Next Article in Journal
Ru-Controlled Thymine Tautomerization Frozen by a k1(O)-, k2(N,O)-Metallacycle: An Experimental and Theoretical Approach
Next Article in Special Issue
Hydrothermally Synthesized ZnCr- and NiCr-Layered Double Hydroxides as Hydrogen Evolution Photocatalysts
Previous Article in Journal
Fatty Acid-Rich Extract from Holothuria atra for Hyperuricemia via Expressions Modulation of GLUT9a and GLUT9b in Rat Model
Previous Article in Special Issue
S-Scheme 2D/2D Heterojunction of ZnTiO3 Nanosheets/Bi2WO6 Nanosheets with Enhanced Photoelectrocatalytic Activity for Phenol Wastewater under Visible Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in the Application of Bi-Based Compounds in Photocatalytic Reduction of CO2

College of Chemistry & Chemical and Environmental Engineering, Weifang University, Weifang 261061, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(10), 3982; https://doi.org/10.3390/molecules28103982
Submission received: 15 April 2023 / Revised: 5 May 2023 / Accepted: 8 May 2023 / Published: 9 May 2023

Abstract

:
Bi-based semiconductor materials have special layered structure and appropriate band gap, which endow them with excellent visible light response ability and stable photochemical characteristics. As a new type of environment-friendly photocatalyst, they have received extensive attention in the fields of environmental remediation and energy crisis resolution and have become a research hotspot in recent years. However, there are still some urgent issues that need to be addressed in the practical large-scale application of Bi-based photocatalysts, such as the high recombination rate of photogenerated carriers, limited response range to visible spectra, poor photocatalytic activity, and weak reduction ability. In this paper, the reaction conditions and mechanism of photocatalytic reduction of CO2 and the typical characteristics of Bi-based semiconductor materials are introduced. On this basis, the research progress and application results of Bi-based photocatalysts in the field of reducing CO2, including vacancy introduction, morphological control, heterojunction construction, and co-catalyst loading, are emphasized. Finally, the future prospects of Bi-based photocatalysts are prospected, and it is pointed out that future research directions should be focused on improving the selectivity and stability of catalysts, deeply exploring reaction mechanisms, and meeting industrial production requirements.

1. Introduction

With the rapid development of the economy, a large amount of fossil fuels are consumed. It leads to the release of a large amount of carbon dioxide, causing the greenhouse effect and thus breaking the world’s ecological balance. Additionally. fossil fuels are also non-renewable energy [1]. The concentration of carbon dioxide increases is seriously impacting human habitats and the earth’s ecosystem. Therefore, it has become one of the most important research topics in the world to explore how to reduce the amount of CO2 in the atmosphere and use it rationally. Photocatalysis technology offers an excellent solution for the conversion of CO2 into important chemical fuels such as CH4 [2]. At the beginning of photocatalysis research, researchers used TiO2 electrodes to decompose water under visible light to produce hydrogen, thus triggering a great deal of interest in the direction of photocatalysis [3]. Therefore, it is essential to use photocatalysis to reduce CO2 in the atmosphere and collect and store solar energy in chemical fuels. Currently, the catalysts used in photocatalytic CO2 reduction, such as noble metals (Pt and Au) [3,4], non-precious metals (Cu, Fe, Ni, and g-C3N4) [5,6,7,8], metal oxides (e.g., TiO2 and Ga2O3) [3,9,10], metal sulfides (e.g., CdS and MoS2) [11,12], and graphene [1,13,14] have been discovered. The successful conversion of CO2 into CO, CH3OH, CH4, and other available chemicals has been reported in the literature [15,16,17,18,19]. However, the photocatalytic efficiency of most semiconductors is relatively low and not can be applied industrially. The main reasons are as follows [18,20]: (1) the separation and migration efficiency of photo-generated carriers is low, and most of the photo-generated electrons and holes will be recombined in the process of migration to the semiconductor surface, which dramatically reduces the photocatalytic performance; (2) the reaction active site is insufficient, and due to the low specific surface area, only a tiny amount of photo-generated carriers can reach the catalyst surface to react with a small amount of adsorbed CO2 on the surface; (3) in general, the catalytic activity of precious metals (such as Pt, Ag) is higher than that of non-precious or non-metallic metals, while the cost is high. This hinders the industrialization of photocatalysis. In this respect, the rational design of an efficient photocatalyst is necessary for the practical application of photocatalytic CO2 reduction.
Bi-based catalysts are widely used for photocatalytic CO2 reduction with the advantages of low price, excellent photoelectric activity, and environmental friendliness. For example, bismuth halide oxide (BiOX, X = Cl, Br, I) has been extensively studied in photocatalysis due to its excellent photocatalytic properties. The layered structure of BiOX provides enough space to polarize the associated atoms and orbitals, which excites the formation of an internal electric field between [Bi2O2] and halogen. The internal electric field could accelerate the separation and migration of photoexcited electron-hole pairs, and the photocatalytic activity of BiOX is significantly improved [21,22,23,24,25]. To improve the activity of Bi-based catalysts for photocatalytic reduction of CO2, researchers have modified the catalyst morphology, heterojunction structure, loaded co-catalysts, and introduced vacancies. Among them, introducing vacancies on the catalyst surface is an effective way to improve photocatalytic efficiency. Vacancies could effectively change the charge distribution and electron energy level of the catalyst metal elements, thus creating a synergistic effect between the elements, promoting photo-generated charge separation, and generating enough photo-generated electrons for the photocatalytic reduction of CO2 [26,27,28,29,30].
To our knowledge, although the evaluation of Bi-based materials continues to emerge, there are few review articles on Bi-based materials applied in photocatalytic CO2 reduction. Therefore, this paper systematically summarizes the recent progress of catalysts used for photocatalytic CO2 reduction: firstly, the basic principles of photocatalytic CO2 reduction are briefly introduced; secondly, the preparation methods of catalysts are introduced, focusing on the summary of the improvement strategies of photocatalytic CO2 reduction performance; then, several bismuth-based photocatalysts for CO2 reduction are introduced. Finally, the prospects and challenges of photocatalysts in CO2 reduction are presented.

2. Photocatalytic CO2 Reduction Foundation

2.1. Basic Principles of Photocatalysis

With the massive consumption of oil and fossil energy sources, CO2 emissions continue to rise. It has caused severe harm to the future human living environment and the earth’s ecosystem. For this reason, it has become the focus of current research to explore how to effectively reduce the amount of CO2 in the atmosphere and utilize it wisely. On the other hand, The resource utilization of CO2 requires the activation of CO2 molecules. In contrast, the chemical properties of CO2 were highly stable. The conventional CO2 resource utilization technology requires very high energy input to activate and convert it, which has the disadvantages of high economic cost, high energy consumption, long conversion period, and harsh reaction conditions [31]. Therefore, exploring an economical, practical, and simple CO2 conversion method to alleviate carbon emissions and energy shortage is strategically essential. Photocatalytic CO2 reduction is a relatively clean, economical, and efficient method.
With inexhaustible sunlight as the only input source and the use of suitable catalysts, photocatalytic CO2 reduction is possible. In the photocatalytic process (Figure 1), the absorbed photons could stimulate semiconductor (SC) photocatalysts to produce electron-hole pairs. The internal electrons will be transferred from the valence band (VB) to the conduction band (CB) inside the semiconductor after absorbing photon energy. Negatively charged electrons will be produced at the CB, and positively charged holes will be produced at the VB. Subsequently, the light-induced e-h+ will separate and migrate to the surface of the photocatalyst [32,33,34,35,36].

2.2. Mechanism of Photocatalytic CO2 Reduction

Thermodynamically, CO2 is one of the most stable linear molecules. The carbon atom in CO2 shows the highest oxidation state; the corresponding C = O bond energy is up to 750 kJ·mol−1; and the gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) is large (13.7 eV). The Gibbs free energy △G0 is −394.4 kJ·mol−1 [37]. Therefore, activating CO2 molecules requires a high input of energy. In addition, due to the optical inertness of CO2 in the UV visible spectral range, the photocatalytic reduction of CO2 requires the involvement of catalysts.
The photocatalytic reduction of CO2 is a process that mimics the photosynthesis of plants to convert CO2 into a series of hydrocarbons such as methanol, methane, and derivatives. To reduce dependence on oil and fossil fuels while reducing atmospheric CO2 concentrations, photocatalytic CO2 reduction is an excellent clean and environmentally friendly method. Photocatalytic CO2 reduction goes through the following steps (Figure 2): (1) CO2 molecules are chemically adsorbed on the surface of the photocatalyst; (2) the electrons of the semiconductor photocatalyst are activated by light with a specific wavelength and migrate from VB to CB, leaving photo-generated holes (h+) in VB; (3) photo-generated electrons (e) and h+ migrate to the catalyst surface, respectively; (4) e are used to activate CO2 and reduce it to solar fuel, while h+ involved in the oxidation of H2O is consumed [38]. Therefore, from the perspective of photocatalytic reaction steps, an efficient photocatalyst should have the following characteristics: (1) a large specific surface area to increase the active surface sites and improve the adsorption capacity of CO2; (2) a narrow bandgap and appropriate band position improve solar energy utilization efficiency; (3) a nanostructure that promotes electron migration and photo-generated carrier separation is necessary; (4) rich surface defects (such as vacancies) alter the electronic and chemical properties of semiconductor, promoting the adsorption and activation of CO2. In addition, co-catalysts are usually added to photocatalysts to promote the separation and migration of photo-generated charge carriers and effectively reduce the reaction energy barrier for CO2 activation and reduction [39,40].
The photocatalytic CO2 reduction reaction is a proton-coupled multi-electron transfer process, and the reduction products vary with the number of electrons participating in the reaction. Two to eight electrons and corresponding numbers of protons participate in the reaction to generate HCOOH, CO, HCHO, CH3OH, and CH4, respectively. The different reduction products and corresponding electrode potentials obtained from the photocatalytic reduction of CO2 in an aqueous solution are shown in Table 1 [41]. Therefore, CO2 can be reduced to hydrocarbon fuel only when the semiconductor’s CB and VB positions meet the photocatalytic reaction’s thermodynamic requirements. Specifically, the CB position must be more negative than the CO2 reduction potential to transfer electrons from the semiconductor to the surface-adsorbed CO2. Furthermore, VB position is more positive than the H2O oxidation reaction. Only in this way can the reaction process of photocatalytic reduction of CO2 be achieved.

2.3. Bi-Based Catalysts

Bismuth-based catalysts have excellent electronic structure and unique redox properties. They could significantly improve the photocatalytic efficiency and the yield of products. Experience shows that the width of the band gap will affect the migration and recombination of electrons. Catalysts with wide band gaps can affect the absorption of light energy with longer wavelengths and have low utilization rates of light energy. Narrow band gap catalysts increase the probability of electron migration and recombination, reducing photocatalytic efficiency [42,43,44,45]. The band gap structures of common bismuth-based catalysts are shown in Figure 3. Most bismuth-based catalysts have a band gap of less than 3.0 eV, which means that they could be excited by solar irradiation [46,47,48]. Meanwhile, since the CB values of most bismuth-based catalysts are not negative enough, the photoinduced electrons do not have enough reduction capacity to initiate the photocatalytic reaction. The band gap of most bismuth-based catalysts satisfies the band potential energy, and the construction of heterostructures with other semiconductors could promote photo-generated electron transfer and accumulate enough electrons to participate in the photocatalytic reduction of CO2. Therefore, constructing intercalation structures between bismuth-based and other catalysts is a feasible way to improve photocatalytic performance [49,50].

3. Optimization Strategy of Bi-Based Photocatalyst

Photocatalytic CO2 reduction promotes the production of energy-rich hydrocarbon fuels (CO, CH4, and CH3OH) and alleviates environmental problems. Several studies have shown that atmospheric CO2 will increase dramatically to 550 PPM by 2100. The development of the effective surface area and potential photocatalysts on exposed crystal surfaces effectively prevents the adsorption, activation, and reduction of CO2 molecules.

3.1. Introduction of Vacancies

Vacancies or defects play a crucial role in photocatalytic applications. They could effectively change the charge distribution and electron energy level of catalysts, improving the catalyst activity. In particular, vacancies introduced in the z-structure contribute to effective interfacial interactions, thus creating a synergistic effect between elements, promoting photo-generated charge separation, and generating sufficient photo-generated electrons for the photocatalytic reduction of CO2. The modification effect of introducing vacancies in semiconductor catalysts is generally reflected in three aspects [51,52]: (1) the generated vacancies can act as trapping centers for photo-generated electrons or holes and inhibit photo-generated charge complexes; (2) they act as active centers and promote the adsorption and activation of reactant molecules; (3) they adjust the band gap energy of catalysts and enhance the light absorption capacity.
Researchers assembled ultra-thin HNb3O8 nanosheets and BiOBr nanosheets to prepare BiOBr/HNb3O8 Z-scheme heterojunctions with abundant oxygen vacancies [53]. The photocatalytic CO2 reduction performance of the photocatalyst was greatly improved, which was attributed to various synergies: (1) the 2D–3D structure produced a more active site; (2) abundant oxygen vacancies enhanced the light absorption ability; (3) the combination of Z-scheme heterostructures and oxygen vacancies promoted the separation and transfer of photogenerated charge carriers. The best BiOBr VO/HNB3O8 NS (50%-BiOBr VO/HNB3O8 NS) photocatalyst showed a CO yield of up to 164.61 μmol g−1 and a high selectivity of 98.7%. In addition, after five cycles, the catalyst exhibited excellent stability, indicating that the number of induced oxygen vacancies remained relatively balanced during the photocatalytic reaction due to the inhibitory effect of the Z-scheme heterostructure on the oxygen vacancy reduction process. Yang et al. synthesized Bi2MoO6 (BMO) with exposed {001} crystal planes by a simple solvothermal method [54]. Compared with Bi2MoO6, the yields of CO and CH4 on Bi2MoO6 containing oxygen vacancies (BMO OVs) reached 0.27 and 2.01 μmol g−1 h−1, respectively. Additionally, the corresponding CH4 selectivity was up to 96.7%. On this basis, the team explored the CO2 adsorption mode and reaction process, and for the first time revealed the hidden mechanism of CO2 selective conversion on oxygen vacancies. The calculation result of charge density difference showed that 0.15 electrons were returned from the slab to B1−CO2 after CO2 adsorption on the surface of BMO OVs, which depleted the local electrons of the two Bi atoms around the oxygen vacancy, and more electrons accumulated on B1−CO2 (Figure 4). In contrast, 0.07 electrons are transferred from CO2 to the slab in BMO. In addition, the C−O bond lengths of CO2 adsorbed on BMO OVs and BMO increase to 1.26 Å or 1.29 Å and 1.26 Å or 1.30 Å, respectively. In an overall analysis, the OVs on the surface of Bi2MoO6 {001} are more favorable to the chemisorption of CO2.
Vacancies on the catalyst surface are rich in localized electrons, which can transfer electrons to the adsorbed gas molecules on the surface and form coordination bonds between nearby metal atoms and reactants, activating the reactants and driving the photocatalytic reaction to proceed. Although the vacancy design can improve the CO2 reduction activity of catalysts, the long-term efficient stabilization of vacancy defects remains a challenge.

3.2. Morphological Control

There is a constitutive relationship between nanomaterials’ size, morphological structure, and properties. Particle size affects the band gap energy, light absorption capacity, and the material’s average free range of photo-generated charges. When the particle size of the material is smaller than the thickness of its space charge layer, the space charge layer is negligible, and the photo-generated charge migrates from the bulk phase to the material surface through simple diffusion and participates in the surface redox reaction. Photo-generated charge transfer and separation efficiency is improved by shortening the charge migration distance [55]. On the other hand, the smaller particle size with a high specific surface area and excellent adsorption properties can promote the interaction between catalysts and reactants while providing abundant active sites and a wide light absorption area for the redox reaction, enhancing photocatalytic performance [56]. The researchers prepared AgCl/δ-Bi2O3 nanosheets with a thickness of about 2.7 nm by the hydrothermal precipitation method [57]. The structural properties of the catalysts, including morphology, crystallinity, optical properties, and energy band structure, were analyzed by SEM, TEM, and AFM tests. The characterization results show that the ultrathin two-dimensional nanomaterials have a larger specific surface area and two-dimensional anisotropy, exposing more surface unsaturated paired atoms, increasing the active surface centers, and shortening the charge migration distance to achieve effective separation of electron-hole pairs. At the same time, surface atoms tend to escape from the lattice during the size reduction of nanomaterials, inducing the generation of surface defects. Jiang et al. [58] prepared S-Scheme 2D/2D heterojunction ZnTiO3 nanosheets/Bi2WO6 nanosheets as catalysts. The catalyst structures are shown in Figure 5.
Wu et al. [59] prepared hierarchical Z-scheme BiVO4/hm-C4N3 structures. The semiconductor material BiVO4 has a large specific surface area, which is favorable for trapping and absorbing incident photons. Due to the darker appearance of hm-C4N3, the BiVO4/hm-C4N3 composite exhibits significantly higher light absorption in the UV and visible regions. Z-scheme structure promotes the separation of e-h+ and enhances the reduction ability of hm-C4N3. BiVO4/hm-C4N3 composites achieve photocatalytic CO2 reduction to CO (48.0 μmol·g−1h−1) with a selectivity of more than 97%. Lu et al. [60] prepared 2D/2D g-C3N4/BiVO4 Z-scheme catalysts. Due to the 2D–2D structure of g-C3N4 and BiVO4 nanosheets, without obvious boundaries, thus providing abundant active sites with effectively enhanced interfacial charge migration. It was shown that the prepared catalysts achieved CO and CH4 conversions of 5.19 μmol·g−1h−1 and 4.57 μmol·g−1h−1, respectively.

3.3. Heterojunction Construction

Heterostructure refers to the interfacial region formed by the contact of two different semiconductors. The construction of heterojunction enhances the light absorption ability of catalysts, improves the quantum efficiency by the window effect, and also promotes the separation of photo-generated electrons and holes by the interfacial effect. The structure is characterized by the fact that [61,62] (1) the surfaces of both components are exposed, which gives both particles the possibility to interact with the environment; (2) both materials form nanoparticles independently, making them more functional and have dual component properties. The heterostructure is a compounding of electrons from the more positive conduction band position and holes from the more negative valence band position of two semiconductors at the heterojunction interface, respectively. The remaining photo-generated electrons in the more negative CB and the remaining photo-generated holes in the more positive VB are simultaneously retained, and they have excellent reduction and oxidation abilities. Semiconductor photocatalysts with more negative CB positions can be considered excellent reduced photocatalysts, while those with more positive VB positions can be considered better oxidation photocatalysts. The composite heterojunction of reduced and oxidized photocatalysts could take full advantage of the high reduction and oxidation ability, thus significantly enhancing the photocatalytic performance.
Z-scheme heterojunction can achieve spatial separation of electrons and holes on different semiconductor materials, with advantages such as wide spectral response, high charge separation efficiency, strong redox ability, and high stability. Z-scheme heterojunction has broad prospects in the application of photocatalytic CO2 reduction. The common Z-scheme heterojunction consists of two band matching semiconductors (I and Ⅱ) and redox medium. The CB and VB positions of semiconductor I are higher than those of semiconductor Ⅱ, and the redox medium serves as the interface charge transfer channel between the two. Under light excitation conditions, both semiconductors are excited to generate electrons and holes, which remain in their CB and VB respectively. The excited electrons in the CB of semiconductor I have strong reduction ability and can catalyze the reduction of CO2; the photogenerated holes in the VB of semiconductor Ⅱ have strong oxidation ability and can catalyze the oxidation of H2O. The electrons in the CB of semiconductor Ⅱ and the holes in the VB of semiconductor I will be transferred through the electronic medium and annihilated. Therefore, the Z-scheme heterojunction effectively inhibits the recombination of electron hole pairs and significantly improves photocatalytic performance. Bai et al. [63] reported a g-C3N4/Bi4O5I2 heterojunction photocatalyst prepared by precursor hydrolysis and applied to the visible light reduction of CO2 for fuel production. Oxygen-active species quantification experiments, confirmed that the heterojunction enhances the efficiency of photocatalytic CO2 reduction by generating an I/I3− redox mediator with a direct Z-scheme structure. Combined with the test results of ·OH, it can be inferred that the H2O oxidation position of g-C3N4/Bi4O5I2 should be the VB of g-C3N4 rather than the VB of Bi4O5I2. So the charge transfer mechanism in Figure 6a does not work. This finding further confirms the previous inference about the direct Z-scheme structure of the bismuth iodide oxygen composite (Figure 6b). Under visible light irradiation, the photogenerated electrons in the photocatalyst are transferred from the CB of g-C3N4 to the VB of Bi4O5I2 through the I3/I redox medium, promoting the separation of photogenerated electrons and holes while retaining strong redox ability. Therefore, g-C3N4/Bi4O5I2 exhibits high photocatalytic reduction performance, with a yield of 45.6 μmol h−1 g−1 for CO2 conversion to CO.
Guo et al. [64] constructed in situ S-scheme (Figure 7a–c) BiOBr/Bi2WO6 heterojunction photocatalysts with tight interfacial contacts using a one-step hydrothermal method. The unique nanoflower morphology was obtained by tuning the synthesis conditions. The material was further calcined under a nitrogen atmosphere to introduce surface oxygen vacancies. Without any sacrificial agent and co-catalyst, the heterojunction material exhibited excellent photocatalytic CO2 reduction with a product CO production rate of 55.17 μmol·g−1 h−1, higher than most reported photocatalysts. The reason is that the construction of S-scheme heterojunction greatly facilitates photo-generated charges’ separation and transfer efficiency while having an efficient redox capability. The nanoflower morphology substantially enhances the CO2 adsorption capacity of the photocatalyst due to its large specific surface area. The efficiency of photocatalytic CO2 reduction is greatly improved. Miao et al. [65] prepared a catalyst with BiOBr/Bi2S3 S-scheme structure by the hydrothermal method. Since the Fermi energy level of Bi2S3 is larger than BiOBr, and the power function is smaller than BiOBr, when the electrons of Bi2S3 come into contact with BiOBr, they could transfer electrons to BiOBr and form an internal electric field. Under light irradiation, the photo-generated holes on Bi2S3 VB are recombined with the photo-generated electrons on BiOBr CB driven by the internal electric field, and the photo-generated electrons on BiOBr CB are used for the CO2 reduction reaction. The CO and CH4 yields of the best photocatalyst were as high as 100.8 and 8.5 μmol·g−1h−1, respectively.

3.4. Co-Catalyst Loading

Photocatalytic CO2 conversion could be broadly divided into three processes [66,67,68]: (i) absorption of light by the catalyst to generate electron–hole pairs, (ii) separation and transfer of electron–hole pairs to the catalyst surface, and (iii) surface reactions of H2O oxidation and CO2 reduction. Adding a co-catalyst could enhance the process (iii) and improve the photocatalytic activity. Introducing a co-catalyst into the host lattice of a semiconductor induces defective states in the electronic and chemical structure, which in turn affects the overall performance of the catalyst. In photocatalytic CO2 reduction, the critical roles of doping sites are to act as active centers for CO2 adsorption activation and to promote photo-generated charge separation. The co-catalyst has the following three critical roles: (1) promoting the separation of photoexcited electron–hole pairs, (2) inhibiting side reactions, and (3) improving the selectivity of the target product. Jiang et al. [69] loaded Pt nanoparticles as co-catalysts onto the surface of the prepared CsPbBr3/Bi2WO6 catalyst. The ultrathin nanosheet structure gave the catalyst a short charge transfer distance, thus facilitating charge separation. In addition, Bi2WO6 NSs provide additional electron transfer channels, which effectively inhibit charge recombination. It was shown that the selectivity of CO and CH4 accounted for 11.4% and 84.3%, respectively. The activity of the loaded catalysts was increased by 2.4 times compared with those prepared without the Pt nanoparticle co-catalysts.
In addition to the noble metal Pt, other noble metals (such as Au and Ag) could be loaded to enhance the photocatalytic CO2 reduction activity. Precious metal co-catalysts have a selective effect on the reactant intermediates, which in turn affects the selectivity of the products. Yoshino et al. [70] loaded Cu, Ag, Ru, Au, and Pt onto BiVO4 to photocatalytically reduce CO2. The results showed that the catalysts prepared by loading Ag or Au exhibited excellent photocatalytic CO2 reduction performance compared to Ru, Cu, and Pt. The reason is that Ag or Au has excellent photoelectric activity.
Therefore, the activity, selectivity, and stability of the co-catalysts depend primarily on the preparation and dispersion methods, as these methods directly affect the physicochemical properties (e.g., chemical composition, size, morphology) of the co-catalysts.

4. Multiple Bi-Based Photocatalysts for CO2 Reduction

Over the past 50 years, more than 150 semiconductors, including metal oxides, metal sulfides, carbon based materials, and MOF, have made significant progress in environmental remediation. Among them, Bi-based semiconductors have attracted much attention due to their unique layered structure, tunable electronic and visible light response performance, and great potential in energy conversion and environmental remediation applications.

4.1. Binary Bi-Based Semiconductor

Compared with other Bi-based materials, Bi2O3 has a suitable band gap (2.1–2.8 eV) [71]. Bi2O3 has α-, β-, γ-, δ-, ω-, ε- six polymorphic forms. Among them, β-Bi2O3 exhibits relatively excellent photocatalytic activity, which is attributed to its narrow bandgap and strong light absorption ability. However, Bi2O3 faces the problem of easy recombination of photo generated carriers when used as a photocatalyst, so it is often coupled with other semiconductors to construct heterojunctions for use. Liu et al. prepared Bi2Al4O7/β-Bi2O3 heterostructures through the one-step in situ auto-combustion method and successfully introduced oxygen vacancy in the synthesis process [72]. Bi2Al4O7/β-Bi2O3 showed high photocatalytic capacity in the photocatalytic reduction of CO2 to CO under the synergistic action of heterojunction photogenerated carrier separation and oxygen vacancy activation of CO2. Compared with the original β-Bi2O3, the CO yield of 0.14BAB catalyst is 13.2 μmol/g, which is an 8-fold increase. Compared with Bi2O3, Bi2S3 has a narrower bandgap (~1.3 eV) and higher absorption coefficient. Therefore, Bi2S3 often acts as a visible light absorber in the composition of photocatalysts. For example, Bi2S3 combines with TiO2, g-C3N4, and ZnIn2S4 to photocatalytic reduce CO2 to value-added chemicals such as methanol [73].

4.2. Ternary Bi-Based Semiconductors

Bismuth halide oxide (BiOX, X = Cl, Br, I) has been extensively studied in photocatalysis due to its excellent photocatalytic properties, environmental friendliness, and low price. The layered structure of BiOX provides enough space to polarize the associated atoms and orbitals, which excites the formation of an internal electric field between [Bi2O2] and halogen. The internal electric field can accelerate the separation and migration of photoexcited electron–hole pairs, and the photocatalytic activity of BiOX is significantly improved. Kong et al. [74] prepared BiOBr nanosheets containing oxygen vacancies to enhance the catalyst activity by ethylene glycol-assisted solvothermal method. The total CH4 yields were obtained at 4.86 µmol·g−1 and 9.58 µmol·g−1 for visible light irradiation and simulated sunlight irradiation, respectively. In contrast, the total CH4 yields were only 1.58 and 2.99 µmol·g−1 for BiOBr without oxygen vacancies, respectively. The surface oxygen vacancies adsorb and activate CO2, which reduces the energy barrier for charge transfer. It also has a strong ability to trap photo-generated electrons, which reduces the probability of electron–hole complexation. Therefore, in the preparation of bismuth halide oxide catalysts, oxygen vacancies could be introduced on the catalyst surface to improve the activity of the catalyst.

4.3. Quaternary Bi-Based Semiconductors

Bi2WO6 is one of the simplest Aurivillius oxides, with a perovskite-like [WO4]2− layer sandwiched between bismuth oxide [Bi2O2]2+ layers [75]. The CB of Bi2WO6 is composed of W 5d orbitals, while its VB is composed of a hybrid of O 2p and Bi 6s orbitals, resulting in a narrower bandgap and a visible light response ability. Bi2WO6 has very interesting physical and chemical properties, such as ferroelectricity, catalytic activity and nonlinear dielectric magnetic susceptibility. In recent years, Bi2WO6 has attracted widespread attention as a potential candidate for visible light-induced photocatalysts and photoelectrochemical (PEC) CO2 reduction. For example, the Bi2WO6 layer was prepared by Liang et al. [76] using oleate reacted with Bi to form a layered Bi-oleic acid complex and then adding sodium tungstate, which has excellent photocatalytic activity. The CO2 reduction activity was increased by nearly 130 times. Wang’s team prepared Bi2WO6/BiOCl heterojunction (BW-X) in situ on F-SnO2 transparent conductive glass by hydrothermal method and applied it to PEC CO2 reduction [77]. The BCW-X heterojunction had a excellent 2D layered/3D flower structure, and the exposed crystal surface of BiOCl had changed from the original (101) to (112) in the heterojunction, improving the separation efficiency of photogenerated electron–hole. Under simulated sunlight exposure, the BCW-X electrode in the BCW-X|KHCO3|BiVO4 PEC cell showed excellent ability to convert water and CO2 molecules into hydrocarbons at −1.0 V. Ethanol was produced at a rate of 11.4 μM h1 cm2 (600 μmol h1 g1) with a selectivity of 80.0%. The apparent quantum efficiency was as high as 0.63%, about 3 times that of the composite BiOCl-Bi2WO6 photocathode.
BiVO4 is an N-type semiconductor material with a narrow band gap of ~2.4 eV. BiVO4 has low toxicity and excellent visible-light responsive activity, exhibiting high activity for solar-driven water splitting and organic decomposition, which has attracted widespread interest in the scientific community. More attractive is that BiVO4 has an appropriate band edge, which makes it suitable for water oxidation coupled with CO2 reduction [78]. Li et al. [79] prepared 2D/2D BiVO4/Ti3C2Tx catalysts to verify the role of the surface contact interface on photocatalytic CO2 reduction. Photoexcitation of BiVO4 effectively transfers electrons to Ti3C2Tx through the formation of Schottky barriers. TEM observed electrostatic positive and negative charges on BiVO4 nanosheets and Ti3C2Tx flakes, and the prepared catalysts showed a robust surface contact interface. In addition, the transband alignment of BiVO4 with water redox levels makes it a valuable visible active photoanode for CO2 reduction systems [80]. However, BiVO4 still needs to overcome serious internal obstacles, such as poor electron transport and slow oxidation kinetics, to meet practical application requirements. For efficient CO2 reduction performance, a suitable photoanode with a low initial potential value is required to minimize the use of an external power source and drive the process primarily through solar energy. Therefore, BiVO4-based photoelectrodes with reduced initial potential and improved photocurrent are expected to develop superior CO2 reduction systems.
A number of specialized words appear in the paper, and the following table shows the abbreviations of the professional terms (Table 2). In addition, recent reports on the performance efficiencies associated with Bi-based photocatalysts are summarized in Table 3.

5. Conclusions

Photocatalytic CO2 reduction is a promising way to utilize CO2. It can use inexhaustible solar energy to convert CO2 into high-value-added chemicals. Impressive results have been achieved. The selection of highly active, selective, and stable materials is necessary to improve CO2 reduction activity. Bi-based photocatalysts have great potential in reducing CO2 to fuel under visible light due to unique electronic structure, crystal structure, and physical and chemical properties. In this paper, the recent progress of modification of Bi-based photocatalyts to improve their photocatalytic properties is reviewed. Many optimization strategies were discussed in detail to achieve strong photocatalytic performance, such as heterojunction, introduction of vacancies, morphology adjustment, and co-catalyst loading. Although significant improvements in photocatalytic efficiency have been achieved, there are still many opportunities and challenges for Bi-based photocatalysts.
1. There is currently no unified explanation about the reaction mechanism of Bi-based photocatalyst. The active species of catalyst in the reaction is controversial. For example, Bi is easily oxidized to form a thin layer of bismuth oxide in the surrounding environment. The presence of bismuth oxide on the catalyst surface and its influence on the photocatalytic process remain unclear. In the future, more attention should be paid to key issues such as the active site of reactants, charge transfer dynamics, and molecular orbital. Some advanced characterization techniques with high spatial, temporal, and spectral resolution, such as in situ XPS, X-ray absorption near-side structures XANES, and EXAFS, need to be introduced into the research to analyze the reaction process from the microscopic level. In addition, in situ analysis technology can be used to monitor the changes of catalyst and product in real time during the reaction process, thereby inferring the reaction pathway.
2. Due to the multiple reaction steps involved in CO2 reduction reaction, various products are formed, leading to selectivity issues. From recent research trends, it is not difficult to find that the main products of Bi-based materials are concentrated in C1 derivatives, such as CO, methane, formic acid, methanol. C2+ derivatives with higher added value, such as ethylene, ethanol, ethane, propanol, and acetone, are rarely reported due to their more difficult generation. Therefore, an effective Bi-based photocatalyst should be designed to selectively produce C2+ products.
3. The stability of Bi-based materials is a fundamental factor determining the lifespan and performance of photocatalysts. The prerequisite for developing efficient and stable photocatalysts is to suppress the poisoning and deactivation of Bi-based materials. Usually, researchers conduct multiple cyclic experiments in performance testing to evaluate the stability of the catalyst. However, the testing time is often limited to the range of a few hours to dozens of hours, far from meeting the requirements of industrial applications. It is insufficient to rely solely on the final reaction results in loop testing. The changes in the surface structure of Bi-based materials during the reaction process also need to be monitored and studied. In addition, the impact of photo corrosion on Bi-based materials in photocatalytic reactions also needs to be further studied.
4. The ultimate goal of photocatalytic research is to achieve industrial applications, and the preparation of catalysts in a sustainable manner is an inevitable problem to be faced and solved. At present, the quantity of Bi-based materials prepared is still relatively small, and the yield of photocatalysts prepared by existing methods is often less than a few grams. The method of large-scale production of Bi-based materials with excellent performance is not yet mature. Therefore, in order to meet the production requirements, a synthesis method with low cost and simple preparation process should be developed to prepare high efficiency and high stability of Bi-based photocatalyst, making it possible to realize practical application in the field of photocatalysis and energy sustainability.

Author Contributions

C.Z.: conceptualization, methodology, software, investigation, and writing—original draft; Q.S.: methodology, validation, formal analysis, and visualization; Z.J.: supervision, data curation, funding, acquisition, and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support for carrying out this work was provided by the Doctoral Research Foundation of Weifang University (2022BS13).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. 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]
  2. Shinde, G.Y.; Mote, A.S.; Gawande, M.B. Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol. Catalysts 2022, 12, 94. [Google Scholar] [CrossRef]
  3. Fujishima, A.; Honda, L. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  4. Li, X.; Fang, Y.; Wang, J.; Fang, H.; Xi, S.; Zhao, X.; Xu, D.; Xu, H.; Yu, W.; Hai, X.; et al. Ordered clustering of single atomic Te vacancies in atomically thin PtTe2 promotes hydrogen evolution catalysis. Nat. Commun. 2021, 12, 2351. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, Z.Y.; Sun, W.; Miao, W.K. Living Atomically Dispersed Cu Ultrathin TiO2 Nanosheet CO2 Reduction Photocatalyst. Adv. Sci. 2019, 6, 1900289. [Google Scholar] [CrossRef] [PubMed]
  6. Kumar, A.; Sharma, G.; Naushad, M.; Ahamad, T.; Veses, R.C.; Stadler, F.J. Highly visible active Ag2CrO4/Ag/BiFeO3@RGO nano-junction for photoreduction of CO2 and photocatalytic removal of ciprofloxacin and bromate ions: The triggering effect of Ag and RGO. Chem. Eng. J. 2019, 370, 148–165. [Google Scholar] [CrossRef]
  7. Zhao, Q.; Sun, J.; Li, S.C.; Huang, C.P.; Yao, W.F.; Chen, W.; Zeng, T.; Wu, Q.; Xu, Q.J. Single nickel atoms anchored on nitrogen-doped graphene as a highly active cocatalyst for photocatalytic H2 evolution. ACS Catal. 2018, 8, 11863–11874. [Google Scholar] [CrossRef]
  8. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal B 2019, 243, 556–565. [Google Scholar] [CrossRef]
  9. Yu, J.G.; Low, J.X.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced photocatalytic CO2 reduction activity of anatase TiO2 by coexposed {001} and {101} facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [CrossRef]
  10. Wang, Z.; Teramura, K.; Huang, Z.; Hosokawa, S.; Sakata, Y.; Tanaka, T. Tuning the selectivity toward CO evolution in the photocatalytic conversion of CO2 with H2O through the modification of Ag-loaded Ga2O3 with a ZnGa2O4 layer. Catal. Sci. Technol. 2016, 6, 1025–1032. [Google Scholar] [CrossRef]
  11. Yu, J.; Jin, J.; Cheng, B.; Jaroniec, M.A. noble metal-free reduced graphene oxide-CdS nanorod composite for the enhanced visible-light photocatalytic reduction of CO2 to solar fuel. J. Mater. Chem. A 2014, 2, 3407. [Google Scholar] [CrossRef]
  12. Yuan, Y.; Guo, R.-T.; Hong, L.-F.; Ji, X.-Y.; Li, Z.-S.; Lin, Z.-D.; Pan, W.-G. Recent advances and perspectives of MoS2-based materials for photocatalytic dyes degradation: A review. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125836. [Google Scholar] [CrossRef]
  13. An, X.; Li, K.; Tang, J. Inside Cover Picture: Cu2O/Reduced Graphene Oxide Composites for the Photocatalytic Conversion of CO2. ChemSusChem 2014, 7, 1086. [Google Scholar] [CrossRef] [PubMed]
  14. Shown, I.; Hsu, H.C.; Chang, Y.C.; Lin, C.H.; Roy, P.K.; Ganguly, A.; Wang, C.H.; Chang, J.K.; Wu, C.I.; Chen, L.C.; et al. Highly Efficient Visible Light Photocatalytic Reduction of CO2 to Hydrocarbon Fuels by Cu-Nanoparticle Decorated Graphene Oxide. Nano Lett. 2014, 14, 6097–6103. [Google Scholar] [CrossRef]
  15. Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.C.; Zou, Z.G. High-yield snthesis 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]
  16. Gondal, M.A.; Dastageer, M.A.; Oloore, L.E.; Baig, U. Laser induced selective photo-catalytic reduction of CO2 into methanol using In2O3-WO3 nano-composite. J. Photochem. Photobiol. A 2017, 343, 40–50. [Google Scholar] [CrossRef]
  17. 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]
  18. Zhu, C.Q.; Wang, Q.N.; Wu, C. Rapid and scalable synthesis of bismuth dendrites on copper mesh as a high-performance cathode for electroreduction of CO2 to formate. J. CO2 Util. 2022, 36, 96–104. [Google Scholar] [CrossRef]
  19. Wu, D.; Chen, W.Y.; Wang, X.W.; Fu, X.Z.; Luo, J.L. Metal-support interaction enhanced electrochemical reduction of CO2 to formate between graphene and Bi nanoparticles. J. CO2 Util. 2020, 37, 353–359. [Google Scholar] [CrossRef]
  20. Choi, K.M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C.A.; Barmanbek, J.T.; Alshammari, A.S.; Yang, P.; Yaghi, O.M. Plasmon-enhanced photocatalytic CO2 conversion within metal-organic frameworks under visible light. J. Am. Chem. Soc. 2017, 139, 356–362. [Google Scholar] [CrossRef]
  21. Zhang, Y.; Wang, L.; Dong, F.; Chen, Q.; Jiang, H.; Xu, M.; Shi, J. Non-additional carbon source one-step synthesis of Bi2O2CO3-based ternary composite for efficient Z-scheme photocatalysis. J. Colloid. Interf. Sci. 2019, 536, 575–585. [Google Scholar] [CrossRef] [PubMed]
  22. Zhao, Z.Y.; Dai, W.W. Electronic structure and optical properties of BiOI ultrathin films for photocatalytic water splitting. Inorg. Chem. 2015, 54, 10732–10737. [Google Scholar] [CrossRef] [PubMed]
  23. Hong, L.F.; Guo, R.T.; Yuan, Y.; Ji, X.Y.; Lin, Z.D.; Li, Z.S.; Pan, W.G. Recent progress of transition metal phosphides for photocatalytic hydrogen evolution. ChemSusChem 2021, 14, 539–557. [Google Scholar] [CrossRef] [PubMed]
  24. Marcì, G.; García-López, E.I.; Palmisano, L. Photocatalytic CO2 reduction in gassolid regime in the presence of H2O by using GaP/TiO2 composite as photocatalyst under simulated solar ligh. Catal. Commun. 2014, 53, 38–41. [Google Scholar] [CrossRef]
  25. Sun, Z.; Wang, H.; Wu, Z.; Wang, L. G-C3N4 based composite photocatalysts for photocatalytic CO2 reduction. Catal. Today 2018, 300, 160–172. [Google Scholar] [CrossRef]
  26. Que, M.; Cai, W.; Chen, J.; Zhu, L.; Yang, Y. Recent advances in g-C3N4 composites within four types of heterojunctions for photocatalytic CO2 reduction. Nanoscale 2021, 13, 6692–6712. [Google Scholar] [CrossRef]
  27. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2008, 8, 76–80. [Google Scholar] [CrossRef]
  28. Cao, S.; Yu, J. G-C3N4-based photocatalysts for hydrogen generation. J. Phys. Chem. Lett. 2014, 5, 2101–2107. [Google Scholar] [CrossRef]
  29. Wu, J.; Xie, Y.; Ling, Y.; Si, J.C.; Li, X.; Wang, J.L.; Ye, H.; Zhao, J.S.; Li, S.Q.; Zhao, Q.D. One-step synthesis and Gd3+ decoration of BiOBr microspheres consisting of nanosheets toward improving photocatalytic reduction of CO2 into hydrocarbon fuel. Chem. Eng. J. 2020, 400, 125944. [Google Scholar] [CrossRef]
  30. Zhao, Z.Y.; Dai, W.W. Structural, electronic, and optical properties of Eu-doped BiOX (X = F, Cl, Br, I): A DFT+U study. Inorg. Chem. 2014, 53, 13001–13011. [Google Scholar] [CrossRef]
  31. Xie, S.; Ma, W.; Wu, X.; Zhang, H.; Zhang, Q.; Wang, Y.; Wang, Y. Photocatalytic and Electrocatalytic Transformations of C1Molecules Involving C-C Coupling. Energy Environ. Sci. 2021, 14, 37–89. [Google Scholar] [CrossRef]
  32. Ajmal, S.; Yang, Y.; Tahir, M.A.; Li, K.; Bacha, A.U.R.; Nabi, I.; Liu, Y.; Wang, T.; Zhang, L. Boosting C2 Products in Electrochemical CO2 Reduction over Highly Dense Copper Nanoplates. Catal. Sci. Technol. 2020, 10, 4562–4570. [Google Scholar] [CrossRef]
  33. Hu, X.; Guo, R.T.; Chen, X.; Bi, Z.X.; Wang, J.; Pan, W.G. Bismuth-based Z-scheme structure for photocatalytic CO2 reduction: A review. J. Environ. Chem. Eng. 2022, 10, 108582. [Google Scholar] [CrossRef]
  34. Hirano, K.; Inoue, K.; Yatsu, T. Photocatalysed reduction of CO2 in aqueous TiO2 suspension mixed with copper powder. J. Photochem. Photobiol. A 1992, 64, 255. [Google Scholar] [CrossRef]
  35. Adachi, K.; Ohta, K.; Mizuno, T. Photocatalytic reduction of carbon dioxide to hydrocarbon using copper-loaded titanium dioxide. Sol. Energy 1994, 53, 187. [Google Scholar] [CrossRef]
  36. Wang, Z.-Y.; Chou, H.-C.; Wu, J.C.; Tsai, D.P.; Mul, G. CO2 photoreduction using NiO/InTaO4 in optical-fiber reactor for renewable energy. Appl. Catal. A Gen. 2010, 380, 172–177. [Google Scholar] [CrossRef]
  37. Li, Y.Y.; Fan, J.S.; Tan, R.Q.; Yao, H.C.; Peng, Y.; Liu, Q.C.; Li, Z.J. Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels. Nanomaterials 2020, 10, 337. [Google Scholar]
  38. Song, Y.; Chen, W.; Wei, W.; Sun, Y. Advances in Clean Fuel Ethanol Production from Electro-, Photo- and Photoelectro-Catalytic CO2 Reduction. Catalysts 2020, 10, 1287. [Google Scholar] [CrossRef]
  39. Xie, Q.; He, W.; Liu, S.; Li, C.; Zhang, J.; Wong, P.K. Bifunctional S-scheme g-C3N4/Bi/BiVO4 hybrid photocatalysts toward artificial carbon cycling. Chin. J. Catal. 2020, 41, 140–153. [Google Scholar] [CrossRef]
  40. Liu, L.; Dai, K.; Zhang, J.; Li, L. Plasmonic Bi-enhanced ammoniated alpha-MnS/Bi2MoO6 S-scheme heterostructure for visible-light-driven CO2 reduction. J. Colloid Interface Sci. 2021, 604, 844–855. [Google Scholar] [CrossRef]
  41. Wang, Q.T.; Fang, Z.X.; Zhang, W.; Zhang, D. High-Efciency g-C3N4 Based Photocatalysts for CO2 Reduction: Modifcation Methods. Adv. Fiber Mater. 2020, 4, 342–360. [Google Scholar] [CrossRef]
  42. Li, X.; Li, W.; Zhuang, Z.; Zhong, Y.; Li, Q.; Wang, L. Photocatalytic Reduction of Carbon Dioxide to Methane over SiO2-Pillared HNb3O8. J. Phys. Chem. C 2012, 116, 16047. [Google Scholar] [CrossRef]
  43. Li, K.; Peng, B.; Peng, T. Preparation of AgIn5S8/TiO2 Heterojunction Nanocomposite and Its Enhanced Photocatalytic H2 Production Property under Visible Light. ACS Catal. 2016, 6, 7485. [Google Scholar] [CrossRef]
  44. Yu, Y.Y.; Dong, X.A.; Chen, P.; Geng, Q.; Wang, H.; Li, J.Y.; Zhou, Y.; Dong, F. Synergistic Effect of Cu Single Atoms and Au-Cu Alloy Nanoparticles on TiO2 for Efficient CO2 Photoreduction. ACS Nano 2021, 15, 14453–14464. [Google Scholar] [CrossRef] [PubMed]
  45. Inoue, H.; Moriwaki, H.; Maeda, K.; Yoneyama, H. Photoreduction of carbon dioxide using chalcogenide semiconductor microcrystals. J. Photochem. Photobiol. A 1995, 86, 191. [Google Scholar] [CrossRef]
  46. Kim, C.; Cho, K.M.; Al-Saggaf, A.; Gereige, I.; Jung, H.-T. Z-scheme photocatalytic CO2 conversion on three-dimensional BiVO4/Carbon-coated Cu2O nanowire arrays under visible light. ACS Catal. 2018, 8, 4170–4177. [Google Scholar] [CrossRef]
  47. Jo, W.-K.; Kumar, S.; Eslava, S.; Tonda, S. Construction of Bi2WO6/RGO/g-C3N4 2D/2D/2D hybrid Z-scheme heterojunctions with large interfacial contact area for efficient charge separation and high-performance photoreduction of CO2 and H2O into solar fuels. Appl. Catal. B 2018, 239, 586–598. [Google Scholar] [CrossRef]
  48. Hu, X.; Hu, J.; Peng, Q.; Ma, X.; Dong, S.; Wang, H. Construction of 2D all-solid-state Z-scheme g-C3N4/BiOI/RGO hybrid structure immobilized on Ni foam for CO2 reduction and pollutant degradation. Mater. Res. Bull. 2020, 122, 110682. [Google Scholar] [CrossRef]
  49. 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]
  50. Dong, F.; Li, Q.; Sun, Y.; Ho, W.-K. Noble metal-like behavior of plasmonic Bi particles as a cocatalyst deposited on (BiO)2CO3 microspheres for efficient visible light photocatalysis. ACS Catal. 2014, 4, 4341–4350. [Google Scholar] [CrossRef]
  51. Zhao, Y.; Chen, G.; Bian, T.; Zhou, C.; Waterhouse, G.I.; Wu, L.Z.; Tung, C.H.; Smith, L.J.; O’Hare, D.; Zhang, T. Defect-rich ultrathin ZnAl-layered double hydroxide nanosheets for efficient photoreduction of CO2 to CO with water. Adv. Mater. 2015, 27, 7824–7831. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, S.; Pan, L.; Song, J.J.; Mi, W.; Zou, J.J.; Wang, L.; Zhang, X. Titanium-defected undoped anatase TiO2 with p-type conductivity, room-temperature ferromagnetism, and remarkable photocatalytic performance. J. Am. Chem. Soc. 2015, 137, 2975–2983. [Google Scholar] [CrossRef] [PubMed]
  53. Zhou, C.; Shi, X.; Li, D.; Song, Q.; Zhou, Y.; Jiang, D.; Shi, W. Oxygen vacancy engineering of BiOBr/HNb3O8 Z-scheme hybrid photocatalyst for boosting photocatalytic conversion of CO2. J. Colloid Interface Sci. 2021, 599, 245–254. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, X.L.; Wang, S.Y.; Yang, N.; Zhou, W.; Wang, P.; Jiang, K.; Li, S.; Song, H.; Ding, X.; Chen, H.; et al. Oxygen vacancies induced special CO2 adsorption modes on Bi2MoO6 for highly selective conversion to CH4. Appl. Catal. B Environ. 2019, 259, 118088. [Google Scholar] [CrossRef]
  55. Cao, Y.; Gao, Q.; Li, Q.; Jing, X.; Wang, S.; Wang, W. Synthesis of 3D porous MoS2/g-C3N4 heterojunction as a high efficiency photocatalyst for boosting H2 evolution activity. RSC Adv. 2017, 7, 40727–40733. [Google Scholar] [CrossRef]
  56. Gordon, T.R.; Cargnello, M.; Paik, T.; Mangolini, F.; Weber, R.T.; Fornasiero, P.; Murray, C.B. Nonaqueous synthesis of TiO2 nanocrystals using TiF4 to engineer morphology, oxygen vacancy concentration, and photocatalytic activity. J. Am. Chem. Soc. 2012, 134, 6751–6761. [Google Scholar] [CrossRef]
  57. Gao, X.; Shang, Y.; Liu, L.; Fu, F. Chemisorption-enhanced photocatalytic nitrogen fixation via 2D ultrathin p–n heterojunction AgCl/δ-Bi2O3 nanosheets. J. Catal. 2019, 371, 71–80. [Google Scholar] [CrossRef]
  58. Zuo, C.; Tai, X.; Jiang, Z.; Liu, M.; Jiang, J.; Su, Q.; Yan, X. S-Scheme 2D/2D Heterojunction of ZnTiO3 Nanosheets/Bi2WO6 Nanosheets with Enhanced Photoelectrocatalytic Activity for Phenol Wastewater under Visible Light. Molecules 2023, 28, 3495. [Google Scholar] [CrossRef]
  59. Wu, J.J.; Xiong, L.J.; Hu, Y.J. Organic half-metal derived erythroid-like BiVO4/hm-C4N3 Z-Scheme photocatalyst: Reduction sites upgrading and rate-determining step modulation for overall CO2 and H2O conversion. Appl.Catal. B Environ. 2021, 295, 120277. [Google Scholar] [CrossRef]
  60. Lu, M.; Li, Q.; Zhang, C.; Fan, X.; Li, L.; Dong, Y.; Chen, G.; Shi, H. Remarkable photocatalytic activity enhancement of CO2 conversion over 2D/2D g-C3N4/ BiVO4 Z-scheme heterojunction promoted by efficient interfacial charge transfer. Carbon 2020, 160, 342–352. [Google Scholar] [CrossRef]
  61. Yuan, Y.; Guo, R.T.; Hong, L.F.; Lin, Z.D.; Ji, X.Y.; Pan, W.G. Fabrication of a dual Sscheme Bi7O9I3/g-C3N4/Bi3O4Cl heterojunction with enhanced visible-lightdriven performance for phenol degradation. Chemosphere 2022, 287, 132241. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, X.Y.; Guo, R.T.; Qin, H.; Wang, Z.Y.; Shi, X.; Pan, W.G.; Tang, J.Y.; Jia, P.Y.; Miao, Y.F.; Gu, J.W. Fabrication of Bi2O2(OH)NO3/g-C3N4 nanocomposites for efficient CO2 photocatalytic reduction. Colloids Surf. A Physicochem. Eng. Asp. 2019, 580, 123782. [Google Scholar] [CrossRef]
  63. Bai, Y.; Ye, L.Q.; Wang, L.; Shi, X.; Wang, P.Q.; Wei, B. g-C3N4/Bi4O5I2 heterojunction with I3−/I redox mediator for enhanced photocatalytic CO2 conversion. Appl. Catal. B Environ. 2016, 194, 98–104. [Google Scholar] [CrossRef]
  64. Wu, J.; Li, K.; Yang, S.; Song, C.; Guo, X. In-situ construction of BiOBr/Bi2WO6 S-scheme heterojunction nanoflowers for highly efficient CO2 photoreduction: Regulation of morphology and surface oxygen vacancy. Chem. Eng. J. 2023, 452, 139493. [Google Scholar] [CrossRef]
  65. Miao, Z.; Zhang, Y.; Wang, N.; Xu, P.; Wang, X. BiOBr/Bi2S3 heterojunction with Sscheme structureand oxygen defects: In-situ construction and photocatalytic behavior for reduction of CO2 with H2O. J. Colloid Interface Sci. 2022, 620, 407–418. [Google Scholar] [CrossRef]
  66. Iizuka, K.; Wato, T.; Miseki, Y.; Saito, K.; Kudo, A. Photocatalytic Reduction of Carbon Dioxide over Ag Cocatalyst-Loaded ALa4Ti4O15(A=Ca, Sr, and Ba) Using Water as a Reducing Reagent. J. Am. Chem. Soc. 2011, 133, 20863. [Google Scholar] [CrossRef]
  67. Bai, S.; Wang, X.; Hu, C.; Xie, M.; Jiang, J.; Xiong, Y. Two-dimensional g-C3N4: An ideal platform for examining facet selectivity of metal co-catalysts in photocatalysis. Chem. Commun. 2014, 50, 6094. [Google Scholar] [CrossRef]
  68. Yang, G.; Miao, W.; Yuan, Z.; Jiang, Z.; Huang, B.; Wang, P.; Chen, J. Bi quantum dots obtained via in situ photodeposition method as a new photocatalytic CO2 reduction cocatalyst instead of noble metals: Borrowing redox conversion between Bi2O3 and Bi. Appl. Catal. B 2018, 237, 302–308. [Google Scholar] [CrossRef]
  69. Jiang, Y.; Chen, H.Y.; Li, J.Y.; Liao, J.F.; Zhang, H.H.; Wang, X.D.; Kuang, D.B. Z-scheme 2D/2D heterojunction of CsPbBr3/Bi2WO6 for improved photocatalytic CO2 reduction. Adv. Funct. Mater. 2020, 30, 2004293. [Google Scholar] [CrossRef]
  70. Yoshino, S.; Sato, K.; Yamaguchi, Y.; Iwase, A.; Kudo, A. Z-schematic CO2 reduction to CO through interparticle electron transfer between SrTiO3: Rh of a reducing photocatalyst and BiVO4 of a water oxidation photocatalyst under visible light. ACS Appl. Energy Mater. 2020, 3, 10001–10007. [Google Scholar] [CrossRef]
  71. Xie, Z.K.; Xu, Y.Y.; Li, D.; Meng, S.C.; Chen, M.; Jiang, D.L. Covalently Bonded Bi2O3 Nanosheet/Bi2WO6 Network Heterostructures for Efficient Photocatalytic CO2 Reduction. ACS Appl. Energ. Mater. 2021, 3, 12194–12203. [Google Scholar] [CrossRef]
  72. Liu, Y.; Guo, J.G.; Wang, Y.; Hao, Y.J.; Liu, R.H.; Li, F.T. One-step Synthesis of Defected Bi2Al4O7/β-Bi2O3 Heterojunctions for Photocatalytic Reduction of CO2 to CO. Green. Energy. Environ. 2021, 6, 244–252. [Google Scholar] [CrossRef]
  73. Jin, J.R.; He, T. Facile synthesis of Bi2S3 nanoribbons for photocatalytic reduction of CO2 into CH3OH. Appl. Surf. Sci. 2017, 394, 364–370. [Google Scholar] [CrossRef]
  74. Kong, X.Y.; Lee, W.P.C.; Ong, W.-J.; Chai, S.-P.; Mohamed, A.R. Oxygen-Deficient BiOBr as a Highly Stable Photocatalyst for Efficient CO2 Reduction into Renewable Carbon-Neutral Fuels. ChemCatChem 2016, 8, 3074–3081. [Google Scholar] [CrossRef]
  75. Dai, W.L.; Yu, J.J.; Deng, Y.Q.; Hu, X.; Wang, T.Y.; Luo, X.B. Facile synthesis of MoS2/Bi2WO6 nanocomposites for enhanced CO2 photoreduction activity under visible light irradiation. Appl. Surf. Sci. 2017, 403, 230–239. [Google Scholar] [CrossRef]
  76. Liang, L.; Lei, F.C.; Gao, S.; Sun, Y.F.; Jiao, X.C.; Wu, J.; Qamar, S.; Xie, Y. Single Unit Cell Bismuth Tungstate Layers Realizing Robust Solar CO2 Reduction to Methanol. Angew. Chem. Int. Ed. 2015, 54, 13971–13974. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, J.X.; Wei, Y.; Yang, B.J.; Wang, B.; Chen, J.Z.; Jing, H.W. In situ grown heterojunction of Bi2WO6/BiOCl for efficient photoelectrocatalytic CO2 reduction. J. Catal. 2019, 377, 209–217. [Google Scholar] [CrossRef]
  78. Wei, Z.H.; Wang, Y.F.; Li, Y.Y.; Zhang, L.; Yao, H.C.; Li, Z.J. Enhanced photocatalytic CO2 reduction activity of Z-scheme CdS/BiVO4 nanocomposite with thinner BiVO4 nanosheets. J. CO2 Util. 2018, 28, 15–25. [Google Scholar] [CrossRef]
  79. Li, L.; Yang, Y.; Zhou, B.; Zhou, Y.; Zou, Z.G. Dimensional matched ultrathin BiVO4/Ti3C2Tx heterosystem for efficient photocatalytic conversion of CO2 to methanol. Mater. Lett. 2022, 306, 130937. [Google Scholar] [CrossRef]
  80. Mane, P.; Bagal, I.V.; Bae, H.; Kadam, A.N.; Burungale, V.; Heo, J.; Byu, S.W.; Ha, J.S. Recent trends and outlooks on engineering of BiVO4 photoanodes toward efficient photoelectrochemical water splitting and CO2 reduction: A comprehensive review. Int. J. Hydrogen Energy 2022, 47, 39796–39828. [Google Scholar] [CrossRef]
  81. Bai, Y.; Yang, P.; Wang, L.; Yang, B.; Xie, H.Q.; Zhou, Y.; Ye, L.Q. Ultrathin Bi4O5Br2 nanosheets for selective photocatalytic CO2 conversion into CO. Chem. Eng. J. 2019, 360, 473–482. [Google Scholar] [CrossRef]
  82. Duan, Z.Y.; Zhao, X.J.; Wei, C.W.; Chen, L.M. Ag-Bi/BiVO4 chain-like hollow microstructures with enhanced photocatalytic activity for CO2 conversion. Appl. Catal. A Gen. 2020, 594, 117459. [Google Scholar] [CrossRef]
  83. Zhao, X.Z.; Xia, Y.G.; Li, H.P.; Wang, X.; Wei, J.; Jiao, X.L.; Chen, D.R. Oxygen vacancy dependent photocatalytic CO2 reduction activity in liquid-exfoliated atomically thin BiOCl nanosheets. Appl. Catal. B Environ. 2021, 297, 120426. [Google Scholar] [CrossRef]
  84. Guan, X.S.; Zhang, X.C.; Zhang, C.M.; Li, R.; Wang, Y.F.; Fan, C.M. One-step synthesis of novel Bi/Bi2SiO5 flower-like composites with highly-efficient CO2 photoreduction performance. Compos. Commun. 2020, 20, 100366. [Google Scholar] [CrossRef]
  85. Huang, L.; Duan, Z.Y.; Song, Y.Y.; Li, Q.S.; Chen, L.M. BiVO4 Microplates with Oxygen Vacancies Decorated with Metallic Cu and Bi Nanoparticles for CO2 Photoreduction. Appl. Nano Mater. 2021, 4, 3576–3585. [Google Scholar] [CrossRef]
  86. Guan, X.S.; Zhang, X.C.; Zhang, C.M.; Li, R.; Liu, J.X.; Wang, Y.F.; Wang, Y.W.; Fan, C.M.; Li, Z. In Situ Hydrothermal Synthesis of Metallic Bi Self-Deposited Bi2SiO5 with Enhanced Photocatalytic CO2 Reduction Performance. Sol. RRL 2022, 6, 2200346. [Google Scholar] [CrossRef]
  87. Wu, Y.Z.; Yue, X.Y.; Fan, J.J.; Hao, X.M.; Xiang, Q.J. In situ fabrication of plasmonic Bi/CsPbBr3 composite photocatalyst toward enhanced photocatalytic CO2 reduction. Appl. Surf. Sci. 2023, 609, 155391. [Google Scholar] [CrossRef]
  88. Chen, Q.; Ma, Y.C.; Qi, B.N.; Zhang, T.F.; Wang, L.L.; Shi, J.S.; Lan, X.F. Z-scheme Bi/AgBiS2/P25 for enhanced CO2 photoreduction to CH4 and CO with photo-themal synergy. Appl. Surf. Sci. 2021, 555, 149648. [Google Scholar] [CrossRef]
  89. Miao, Y.F.; Guo, R.T.; Gu, J.W.; Liu, Y.Z.; Wu, G.L.; Duan, C.P.; Pan, W.G. Z-Scheme Bi/Bi2O2CO3/Layered Double-Hydroxide Nanosheet Heterojunctions for Photocatalytic CO2 Reduction under Visible Light. ACS Appl. Nano Mater. 2021, 4, 4902–4911. [Google Scholar] [CrossRef]
  90. Zhu, J.Y.; Li, Y.P.; Zhao, J.; Wu, Y.S.; Li, F.T. Simultaneous Phosphorylation and Bi Modification of BiOBr for Promoting Photocatalytic CO2 Reduction. ACS Sustain. Chem. Eng. 2019, 7, 14953–14961. [Google Scholar] [CrossRef]
  91. Jiang, L.S.; Li, Y.; Wu, X.Y.; Zhang, G.K. Rich oxygen vacancies mediated bismuth oxysulfide crystals towards photocatalytic CO2-to-CH4 conversion. Sci China Mater. 2021, 64, 2230–2241. [Google Scholar] [CrossRef]
  92. Zhang, Y.Z.; Zhi, X.; Harmer, J.R.; Xu, H.L.; Davey, K.; Ran, J.R.; Qian, S.Z. Facet-specific Active Surface Regulation of BixMOy (M=Mo, V, W) Nanosheets for Boosted Photocatalytic CO2 reduction. Angew. Chem. Int. Ed. 2022, 61, e202212355. [Google Scholar] [CrossRef] [PubMed]
  93. Sun, N.C.; Zhou, M.; Ma, X.X.; Cheng, Z.H.; Wu, J.; Qi, Y.F.; Sun, Y.J.; Zhou, F.H.; Shen, Y.X.; Lu, S.Y. Self-assembled spherical In2O3/BiOI heterojunctions for enhanced photocatalytic CO2 reduction activity. J. CO2 Util. 2022, 65, 102220. [Google Scholar] [CrossRef]
  94. Li, R.J.; Luan, Q.J.; Dong, C.; Dong, W.J.; Tang, W.; Wang, G.; Lu, Y.F. Light-facilitated structure reconstruction on self-optimized photocatalyst TiO2@BiOCl for selectively efficient conversion of CO2 to CH4. Appl. Catal. B Environ. 2021, 286, 119832. [Google Scholar] [CrossRef]
  95. Jiang, L.S.; Wang, D.Y.; Hu, Y.; Guo, T.; Liu, C.Y.; Liang, C.; Du, W.; Li, X.Q.; Liu, W. Surface-iodination-induced efficient charge separation in bismuth oxysulfide crystals for enhanced photocatalytic CO2 conversion. Chem. Eng. J. 2023, 453, 139848. [Google Scholar] [CrossRef]
  96. Cao, S.W.; Shen, B.J.; Tong, T.; Fu, J.W.; Yu, J.G. 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  97. Zhao, D.W.; Xuan, Y.M.; Zhang, K.; Liu, X.L. Highly Selective Production of Ethanol over Hierarchical Bi@Bi2MoO6 composite via Bicarbonate-Assisted Photocatalytic CO2 Reduction. ChemSusChem 2021, 14, 3293–3302. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanism of photocatalytic process.
Figure 1. Mechanism of photocatalytic process.
Molecules 28 03982 g001
Figure 2. Mechanism of photocatalytic CO2 reduction [33]. Copyright 2022, Elsevier.
Figure 2. Mechanism of photocatalytic CO2 reduction [33]. Copyright 2022, Elsevier.
Molecules 28 03982 g002
Figure 3. The bandgap structure of common Bi-based catalysts.
Figure 3. The bandgap structure of common Bi-based catalysts.
Molecules 28 03982 g003
Figure 4. (a) Absorption of B1−CO2 on BMO OVs and (b) B2−CO2 on BMO [54]. Copyright 2019, Elsevier.
Figure 4. (a) Absorption of B1−CO2 on BMO OVs and (b) B2−CO2 on BMO [54]. Copyright 2019, Elsevier.
Molecules 28 03982 g004
Figure 5. (a,b) SEM imagesZnTiO3/Bi2WO6 heterojunction sample, (c) TEM image and (d) HRTEM image of ZnTiO3/Bi2WO6 heterojunction sample [58].
Figure 5. (a,b) SEM imagesZnTiO3/Bi2WO6 heterojunction sample, (c) TEM image and (d) HRTEM image of ZnTiO3/Bi2WO6 heterojunction sample [58].
Molecules 28 03982 g005
Figure 6. Photocatalytic enhancement mechanism for reducing CO2: (a) charge transfer mechanism, and (b) Z-scheme mechanism [63]. Copyright 2016, Elsevier.
Figure 6. Photocatalytic enhancement mechanism for reducing CO2: (a) charge transfer mechanism, and (b) Z-scheme mechanism [63]. Copyright 2016, Elsevier.
Molecules 28 03982 g006
Figure 7. Charge transfer processes in the S-scheme structure [33]. Copyright 2022, Elsevier.
Figure 7. Charge transfer processes in the S-scheme structure [33]. Copyright 2022, Elsevier.
Molecules 28 03982 g007
Table 1. Photocatalytic CO2 reduction products and corresponding electrode potentials.
Table 1. Photocatalytic CO2 reduction products and corresponding electrode potentials.
ReactionReduction Potential (vs. NHE at pH = 7)
CO2 + 2H+ + 2e → HCOOH−0.61
CO2 + 2H+ + 2e → CO + H2O−0.53
CO2 + 4H+ + 4e → HCHO + H2O−0.48
CO2 + 4H+ + 4e → C + 2H2O−0.20
CO2 + 6H+ + 6e → CH3OH + H2O−0.38
CO2 + 8H+ + 8e → CH4 + 2H2O−0.24
2CO2 + 2H+ + 2e → C2H4 + 4H2O−0.34
CO2 + 2H+ + 2e → CH3OH + 3H2O−0.33
2CO2 + 14H+ + 14e → C2H6 + 4H2O−0.27
2H+ + 2e → H2−0.42
Table 2. Abbreviations of professional terms.
Table 2. Abbreviations of professional terms.
NamesAbbreviations
Valence bandVB
Conduction bandCB
Lowest unoccupied molecular orbitalLUMO
Highest occupied molecular orbitalHOMO
Table 3. Summary of photocatalytic CO2 reduction performance with Bi-based materials.
Table 3. Summary of photocatalytic CO2 reduction performance with Bi-based materials.
CatalystsPreparation MethodReaction ConditionMajor ProductsProduction Rate
/µmol gcat−1 h−1
Ref.
Light SourceSolution
ultrathin Bi4O5Br2precursor method300 W high-pressure xenon lampCO2/H2O vaporCO31.57 μmol g−1 h−1[81]
Ag-Bi/BiVO4galvanic replacement reaction300 W Xe-illuminator with a light cutoff filter (λ > 420 nm)0.5 mL H2O deionized waterCO~5.19 μmol g−1 h−1[82]
a-BiOClliquid exfoliation300 W Xenon arc lamp with a filter (AM 1.5 G)50 mg of catalyst and 100 mL of Milli-Q waterCO8.99 µmol g−1 h−1[83]
Bi/Bi2SiO5one-step hydrothermal strategy50 mL of deionized water300 W Xe-lampCO33.02 µmol g−1 h−1[84]
Cu-Bi/BiVO4solvothermal method1 mL distill waterXe lamp with a light intensity of 160 mW·cm2 and a wavelength range of 420~780 nmCO11.15 μmol h−1 g−1[85]
Bi/Bi2SiO5OH-assisted hydrothermal controllable route50 mL of deionized water300 W Xe-lamp (200–2500 nm)CO62.70 μmol h−1 g−1[86]
Bi/CsPbBr3in-situ growth method0.5 mL of deionized water300 W Xenon lampCO19.1 µmol g−1 h−1[87]
Bi/AgBiS2/P25one-step solvothermal treatment4 mL of deionized water300 W Xenon lampCH4 (CO)4.31 µmol g−1 h−1 (6.37 µmol g−1 h−1)[88]
Bi2O2CO3/Bi/NiAl-LDHhydrothermal
method
100 mL deionized water300 W Xenon lamp with a cut−800 nm filterCH456.64 μmol gcat−1[89]
P/Bi-BiOBrin-situ bismuth deposition and phosphorus modification2.5 mL H2O300 W xenon lamp with 400 nm filterCH40.62 µmol g−1 h−1[90]
Bi2O2Shydrothermal method1.2 g of Na2CO3 and 2 mL of H2SO4 (1:1 vol.)300 W of Xe lamp with a 420-nm filterCH443.87 µmol g−1 h−1[91]
Bi2MoO6sonication-assisted chemical reduction20 mg of photocatalyst and 5 mL water300 W Xenon lampCH4 (CO)12.4 µmol g−1 h−1 (61.5 µmol g−1 h−1)[92]
In2O3/BiOIsolvothermal methodsTEOA as a sacrificial agen300 W Xe lamp with a cut-off filter (λ ≥ 420 nm)CH4 (CO)5.69 µmol g−1 h−1 (11.98 µmol g−1 h−1)[93]
TiO2@BiOClchemical impregnation and calcination5 μL acetonitrile and 1 mL
H2O
300 W high pressure xenon lampCH4168.5 µmol g−1 h−1[94]
surface iodinated Bi2O2Shydrothermal reaction1.2 g of Na2CO3 and 2 mL H2SO4 (1:1 vol)Xe lamp (300 W) with a 420 nm cutoff filterCH453.35 µmol g−1 h−1[95]
Ti3C2/Bi2WO6etching and ultrasonic exfoliation0.084 g NaHCO3 and 0.3 mL H2SO4 (2 mol L−1)Xe lamp (300 W)CH41.78 µmol g−1 h−1[96]
Bi@Bi2MoO6solvothermal approachXe lamp (λ ≥ 400)30 mL uniform solution containing 50 mg of photocatalyst and 0.42 g NaHCO3C2H5OH17.93 µmol g−1 h−1[97]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zuo, C.; Su, Q.; Jiang, Z. Advances in the Application of Bi-Based Compounds in Photocatalytic Reduction of CO2. Molecules 2023, 28, 3982. https://doi.org/10.3390/molecules28103982

AMA Style

Zuo C, Su Q, Jiang Z. Advances in the Application of Bi-Based Compounds in Photocatalytic Reduction of CO2. Molecules. 2023; 28(10):3982. https://doi.org/10.3390/molecules28103982

Chicago/Turabian Style

Zuo, Cheng, Qian Su, and Zaiyong Jiang. 2023. "Advances in the Application of Bi-Based Compounds in Photocatalytic Reduction of CO2" Molecules 28, no. 10: 3982. https://doi.org/10.3390/molecules28103982

Article Metrics

Back to TopTop