molecules. Xiong et al. [81] developed a method for synthesizing Cu3(BTC)2@TiO2 core-shell structures. The loose shell structure of TiO2 facilitates the passage of CO2 molecules through the shell, while the strong CO2 absorption of Cu3(BTC)2 in the core favors the efficient CO2 reduction of nanocomposites. Wu et al. [82] designed and constructed inorganic perovskite quantum dots and organic MOF complexes. A series of MAPbI(3)@PCN-221(Fe-x) composite photocatalytic materials were prepared by encapsulating quantum dots in MOF channels by a deposition method. In the aqueous phase, the MAPbI(3)@PCN-221(Fe-0.2) composite structure showed a high photocatalytic activity for CO2—much higher than that of PCN-221(Fe-0.2). At the same time, hole oxidation oxidized H2O to O2 to achieve the total decomposition of CO2. Alvaro et al. [83] reported that MOF-5 produced by the coordination of Zn2+ and terephthalic acid has semiconductor-like properties. There is an electron-hole separation and transfer process in which electrons migrate from ligand to adjacent metal nodes for the photocatalytic degradation of phenol. Subsequently, a great deal of research was conducted on the construction and development of MOFs with semiconductor properties. On the one hand, MOFs act as light absorbers. On the other hand, due to their unique structural characteristics, they can be used as highly efficient catalysts integrating light absorption and catalytic active sites [84]. Fang et al. [85] designed and synthesized a pyrazolyl porphyrinic Ni-MOF (PCN-601) with active sites of light capture exposure and high specific surface area, as shown in Figure 6. The experimental results showed that PCN-601 had high-efficiency CO2 photoreduction performance under the condition of CO2-saturated water vapor. Under the condition of simulated solar irradiation (AM 1.5 G), the CH4 yield of PCN-601 reached 92 μmol·h−1·g−1 and the apparent quantum yield (AQY) of CH4 was 2.18%. Therefore, the accumulated research suggests that MOFs have the best interfacial charge transfer and kinetic process among the materials evaluated for photocatalytic CO2 reduction. The optimal catalytic activity was obtained in the catalytic reaction system, which provided a new direction for guiding and controlling the synthesis of new MOF materials.

2.5. Other Semiconductor Photocatalysts

Compared with commonly used photocatalysts (MHPs, MOFs, TiO2, etc.), other efficient photocatalytic materials for the reduction of CO2 need to be developed and studied. The conduction band position of SiC has a comparatively more negative potential, which can produce photogenerated electrons with stronger reduction ability for the photocatalytic reduction of CO2. However, the synthesis of SiC in a high-temperature protective atmosphere is not conducive to the regulation of its nanostructure [86,87]. Layered double hydroxides (LDHs) such as Zn–Al LDH [88], Mg–Al LDH [89] and Zn–Cu–Ga LDH [90] have been used for the photocatalytic reduction of CO2. Teramura et al. [91] synthesized a variety of LDHs with surface alkaline sites for the photocatalytic conversion of CO2 to CO. Their activities tend to be higher than that of pure hydroxide. Graphite carbon nitride (g-C3N4) is a metal-free polymer material considered to be a promising visible-light catalyst [92,93]. Hsu et al. [94] used graphene oxide as a catalyst for the efficient photocatalytic conversion of CO2 to methanol, and synthesized a graphene catalyst with an improved method to improve the activity. The yield of CH3OH was six times that obtained using pure TiO2. The photocatalytic activity of highly porous Ga2O3 for the reduction of CO2 was found to be more than four times that of commercial Ga2O3 without the need for additives [95]. The performance of porous Ga2O3 was improved by doubling the surface area and tripling the adsorption capacity. Tanaka et al. [96] used H2 instead of H2O as a reducing agent to photocatalytically reduce CO2 on Ga2O3, and the product was CO instead of CH4. Notably, nearly 7.3% of the surface-adsorbed CO2 was converted.
An effective strategy is to develop new semiconductor photocatalysts with visible-light response to improve the utilization rate of sunlight, so as to obtain higher photocatalytic activity. Zhou et al. [97] prepared Bi2WO6 square nanoplates by a hydrothermal method, and the product obtained by reducing CO2 was CH4. Figure 7 shows the result of CH4 formation on Bi2WO6 photocatalysts. Cheng et al. [98] successfully prepared hollow Bi2WO6 microspheres, and methanol was obtained by reducing CO2. Xi et al. [99] prepared W18O49 nanowires by a one-step liquid-phase method. Under visible-light irradiation, CO2 was photocatalytically reduced in water vapor to generate CH4. The average formation rate of CH4 was significantly increased by using Pt and Au as cocatalysts.

3. Principle of the Photocatalytic Reduction of CO2

3.1. Process of the Photocatalytic Reduction of CO2

Among many catalytic technologies, green and cheap semiconductor photocatalytic technology provides a new approach to the realization of this process. The inspiration for the semiconductor photocatalytic reduction of CO2 comes from the photosynthesis of green plants in Nature. In Nature, plants absorb solar energy to convert CO2 and H2O into organic matter and release oxygen to sustain their growth. Artificial photosynthesis using synthetic photocatalysts is used to simulate natural photosynthesis and reduce CO2 into chemical products with high added value using sunlight. This is the most promising environment-friendly technology for CO2 reduction at present.
Figure 8 shows a diagram of the mechanism of photocatalytic CO2 reduction using semiconductor materials [100]. It can be seen from the figure that the photocatalytic reduction of CO2 by semiconductor materials involves three basic processes. First, when the semiconductor photocatalyst is irradiated by incident light with an energy greater than or equal to its band gap (E.g.), the electrons in the valence band are excited and transition to the conduction band to form excited-state electrons with strong reduction ability. Holes with strong oxidation capacity equal to the number of electrons are left in the valence band, thus forming electron-hole pairs. The electrons and holes are then transferred to the surface of the semiconductor catalyst. Finally, the electrons reduce the CO2 molecules on the surface of the catalyst, and the holes convert the H2O molecules on the surface of the catalyst into O2. Most of the photogenerated electrons and holes recombine because the recombination rate of electrons and holes is much faster than the rate of surface redox reactions. Energy is released as light or heat, and only a small fraction of the electrons/holes can reach the surface of the catalyst to participate in the redox reaction. Therefore, it is very important to improve the separation efficiency of electrons and holes in the photocatalytic reduction of CO2.

3.2. Types of Photocatalytic CO2 Reduction

The H2O molecule is an ideal reducing agent for the reduction of CO2, providing protons to consume holes and producing oxygen and hydrogen during the reaction. At present, most photocatalytic CO2 reduction reactions need to build reaction systems based on H2O molecules. The reaction system mainly includes two types, namely, gas-solid and liquid-solid systems. Figure 9 shows schematics of gas-solid and liquid-solid systems for photocatalytic CO2 reduction.
In a gas-solid system, the H2O molecule is present as water vapor and the catalyst is uniformly dispersed at the bottom of the reactor or uniformly coated on the substrate and placed at the bottom of the reactor. CO2 gas flows into the reactor as a continuous stream or exists in the closed off-line system as a saturated gas. Zhang et al. [101] found that if TiO2 nanotubes supported by Pt were used as photocatalyst, methane production increased when the concentration of water around the catalyst and the surface –OH group was relatively high. However, when Pt–TiO2 nanoparticles were used as catalyst, the CO2/H2O ratio had no effect on methane production. This indicates that the adsorption energy of water molecules on the catalyst surface affects the activity of the gas-phase photocatalytic system. Due to the microporous structure and surface properties of the materials, the adsorption of CO2 and H2O on the active sites leads to different photocatalytic processes and the formation of different products [102].
In a liquid-solid system, the catalyst is suspended in an aqueous solution. Before the reaction, CO2 gas is bubbled into the water to reach the saturation state [103]. The yield of product when using H2O as reducing agent is very low. It was found that adding triethanolamine or isopropanol into the solution consumes holes to improve the efficiency of photocatalytic CO2 reduction. Kaneco et al. [104] used TiO2 to photocatalytically reduce CO2 in isopropanol solution. The CO2 was reduced to methane, and isopropanol was oxidized to acetone. Another factor with a great influence on the reaction system is the suspension pH. Alkaline substances (NaOH, NaHCO3, etc.) are usually added to the solution to improve the solubility of CO2 and promote the reduction of CO2 [105]. The pH of the suspension decreased from about 14 to 7.5, indicating that the solution mainly contained HCO3 and a small amount of CO32−, which could have accelerated the photocatalytic reduction of CO2. The OH ions in alkaline solution can be used as strong hole-trapping agents to improve photocatalytic activity. Wu et al. [106] significantly increased the yield of CH3OH by adding NaOH solution to the photocatalytic system solution. In addition, the addition of electrolytes increased total production as well as C2 products such as ethanol and acetaldehyde.
The difference between the gas–solid and liquid–solid systems mainly lies in the different concentrations of CO2 and H2O on the surface of the catalyst; the products under different concentrations are also different. The disadvantage of gas-solid systems is that the contact area between reactants and catalyst is small, while the disadvantage of liquid-solid systems is that a large number of H2O molecules occupy adsorption sites for CO2. Therefore, different optimization designs of catalysts are needed for different reaction systems in order to overcome the defects of the reaction system itself and maximize the efficiency of photocatalytic CO2 reduction.

3.3. Reaction Conditions of Photocatalytic CO2 Reduction

Photocatalytic reactions are usually influenced by external environmental and operational conditions, such as the wavelength range and intensity of light, the quality of the catalyst, the pH value of the solution, the pressure of CO2, the reducing agent and the temperature. Hou et al. [107] studied the influence of different irradiation sources in the photocatalytic reduction of CO2 with Au nanoparticle/TiO2 catalyst, and generated different products such as CH4, CO, CH3OH and HCHO. It was found that the photocatalytic reduction product of aqueous CO2 by Au nanoparticle/TiO2 catalyst under different excitation wavelengths (λ = 532 nm, 254 nm and 365 nm) was CH4. However, additional reaction products (including C2H6, CH3OH and HCHO) were observed under UV light (254 nm mercury lamp). The photocatalytic activity and selectivity can also be effectively improved by increasing the concentration of CO2. Mizuno et al. [108] reported that with the increase of CO2 pressure, the adsorption rate of hydrogen on the TiO2 surface gradually exceeded that of CO2. This adsorbed hydrogen began to form low-mass hydrocarbons such as CH4 and CH2CH2. In gas-phase systems, increasing the pressure of CO2 and water vapor can also promote the binding of reactants at the active sites of the catalyst, thus improving the photocatalytic activity of CO2 reduction [101]. A further variable that must be considered in the photocatalytic reduction processes is the system temperature. Fox and Dulay believed that a small temperature change had little effect on the photocatalytic reaction [109]; however, high temperature not only increases the collision frequency and diffusivity, but also increases the rate of thermal activation steps. Guan et al. [110] used a Pt-loaded K2Ti6O13 photocatalyst combined with an Fe-based catalyst to reduce CO2 and water vapor under sunlight and high temperature. It was found that the yield of HCOOH, CH3OH and C2H5OH increased significantly with the increase of reaction temperature.
The crystal phase structure of the catalyst must also be considered. If the structure is the same but the exposed crystal plane is different, the photocatalytic CO2 reduction performance will be different. Yamashita et al. [111] studied the photocatalytic CO2 reduction by TiO2 single crystal with rutile phase, exposed to (100) and (110) surfaces. They found that the activity of the (100) surface was significantly higher than that of the (110) surface, and CH4/CH3OH products could be detected in the TiO2 system exposed to the (100) surface. It can be concluded that optimizing the microstructure of catalytic materials can significantly improve the photocatalytic CO2 reduction activity.

4. Conclusions and Perspectives

Since the first Industrial Revolution, with the increasing frequency of human activities and the massive burning of fossil fuels, the emission of CO2 into the atmosphere has been increasing. This will lead to global warming, sea level rise, land desertification and other extreme climate and environmental problems. At the same time, the rapid development of industry has led to sharp increases in energy consumption worldwide. However, as non-renewable resources, fossil energy reserves are very limited. The development of new renewable energy sources is currently an urgent need of human society. One promising strategy for the mitigation of negative effects brought by CO2 emissions is to simulate natural photosynthesis and use clean and sustainable solar energy to convert CO2 into hydrocarbon fuels with added value; this could simultaneously alleviate environmental pollution and energy shortage in today’s human society. However, because CO2 is a symmetrical linear molecule and carbon is in its highest oxidation state, it has high thermodynamic stability, making the process of CO2 reduction by visible light extremely challenging.
Many studies have been published on the development and design of efficient photocatalysts to achieve photocatalytic CO2 reduction, but the photocatalytic reduction conversion efficiency is far from sufficient to meet the needs of practical application. This is usually caused by the high recombination rate of photogenerated electron-hole pairs and the poor adsorption activation capacity of CO2 molecules in photocatalytic CO2 reduction processes. In order to solve the bottleneck problem of low CO2 adsorption activation capacity in the process of CO2 photoreduction, it is necessary to modify the catalyst. There are many common catalytic material modification methods, such as morphology control, surface microstructure control and grain size control.
In the future, the research focus of semiconductor catalysts for the photocatalytic reduction of CO2 should still be on improving the efficiency of photocatalytic reduction of CO2. The main research direction should be the development of new semiconductor catalysts and optimized reaction conditions. In addition, the study of the reaction mechanism and type of photocatalytic reduction of CO2 on semiconductor catalysts is helpful to find ways to improve the catalyst activity. Meanwhile, optimizing the reaction path is of great significance for the large-scale synthesis of catalysts in the future industrialization process. From an economic perspective, Co/Ni-based catalysts have the lowest cost among the catalysts (TiO2 photocatalyst, Co/Ni-based catalysts, MHPs, MOFs and other semiconductor photocatalysts) mentioned in this review, while MHP catalysts are relatively expensive. Finally, in order to achieve green catalysis, the preparation cost of the catalyst must be considered, and the catalyst should be applied to the industrial photocatalytic reduction of CO2 through improved reactors and other methods.

Author Contributions

Conceptualization, M.C. and Y.C.; formal analysis, P.G.; investigation, Y.C.; resources, M.C.; data curation, P.G.; writing—original draft preparation, Y.C.; writing—review and editing, L.X.; funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 21976094, 22176100) and the National Key Research and Development Project 9 (grant number 2018YFC0213802).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Varga, B.; Vida, G.; Varga-Laszlo, E.; Hoffmann, B.; Veisz, O. Combined Effect of Drought Stress and Elevated Atmospheric CO2 Concentration on the Yield Parameters and Water Use Properties of Winter Wheat (Triticum aestivum L.) Genotypes. J. Agron. Crop. Sci. 2017, 203, 192–205. [Google Scholar] [CrossRef]
  2. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  3. Liddicoat, S.K.; Wiltshire, A.J.; Jones, C.D.; Arora, V.K.; Brovkin, V.; Cadule, P.; Hajima, T.; Lawrence, D.M.; Pongratz, J.; Schwinger, J.; et al. Compatible Fossil Fuel CO2 Emissions in the CMIP6 Earth System Models’ Historical and Shared Socioeconomic Pathway Experiments of the Twenty-First Century. J. Clim. 2021, 34, 2853–2875. [Google Scholar] [CrossRef]
  4. Yin, Y.; Bowman, K.; Bloom, A.A.; Worden, J. Detection of Fossil Fuel Emission Trends in the Presence of Natural Carbon Cycle Variability. Environ. Res. Lett. 2019, 14, 084050. [Google Scholar] [CrossRef]
  5. Vermeer, M.; Rahmstorf, S. Global Sea Level Linked to Global Temperature. Proc. Natl. Acad. Sci. USA 2009, 106, 21527–21532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Chang, J.; Clay, D.E.; Clay, S.A.; Chintala, R.; Miller, J.M.; Schumacher, T. Biochar Reduced Nitrous Oxide and Carbon Dioxide Emissions from Soil with Different Water and Temperature Cycles. Agron. J. 2016, 108, 2214–2221. [Google Scholar] [CrossRef] [Green Version]
  7. Wang, Y.; Han, P.; Lv, X.; Zhang, L.; Zheng, G. Defect and Interface Engineering for Aqueous Electrocatalytic CO2 Reduction. Joule 2018, 2, 2551–2582. [Google Scholar] [CrossRef] [Green Version]
  8. Xie, H.; Wang, Y.; He, Y.; Gou, M.; Liu, T.; Wang, J.; Tang, L.; Jiang, W.; Zhang, R.; Xie, L.; et al. Generation of Electricity from CO2 Mineralization: Principle and Realization. Sci. China Technol. Sci. 2014, 57, 2335–2343. [Google Scholar] [CrossRef]
  9. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D Heterojunction of Ultrathin Mxene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  10. Xia, P.; Zhu, B.; Yu, J.; Cao, S.; Jaroniec, M. Ultra-Thin Nanosheet Assemblies of Graphitic Carbon Nitride for Enhanced Photocatalytic CO2 Reduction. J. Mater. Chem. A 2017, 5, 3230–3238. [Google Scholar] [CrossRef]
  11. Low, J.; Cheng, B.; Yu, J. Surface Modification and Enhanced Photocatalytic CO2 Reduction Performance of TiO2: A Review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  12. 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] [PubMed]
  13. Liu, L.; Zhao, H.; Andino, J.M.; Li, Y. Photocatalytic CO2 Reduction with H2O on TiO2 Nanocrystals: Comparison of Anatase, Rutile, and Brookite Polymorphs and Exploration of Surface Chemistry. ACS Catal. 2012, 2, 1817–1828. [Google Scholar] [CrossRef]
  14. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  15. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. Improved Interfacial Bonding of Graphene-TiO2 with Enhanced Photocatalytic Reduction of CO2 into Solar Fuel. J. Environ. Chem. Eng. 2018, 6, 6947–6957. [Google Scholar] [CrossRef]
  16. 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]
  17. Cai, J.; Shen, F.; Shi, Z.; Lai, Y.; Sun, J. Nanostructured TiO2 for Light-Driven CO2 Conversion into Solar Fuels. APL Mater. 2020, 8, 040914. [Google Scholar] [CrossRef] [Green Version]
  18. Han, B.; Ou, X.; Deng, Z.; Song, Y.; Tian, C.; Deng, H.; Xu, Y.-J.; Lin, Z. Nickel Metal-Organic Framework Monolayers for Photoreduction of Diluted CO2: Metal-Node-Dependent Activity and Selectivity. Angew. Chem. Int. Ed. 2018, 57, 16811–16815. [Google Scholar] [CrossRef]
  19. Song, K.; Qiu, X.; Han, B.; Liang, S.; Lin, Z. Efficient Upcycling Electroplating Sludge and Waste Pet into Ni-MOF Nanocrystals for the Effective Photoreduction of CO2. Environ. Sci.-Nano. 2021, 8, 390–398. [Google Scholar] [CrossRef]
  20. Wang, H.-N.; Sun, H.-X.; Fu, Y.-M.; Meng, X.; Zou, Y.-H.; He, Y.-O.; Yang, R.-G. Varied Proton Conductivity and Photoreduction CO2 Performance of Isostructural Heterometallic Cluster Based Metal-Organic Frameworks. Inorg. Chem. Front. 2021, 8, 4062–4071. [Google Scholar] [CrossRef]
  21. Yu, X.; Ordomsky, V.V.; Khodakov, A.Y. Selective Deposition of Cobalt and Copper Oxides on BiVO4 Facets for Enhancement of CO2 Photocatalytic Reduction to Hydrocarbons. ChemCatChem 2020, 12, 740–749. [Google Scholar] [CrossRef]
  22. 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]
  23. Li, X.; Wei, D.; Ye, L.; Li, Z. Fabrication of Cu2O-RGO/BiVO4 Nanocomposite for Simultaneous Photocatalytic CO2 Reduction and Benzyl Alcohol Oxidation under Visible Light. Inorg. Chem. Commun. 2019, 104, 171–177. [Google Scholar] [CrossRef]
  24. Que, M.; Zhao, Y.; Yang, Y.; Pan, L.; Lei, W.; Cai, W.; Yuan, H.; Chen, J.; Zhu, G. Anchoring of Formamidinium Lead Bromide Quantum Dots on Ti3C2 Nanosheets for Efficient Photocatalytic Reduction of CO2. ACS Appl. Mater. Interfaces 2021, 13, 6180–6187. [Google Scholar] [CrossRef] [PubMed]
  25. Ren, K.; Yue, S.; Li, C.; Fang, Z.; Gasem, K.A.M.; Leszczynski, J.; Qu, S.; Wang, Z.; Fan, M. Metal Halide Perovskites for Photocatalysis Applications. J. Mater. Chem. A 2022, 10, 407–429. [Google Scholar] [CrossRef]
  26. Park, S.; Choi, S.; Kim, S.; Nam, K.T. Metal Halide Perovskites for Solar Fuel Production and Photoreactions. J. Phys. Chem. Lett. 2021, 12, 8292–8301. [Google Scholar] [CrossRef]
  27. Wang, S.; Wang, L.; Huang, W. Bismuth-Based Photocatalysts for Solar Energy Conversion. J. Mater. Chem. A 2020, 8, 24307–24352. [Google Scholar] [CrossRef]
  28. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  29. Yang, Y.; Liu, J.; Liu, F.; Wu, D. Reaction Mechanism of CO2 Methanation over Rh/TiO2 Catalyst. Fuel 2020, 276, 118093. [Google Scholar] [CrossRef]
  30. Do, J.Y.; Kwak, B.S.; Park, S.-M.; Kang, M. Effective Carbon Dioxide Photoreduction over Metals (Fe-, Co-, Ni-, and Cu-) Incorporated TiO2/Basalt Fiber Films. Int. J. Photoenergy 2016, 2016, 5195138. [Google Scholar] [CrossRef] [Green Version]
  31. Yan, Z.-H.; Ma, B.; Li, S.-R.; Liu, J.; Chen, R.; Du, M.-H.; Jin, S.; Zhuang, G.-L.; Long, L.-S.; Kong, X.-J.; et al. Encapsulating a Ni(II) Molecular Catalyst in Photoactive Metal-Organic Framework for Highly Efficient Photoreduction of CO2. Sci. Bull. 2019, 64, 976–985. [Google Scholar] [CrossRef] [Green Version]
  32. Han, B.; Song, J.; Liang, S.; Chen, W.; Deng, H.; Ou, X.; Xu, Y.-J.; Lin, Z. Hierarchical NiCo2O4 Hollow Nanocages for Photoreduction of Diluted CO2: Adsorption and Active Sites Engineering. Appl. Catal. B Environ. 2020, 260, 118208. [Google Scholar] [CrossRef]
  33. Xu, S.; Carter, E.A. Theoretical Insights into Heterogeneous (Photo)Electrochemical CO2 Reduction. Chem. Rev. 2019, 119, 6631–6669. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, H.; Teng, F.; Gu, W.; Liu, Z.; Zhao, Y.; Zhang, A.; Liu, Z.; Teng, Y. A Simple Post-Synthesis Conversion Approach to Zn(OH)F and the Effects of Fluorine and Hydroxyl on the Photodegradation Properties of Dye Wastewater. J. Hazard. Mater. 2017, 333, 250–258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, Z.; Wang, L.; Liu, W.; Yan, Z.; Zhu, Y.; Zhou, S.; Guan, S. Photogenerated-Hole-Induced Rapid Elimination of Solid Tumors by the Supramolecular Porphyrin Photocatalyst. Natl. Sci. Rev. 2021, 8, 155. [Google Scholar] [CrossRef] [PubMed]
  36. Vu, N.-N.; Kaliaguine, S.; Do, T.-O. Critical Aspects and Recent Advances in Structural Engineering of Photocatalysts for Sunlight-Driven Photocatalytic Reduction of CO2 into Fuels. Adv. Funct. Mater. 2019, 29, 1901825. [Google Scholar] [CrossRef]
  37. Chen, C.; Ma, W.; Zhao, J. Semiconductor-Mediated Photodegradation of Pollutants under Visible-Light Irradiation. Chem. Soc. Rev. 2010, 39, 4206–4219. [Google Scholar] [CrossRef]
  38. Li, C.-J.; Xu, G.-R.; Zhang, B.; Gong, J.R. High Selectivity in Visible-Light-Driven Partial Photocatalytic Oxidation of Benzyl Alcohol into Benzaldehyde over Single-Crystalline Rutile TiO2 Nanorods. Appl. Catal. B Environ. 2012, 115, 201–208. [Google Scholar] [CrossRef]
  39. Lertthanaphol, N.; Pienutsa, N.; Chusri, K.; Sornsuchat, T.; Chanthara, P.; Seeharaj, P.; Kim-Lohsoontorn, P.; Srinives, S. One-Step Hydrothermal Synthesis of Precious Metal-Doped Titanium Dioxide-Graphene Oxide Composites for Photocatalytic Conversion of CO2 to Ethanol. ACS Omega 2021, 6, 35769–35779. [Google Scholar] [CrossRef]
  40. Xing, F.; Liu, Q.; Song, M.; Huang, C. Fluorine Modified Boron Carbon Nitride Semiconductors for Improved Photocatalytic CO2 Reduction under Visible Light. ChemCatChem 2018, 10, 5270–5279. [Google Scholar] [CrossRef]
  41. Ma, Z.; Li, P.; Ye, L.; Wang, L.; Xie, H.; Zhou, Y. Selectivity Reversal of Photocatalytic CO2 Reduction by Pt Loading. Catal. Sci. Technol. 2018, 8, 5129–5132. [Google Scholar] [CrossRef]
  42. Bowker, M.; Bahruji, H.; Kennedy, J.; Jones, W.; Hartley, G.; Morton, C. The Photocatalytic Window: Photo-Reforming of Organics and Water Splitting for Sustainable Hydrogen Production. Catal. Lett. 2015, 145, 214–219. [Google Scholar] [CrossRef] [Green Version]
  43. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic Reduction of Carbon Dioxide in Aqueous Suspensions of Semiconductor Powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  44. Moebs, J.; Gerhard, M.; Heine, J. (Hpy)2(Py)CuBi3I12, a Low Bandgap Metal Halide Photoconductor. Dalton Trans. 2020, 49, 14397–14400. [Google Scholar] [CrossRef] [PubMed]
  45. Park, B.-W.; Philippe, B.; Zhang, X.; Rensmo, H.; Boschloo, G.; Johansson, E.M.J. Bismuth Based Hybrid Perovskites A3Bi2I9 (A: Methylammonium or Cesium) for Solar Cell Application. Adv. Mater. 2015, 27, 6806–6813. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  47. Shen, Y.; Ren, F.; Liu, H. Progress in Research on Photocatalytic Performance of Doped-TiO2. Rare Met. Mater. Eng. 2006, 35, 1841–1844. [Google Scholar]
  48. Miodynska, M.; Mikolajczyk, A.; Bajorowicz, B.; Zwara, J.; Klimczuk, T.; Lisowski, W.; Trykowski, G.; Pinto, H.P.; Zaleska-Medynska, A. Urchin-Like TiO2 Structures Decorated with Lanthanide-Doped Bi2S3 Quantum Dots to Boost Hydrogen Photogeneration Performance. Appl. Catal. B Environ. 2020, 272, 118962. [Google Scholar] [CrossRef]
  49. Wang, J.; Shen, Y.; Liu, S.; Zhang, Y. Single 2D Mxene Precursor-Derived TiO2 Nanosheets with a Uniform Decoration of Amorphous Carbon for Enhancing Photocatalytic Water Splitting. Appl. Catal. B Environ. 2020, 270, 118885. [Google Scholar] [CrossRef]
  50. Li, F.; Xiao, X.; Zhao, C.; Liu, J.; Li, Q.; Guo, C.; Tian, C.; Zhang, L.; Hu, J.; Jiang, B. TiO2-on-C3N4 Double-Shell Microtubes: In-Situ Fabricated Heterostructures toward Enhanced Photocatalytic Hydrogen Evolution. J. Colloid Interface Sci. 2020, 572, 22–30. [Google Scholar] [CrossRef]
  51. Rao, Z.; Shi, G.; Wang, Z.; Mahmood, A.; Xie, X.; Sun, J. Photocatalytic Degradation of Gaseous Vocs over Tm3+-TiO2: Revealing the Activity Enhancement Mechanism and Different Reaction Paths. Chem. Eng. J. 2020, 395, 125078. [Google Scholar] [CrossRef]
  52. Feizpoor, S.; Habibi-Yangjeh, A.; Seifzadeh, D.; Ghosh, S. Combining Carbon Dots and Ag6Si2O7 Nanoparticles with TiO2: Visible-Light-Driven Photocatalysts with Efficient Performance for Removal of Pollutants. Sep. Purif. Technol. 2020, 248, 116928. [Google Scholar] [CrossRef]
  53. Porcar-Santos, O.; Cruz-Alcalde, A.; Lopez-Vinent, N.; Zanganas, D.; Sans, C. Photocatalytic Degradation of Sulfamethoxazole Using TiO2 in Simulated Seawater: Evidence for Direct Formation of Reactive Halogen Species and Halogenated by-Products. Sci. Total Environ. 2020, 736, 139605. [Google Scholar] [CrossRef] [PubMed]
  54. Li, K.; Teng, C.; Wang, S.; Min, Q. Recent Advances in TiO2-Based Heterojunctions for Photocatalytic CO2 Reduction with Water Oxidation: A Review. Front. Chem. 2021, 9, 637501. [Google Scholar] [CrossRef]
  55. Nguyen, T.P.; Nguyen DL, T.; Nguyen, V.H.; Le, T.H.; Vo DV, N.; Trinh, Q.T.; Bae, S.-R.; Chae, S.Y.; Kim, S.Y.; Le, Q.V. Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels. Nanomaterials 2020, 10, 337. [Google Scholar] [CrossRef] [Green Version]
  56. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A Critical Review on TiO2 Based Photocatalytic CO2 Reduction System: Strategies to Improve Efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  57. Tahir, M.; Amin, N.S. Advances in Visible Light Responsive Titanium Oxide-Based Photocatalysts for CO2 Conversion to Hydrocarbon Fuels. Energy Convers. Manag. 2013, 76, 194–214. [Google Scholar] [CrossRef]
  58. Wu, L.; Fu, C.; Huang, W. Surface Chemistry of TiO2 Connecting Thermal Catalysis and Photocatalysis. Phys. Chem. Chem. Phys. 2020, 22, 9875–9909. [Google Scholar] [CrossRef]
  59. Tang, J.; Chen, Y.; Dong, Z. Effect of Crystalline Structure on Terbuthylazine Degradation by H2O2-Assisted TiO2 Photocatalysis under Visible Irradiation. J. Environ. Sci. 2019, 79, 153–160. [Google Scholar] [CrossRef]
  60. Jin, J.; Chen, S.; Wang, J.; Chen, C.; Peng, T. One-Pot Hydrothermal Preparation of PbO-Decorated Brookite/Anatase TiO2 Composites with Remarkably Enhanced CO2 Photoreduction Activity. Appl. Catal. B Environ. 2020, 263, 118353. [Google Scholar] [CrossRef]
  61. Cao, M.-Q.; Liu, K.; Zhou, H.-M.; Li, H.-M.; Gao, X.-H.; Qiu, X.-Q.; Liu, M. Hierarchical TiO2 Nanorods with a Highly Active Surface for Photocatalytic CO2 Reduction. J. Cent. South Univ. 2019, 26, 1503–1509. [Google Scholar] [CrossRef]
  62. Tan, D.; Zhang, J.; Shi, J.; Li, S.; Zhang, B.; Tan, X.; Zhang, F.; Liu, L.; Shao, D.; Han, B. Photocatalytic CO2 Transformation to CH4 by Ag/Pd Bimetals Supported on N-Doped TiO2 Nanosheet. ACS Appl. Mater. Interfaces 2018, 10, 24516–24522. [Google Scholar] [CrossRef] [PubMed]
  63. Kar, P.; Zeng, S.; Zhang, Y.; Vahidzadeh, E.; Manuel, A.; Kisslinger, R.; Alam, K.M.; Thakur, U.K.; Mahdi, N.; Kumar, P.; et al. High Rate CO2 Photoreduction Using Flame Annealed TiO2 Nanotubes. Appl. Catal. B Environ. 2019, 243, 522–536. [Google Scholar] [CrossRef]
  64. Yang, L.; Yu, L.E.; Ray, M.B. Degradation of Paracetamol in Aqueous Solutions by TiO2 Photocatalysis. Water Res. 2008, 42, 3480–3488. [Google Scholar] [CrossRef] [PubMed]
  65. Shimoda, T.; Morishima, T.; Kodama, K.; Hirose, T.; Polyansky, D.E.; Manbeck, G.F.; Muckerman, J.T.; Fujita, E. Photocatalytic CO2 Reduction by Trigonal-Bipyramidal Cobalt(II) Polypyridyl Complexes: The Nature of Cobalt(I) and Cobalt(0) Complexes Upon Their Reactions with CO2, CO, or Proton. Inorg. Chem. 2018, 57, 5486–5498. [Google Scholar] [CrossRef]
  66. Wang, Y.; Wang, S.; Lou, X.W. Dispersed Nickel Cobalt Oxyphosphide Nanoparticles Confined in Multichannel Hollow Carbon Fibers for Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 17236–17240. [Google Scholar] [CrossRef]
  67. Xu, J.; Liu, X.; Zhou, Z.; Xu, M. Photocatalytic CO2 Reduction Catalyzed by Metalloporphyrin: Understanding of Cobalt and Nickel Sites in Activity and Adsorption. Appl. Surf. Sci. 2020, 513, 145801. [Google Scholar] [CrossRef]
  68. Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Cobalt Imidazolate Metal-Organic Frameworks Photosplit CO2 under Mild Reaction Conditions. Angew. Chem. Int. Ed. 2014, 53, 1034–1038. [Google Scholar] [CrossRef]
  69. Wang, S.; Ding, Z.; Wang, X. A Stable ZnCo2O4 Cocatalyst for Photocatalytic CO2 Reduction. Chem. Commun. 2015, 51, 1517–1519. [Google Scholar] [CrossRef]
  70. Wang, S.; Hou, Y.; Wang, X. Development of a Stable MnCo2O4 Cocatalyst for Photocatalytic CO2 Reduction with Visible Light. ACS Appl. Mater. Interfaces 2015, 7, 4327–4335. [Google Scholar] [CrossRef]
  71. Wang, Y.; Huang, N.-Y.; Shen, J.-Q.; Liao, P.-Q.; Chen, X.-M.; Zhang, J.-P. Hydroxide Ligands Cooperate with Catalytic Centers in Metal-Organic Frameworks for Efficient Photocatalytic CO2 Reduction. JACS 2018, 140, 38–41. [Google Scholar] [CrossRef] [PubMed]
  72. Niu, K.; Xu, Y.; Wang, H.; Ye, R.; Xin, H.L.; Lin, F.; Tian, C.; Lum, Y.; Bustillo, K.C.; Doeff, M.M.; et al. A Spongy Nickel-Organic CO2 Reduction Photocatalyst for Nearly 100% Selective CO Production. Sci. Adv. 2017, 3, 1700921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zhong, W.; Sa, R.; Li, L.; He, Y.; Li, L.; Bi, J.; Zhuang, Z.; Yu, Y.; Zou, Z. A Covalent Organic Framework Bearing Single Ni Sites as a Synergistic Photocatalyst for Selective Photoreduction of CO2 to CO. JACS 2019, 141, 7615–7621. [Google Scholar] [CrossRef]
  74. Chen, J.; Dong, C.; Idriss, H.; Mohammed, O.F.; Bakr, O.M. Metal Halide Perovskites for Solar-to-Chemical Fuel Conversion. Adv. Energy Mater. 2020, 10, 1902433. [Google Scholar] [CrossRef] [Green Version]
  75. Liang, S.; Zhang, M.; Biesold, G.M.; Choi, W.; He, Y.; Li, Z.; Shen, D.; Lin, Z. Recent Advances in Synthesis, Properties, and Applications of Metal Halide Perovskite Nanocrystals/Polymer Nanocomposites. Adv. Mater. 2021, 33, 2005888. [Google Scholar] [CrossRef] [PubMed]
  76. Hou, J.; Cao, S.; Wu, Y.; Gao, Z.; Liang, F.; Sun, Y.; Lin, Z.; Sun, L. Inorganic Colloidal Perovskite Quantum Dots for Robust Solar CO2 Reduction. Chem. Eur. J. 2017, 23, 9481–9485. [Google Scholar] [CrossRef] [PubMed]
  77. Xu, Y.-F.; Yang, M.-Z.; Chen, B.-X.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2 Reduction. JACS 2017, 139, 5660–5663. [Google Scholar] [CrossRef]
  78. Ou, M.; Tu, W.; Yin, S.; Xing, W.; Wu, S.; Wang, H.; Wan, S.; Zhong, Q.; Xu, R. Amino-Assisted Anchoring of CsPbBr3 Perovskite Quantum Dots on Porous g-C3N4 for Enhanced Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2018, 57, 13570–13574. [Google Scholar] [CrossRef]
  79. Jiang, Y.; Liao, J.-F.; Chen, H.-Y.; Zhang, H.-H.; Li, J.-Y.; Wang, X.-D.; Kuang, D.-B. All-Solid-State Z-Scheme Alpha-Fe2O3/Amine-RGO/CsPbBr3 Hybrids for Visible-Light-Driven Photocatalytic CO2 Reduction. Chem 2020, 6, 766–780. [Google Scholar] [CrossRef]
  80. Zhu, L.; Liu, X.-Q.; Jiang, H.-L.; Sun, L.-B. Metal-Organic Frameworks for Heterogeneous Basic Catalysis. Chem. Rev. 2017, 117, 8129–8176. [Google Scholar] [CrossRef]
  81. Li, R.; Hu, J.; Deng, M.; Wang, H.; Wang, X.; Hu, Y.; Jiang, H.-L.; Jiang, J.; Zhang, Q.; Xie, Y.; et al. Integration of an Inorganic Semiconductor with a Metal-Organic Framework: A Platform for Enhanced Gaseous Photocatalytic Reactions. Adv. Mater. 2014, 26, 4783–4788. [Google Scholar] [CrossRef] [PubMed]
  82. Wu, L.-Y.; Mu, Y.-F.; Guo, X.-X.; Zhang, W.; Zhang, Z.-M.; Zhang, M.; Lu, T.-B. Encapsulating Perovskite Quantum Dots in Iron-Based Metal-Organic Frameworks (MOFs) for Efficient Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 9491–9495. [Google Scholar] [CrossRef] [PubMed]
  83. Alvaro, M.; Carbonell, E.; Ferrer, B.; Llabres i Xamena, F.X.; Garcia, H. Semiconductor Behavior of a Metal-Organic Framework (MOF). Chem. Eur. J. 2007, 13, 5106–5112. [Google Scholar] [CrossRef] [PubMed]
  84. Ma, X.; Wang, L.; Zhang, Q.; Jiang, H.-L. Switching on the Photocatalysis of Metal-Organic Frameworks by Engineering Structural Defects. Angew. Chem. Int. Ed. 2019, 58, 12175–12179. [Google Scholar] [CrossRef]
  85. Fang, Z.-B.; Liu, T.-T.; Liu, J.; Jin, S.; Wu, X.-P.; Gong, X.-Q.; Wang, K.; Yin, Q.; Liu, T.-F.; Cao, R.; et al. Boosting Interfacial Charge-Transfer Kinetics for Efficient Overall CO2 Photoreduction Via Rational Design of Coordination Spheres on Metal-Organic Frameworks. JACS 2020, 142, 12515–12523. [Google Scholar] [CrossRef]
  86. Han, C.; Lei, Y.; Wang, B.; Wu, C.; Zhang, X.; Shen, S.; Sun, L.; Tian, Q.; Feng, Q.; Wang, Y. The Functionality of Surface Hydroxyls on Selective Ch4 Generation from Photoreduction of CO2 over Sic Nanosheets. Chem. Commun. 2019, 55, 1572–1575. [Google Scholar] [CrossRef]
  87. Yang, T.C.; Chang, F.C.; Peng, C.Y.; Wang, H.P.; Wei, Y.L. Photocatalytic Reduction of CO2 with Sic Recovered from Silicon Sludge Wastes. Environ. Technol. 2015, 36, 2987–2990. [Google Scholar] [CrossRef]
  88. Ahmed, N.; Shibata, Y.; Taniguchi, T.; Izumi, Y. Photocatalytic Conversion of Carbon Dioxide into Methanol Using Zinc-Copper-M(III) (M = Aluminum, Gallium) Layered Double Hydroxides. J. Catal. 2011, 279, 123–135. [Google Scholar] [CrossRef]
  89. Gao, G.; Zhu, Z.; Zheng, J.; Liu, Z.; Wang, Q.; Yan, Y. Ultrathin Magnetic Mg-Al Ldh Photocatalyst for Enhanced CO2 Reduction: Fabrication and Mechanism. J. Colloid Interface Sci. 2019, 555, 1–10. [Google Scholar] [CrossRef]
  90. Kawamura, S.; Ahmed, N.; Carja, G.; Izumi, Y. Photocatalytic Conversion of Carbon Dioxide Using Zn-Cu-Ga Layered Double Hydroxides Assembled with Cu Phthalocyanine: Cu in Contact with Gaseous Reactant Is Needed for Methanol Generation. Oil Gas Sci. Technol. 2015, 70, 841–852. [Google Scholar] [CrossRef]
  91. Teramura, K.; Iguchi, S.; Mizuno, Y.; Shishido, T.; Tanaka, T. Photocatalytic Conversion of CO2 in Water over Layered Double Hydroxides. Angew. Chem. Int. Ed. 2012, 51, 8008–8011. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, Y.; Wang, X.; Antonietti, M. Polymeric Graphitic Carbon Nitride as a Heterogeneous Organocatalyst: From Photochemistry to Multipurpose Catalysis to Sustainable Chemistry. Angew. Chem. Int. Ed. 2012, 51, 68–89. [Google Scholar] [CrossRef] [PubMed]
  93. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-Like Carbon Nitride Nanosheets for Improved Photocatalytic Activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  94. Hsu, H.-C.; Shown, I.; Wei, H.-Y.; Chang, Y.-C.; Du, H.-Y.; Lin, Y.-G.; Tseng, C.-A.; Wang, C.-H.; Chen, L.-C.; Lin, Y.-C.; et al. Graphene Oxide as a Promising Photocatalyst for CO2 to Methanol Conversion. Nanoscale 2013, 5, 262–268. [Google Scholar] [CrossRef] [PubMed]
  95. Park, H.-A.; Choi, J.H.; Choi, K.M.; Lee, D.K.; Kang, J.K. Highly Porous Gallium Oxide with a High CO2 Affinity for the Photocatalytic Conversion of Carbon Dioxide into Methane. J. Mater. Chem. 2012, 22, 5304–5307. [Google Scholar] [CrossRef]
  96. Tsuneoka, H.; Teramura, K.; Shishido, T.; Tanaka, T. Adsorbed Species of CO2 and H2 on Ga2O3 for the Photocatalytic Reduction of CO2. J. Phys. Chem. C 2010, 114, 8892–8898. [Google Scholar] [CrossRef]
  97. Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.; Zou, Z. High-Yield Synthesis of Ultrathin and Uniform Bi2WO6 Square Nanoplates Benefitting from Photocatalytic Reduction of CO2 into Renewable Hydrocarbon Fuel under Visible Light. ACS Appl. Mater. Interfaces 2011, 3, 3594–3601. [Google Scholar] [CrossRef]
  98. Cheng, H.; Huang, B.; Liu, Y.; Wang, Z.; Qin, X.; Zhang, X.; Dai, Y. An Anion Exchange Approach to Bi2WO6 Hollow Microspheres with Efficient Visible Light Photocatalytic Reduction of CO2 to Methanol. Chem. Commun. 2012, 48, 9729–9731. [Google Scholar] [CrossRef]
  99. Xi, G.; Ouyang, S.; Li, P.; Ye, J.; Ma, Q.; Su, N.; Bai, H.; Wang, C. Ultrathin W18O49 Nanowires with Diameters Below 1 nm: Synthesis, near-Infrared Absorption, Photoluminescence, and Photochemical Reduction of Carbon Dioxide. Angew. Chem. Int. Ed. 2012, 51, 2395–2399. [Google Scholar] [CrossRef]
  100. Li, K.; Peng, B.; Peng, T. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. Acs Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  101. Zhang, Q.-H.; Han, W.-D.; Hong, Y.-J.; Yu, J.-G. Photocatalytic Reduction of CO2 with H2O on Pt-Loaded TiO2 Catalyst. Catal. Today 2009, 148, 335–340. [Google Scholar] [CrossRef]
  102. Mao, J.; Li, K.; Peng, T. Recent Advances in the Photocatalytic CO2 Reduction over Semiconductors. Catal. Sci. Technol. 2013, 3, 2481–2498. [Google Scholar] [CrossRef]
  103. Mori, K.; Yamashita, H.; Anpo, M. Photocatalytic Reduction of CO2 with H2O on Various Titanium Oxide Photocatalysts. RSC Adv. 2012, 2, 3165–3172. [Google Scholar] [CrossRef]
  104. Kaneco, S.; Shimizu, Y.; Ohta, K.; Mizuno, T. Photocatalytic Reduction of High Pressure Carbon Dioxide Using TiO2 Powders with a Positive Hole Scavenger. J. Photochem. Photobiol. A 1998, 115, 223–226. [Google Scholar] [CrossRef]
  105. Li, H.; Gan, S.; Wang, H.; Han, D.; Niu, L. Intercorrelated Superhybrid of Agbr Supported on Graphitic-C3N4-Decorated Nitrogen-Doped Graphene: High Engineering Photocatalytic Activities for Water Purification and CO2 Reduction. Adv. Mater. 2015, 27, 6906–6913. [Google Scholar] [CrossRef] [PubMed]
  106. Tseng, I.H.; Chang, W.C.; Wu, J. Photoreduction of CO2 Using Sol–Gel Derived Titania and Titania-Supported Copper Catalysts. Appl. Catal. B Environ. 2002, 37, 37–48. [Google Scholar] [CrossRef]
  107. Hou, W.; Hung, W.H.; Pavaskar, P.; Goeppert, A.; Aykol, M.; Cronin, S.B. Photocatalytic Conversion of CO2 to Hydrocarbon Fuels Via Plasmon-Enhanced Absorption and Metallic Interband Transitions. ACS Catal. 2011, 1, 929–936. [Google Scholar] [CrossRef]
  108. Mizuno, T.; Adachi, K.; Ohta, K.; Saji, A. Effect of CO2 Pressure on Photocatalytic Reduction of CO2 Using TiO2 in Aqueous Solutions. J. Photochem. Photobiol. A 1996, 98, 87–90. [Google Scholar] [CrossRef]
  109. Fox, M.A.; Dulay, M.T. Heterogeneous Photocatalysis. Chem. Rev. 1993, 93, 341–357. [Google Scholar] [CrossRef]
  110. Guan, G.; Kida, T.; Yoshida, A. Reduction of Carbon Dioxide with Water under Concentrated Sunlight Using Photocatalyst Combined with Fe-Based Catalyst. Appl. Catal. B Environ. 2003, 41, 387–396. [Google Scholar] [CrossRef]
  111. Ikeue, K.; Yamashita, H.; Anpo, M.; Takewaki, T. Photocatalytic Reduction of CO2 with H2O on Ti-β Zeolite Photocatalysts: Effect of the Hydrophobic and Hydrophilic Properties. J. Phys. Chem. B 2001, 105, 8350–8355. [Google Scholar] [CrossRef]
Figure 1. Band structures of some representative photocatalysts. Reprinted with permission from Ref. [46]. 2020 Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J.
Figure 1. Band structures of some representative photocatalysts. Reprinted with permission from Ref. [46]. 2020 Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J.
Catalysts 12 00372 g001
Figure 2. Photocatalytic CO/CH4 production activities of the single PbO, pristine TiO2 and PbO/TiO2 composites. Reprinted with permission from Ref. [60]. 2020 Jin, J.; Chen, S.; Wang, J.; Chen, C.; Peng, T.
Figure 2. Photocatalytic CO/CH4 production activities of the single PbO, pristine TiO2 and PbO/TiO2 composites. Reprinted with permission from Ref. [60]. 2020 Jin, J.; Chen, S.; Wang, J.; Chen, C.; Peng, T.
Catalysts 12 00372 g002
Figure 3. TOF values for the photocatalytic reduction of CO2 to CO at 1.0 atm (blue) and 0.1 atm (orange). Reprinted with permission from Ref. [71]. 2018 Wang, Y.; Huang, N.-Y.; Shen, J.-Q.; Liao, P.-Q.; Chen, X.-M.; Zhang, J.-P.
Figure 3. TOF values for the photocatalytic reduction of CO2 to CO at 1.0 atm (blue) and 0.1 atm (orange). Reprinted with permission from Ref. [71]. 2018 Wang, Y.; Huang, N.-Y.; Shen, J.-Q.; Liao, P.-Q.; Chen, X.-M.; Zhang, J.-P.
Catalysts 12 00372 g003
Figure 4. (a) Comparison of the free energy of CO2 reduction to CO with and without TpBpy ligand; (b) proposed mechanism for the photocatalytic CO2 reduction by Ni-TpBpy. Reprinted with permission from Ref. [73]. 2019 Zhong, W.; Sa, R.; Li, L.; He, Y.; Li, L.; Bi, J.; Zhuang, Z.; Yu, Y.; Zou, Z.
Figure 4. (a) Comparison of the free energy of CO2 reduction to CO with and without TpBpy ligand; (b) proposed mechanism for the photocatalytic CO2 reduction by Ni-TpBpy. Reprinted with permission from Ref. [73]. 2019 Zhong, W.; Sa, R.; Li, L.; He, Y.; Li, L.; Bi, J.; Zhuang, Z.; Yu, Y.; Zou, Z.
Catalysts 12 00372 g004
Figure 5. Schematic of photocatalytic CO2 reduction over CsPbBr3QDs/GO. Reprinted with permission from Ref. [77]. 2017 Xu, Y.-F.; Yang, M.-Z.; Chen, B.-X.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y.
Figure 5. Schematic of photocatalytic CO2 reduction over CsPbBr3QDs/GO. Reprinted with permission from Ref. [77]. 2017 Xu, Y.-F.; Yang, M.-Z.; Chen, B.-X.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y.
Catalysts 12 00372 g005
Figure 6. Proposed mechanism for CO2 photoreduction by PCN-601 with (a) charge separation and transfer through a coordination sphere; (b) reaction pathways for the products. Reprinted with permission from Ref. [85]. 2020 Fang, Z.-B.; Liu, T.-T.; Liu, J.; Jin, S.; Wu, X.-P.; Gong, X.-Q.; Wang, K.; Yin, Q.; Liu, T.-F.; Cao, R.
Figure 6. Proposed mechanism for CO2 photoreduction by PCN-601 with (a) charge separation and transfer through a coordination sphere; (b) reaction pathways for the products. Reprinted with permission from Ref. [85]. 2020 Fang, Z.-B.; Liu, T.-T.; Liu, J.; Jin, S.; Wu, X.-P.; Gong, X.-Q.; Wang, K.; Yin, Q.; Liu, T.-F.; Cao, R.
Catalysts 12 00372 g006
Figure 7. CH4 yields on Bi2WO6 photocatalysts. Reprinted with permission from Ref. [97]. 2011 Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.; Zou, Z.
Figure 7. CH4 yields on Bi2WO6 photocatalysts. Reprinted with permission from Ref. [97]. 2011 Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.; Zou, Z.
Catalysts 12 00372 g007
Figure 8. Schematic of the photocatalytic reduction of CO2 by semiconductor materials. Reprinted with permission from Ref. [100]. 2016 Li, K.; Peng, B.; Peng, T.
Figure 8. Schematic of the photocatalytic reduction of CO2 by semiconductor materials. Reprinted with permission from Ref. [100]. 2016 Li, K.; Peng, B.; Peng, T.
Catalysts 12 00372 g008
Figure 9. Schematics of gas-solid (a) and liquid-solid (b) systems for photocatalytic CO2 reduction.
Figure 9. Schematics of gas-solid (a) and liquid-solid (b) systems for photocatalytic CO2 reduction.
Catalysts 12 00372 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cui, Y.; Ge, P.; Chen, M.; Xu, L. Research Progress in Semiconductor Materials with Application in the Photocatalytic Reduction of CO2. Catalysts 2022, 12, 372. https://doi.org/10.3390/catal12040372

AMA Style

Cui Y, Ge P, Chen M, Xu L. Research Progress in Semiconductor Materials with Application in the Photocatalytic Reduction of CO2. Catalysts. 2022; 12(4):372. https://doi.org/10.3390/catal12040372

Chicago/Turabian Style

Cui, Yan, Pengxiang Ge, Mindong Chen, and Leilei Xu. 2022. "Research Progress in Semiconductor Materials with Application in the Photocatalytic Reduction of CO2" Catalysts 12, no. 4: 372. https://doi.org/10.3390/catal12040372

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