Next Article in Journal
Co-Adsorption of Alcohols and Water in JUK-8 Studied Using Quasi-Equilibrated Thermodesorption
Next Article in Special Issue
Recent Progress of Ion-Modified TiO2 for Enhanced Photocatalytic Hydrogen Production
Previous Article in Journal
Catalytic Strategies for the Cycloaddition of CO2 to Epoxides in Aqueous Media to Enhance the Activity and Recyclability of Molecular Organocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Effect of Defect Modulation on the Photocatalytic Reduction Performance of Carbon Dioxide

1
College of Chemistry & Chemical and Environmental Engineering, Weifang University, Weifang 261061, China
2
College of Science, Nanjing Forestry University, Nanjing 210037, China
*
Authors to whom correspondence should be addressed.
Submission received: 22 April 2024 / Revised: 9 May 2024 / Accepted: 10 May 2024 / Published: 14 May 2024

Abstract

:
Constructive defect engineering has emerged as a prominent method for enhancing the performance of photocatalysts. The mechanisms of the influence of defect types, concentrations, and distributions on the efficiency, selectivity, and stability of CO2 reduction were revealed for this paper by analyzing the effects of different types of defects (e.g., metallic defects, non-metallic defects, and composite defects) on the performance of photocatalysts. There are three fundamental steps in defect engineering techniques to promote photocatalysis, namely, light absorption, charge transfer and separation, and surface-catalyzed reactions. Defect engineering has demonstrated significant potential in recent studies, particularly in enhancing the light-harvesting, charge separation, and adsorption properties of semiconductor photocatalysts for reducing processes like carbon dioxide reduction. Furthermore, this paper discusses the optimization method used in defect modulation strategy to offer theoretical guidance and an experimental foundation for designing and preparing efficient and stable photocatalysts.

1. Introduction

As the economy rapidly expands, a significant amount of fossil fuels is being consumed, resulting in substantial carbon dioxide emissions. These emissions exacerbate the greenhouse effect and disrupt the global ecological balance. Given that fossil fuels are non-renewable energy sources [1], escalating energy consumption and carbon dioxide concentration are having profound impacts on human habitats and the Earth’s ecosystem. Consequently, a pressing global research focus has emerged on how to reduce atmospheric CO2 levels and utilize them sustainably. Photocatalysis technology has emerged as a promising solution for converting CO2 into crucial chemical fuels such as methane (CH4). Although numerous studies have successfully converted CO2 into clean fuels, including methanol [2,3,4,5], formic acid [6,7,8,9], and methane [10,11,12,13] via photocatalysis, the yields of these products remain low and are inadequate for industrial-scale applications.
Photocatalysis has undergone a remarkable evolution spanning over a century since its inception. In 1911, the concept of “photocatalysis” was first proposed. Baur et al. [14] validated the ability of ZnO to reduce Ag+ to Ag using light energy. TiO2’s capacity to fade Prussian blue was also established [15]. A pivotal moment occurred when Honda et al. [16] demonstrated TiO2’s capacity for photolytic aquatic H2 production, marking a significant milestone in photocatalytic degradation and synthesis reactions. Halman et al. further expanded this field by demonstrating CO2 reduction under photocathode catalysis [17]. The researchers went on to achieve photocatalytic CO2 reduction with TiO2 in water, yielding products like CH3OH and CH4. Subsequently, metal oxides and sulfides such as SrTiO3, WO3, ZnO, and CdS were found to exhibit remarkable CO2 reduction performance [18,19,20]. In addition, various metal oxides and metal sulfides became mainstream catalysts [21,22,23,24]. Metal–organic frameworks (MOFs) have also emerged as promising photocatalysts, due to their impressive specific surface area and CO2 capture capabilities [25]. Recently, Niu et al. [26] introduced a graphite-like g-C3N4 catalyst that exhibits superior photocatalytic performance.
Ideal photocatalysts for CO2 photocatalytic reduction reactions exhibit several key characteristics [27,28,29,30,31,32,33]: (1) a suitable bandgap width, excellent visible and even infrared light absorption ability, and the enhanced utilization of sunlight. (2) Appropriate conduction band and valence band positions to obtain sufficient reduction potential for CO2 reduction and O2 precipitation potential. (3) A large specific surface area, excellent CO2 adsorption performance, and the rapid transport of reactive substances. (4) The fast separation and transport of photogenerated carriers with enhanced carrier separation efficiency. (5) Abundant active sites for CO2 reactions. However, single photocatalysts are more or less flawed and it is difficult to achieve the performance of all desirable catalysts. For most catalysts, it is imperative to optimize the performance of the catalyst through certain strategies to compensate for its deficiencies. The activation of stabilized CO2 molecules is very difficult, due to the extremely high dissociation energy of the C=O bond (~750 kJ mol−1), which inevitably leads to slow reaction kinetics in the photocatalytic reduction of CO2 [34]. Defect engineering provides an effective means to improve the performance of photocatalysts. Defect engineering could improve carrier diffusion efficiency, induce electron enrichment, and promote the chemisorption and activation of CO2 molecules [35]. Table 1 shows the standard redox potentials for various reactions during CO2 reduction [36].
Unlike previous reports in the literature [37,38], this paper offers a comprehensive review of the types of defects introduced in recent photocatalysts, along with their respective mechanisms. Through a detailed analysis of the various components’ roles in the photocatalytic process, this paper elucidates the reasons for the observed improvements in photocatalytic performance. This provides a solid theoretical foundation and experimental guidance for the advancement of novel and efficient photocatalysts. Furthermore, this paper delves into the effects of diverse defect types, such as metallic and non-metallic types, on photocatalyst performance. By combining these theoretical insights with experimental data and reaction principles, it explores their potential applications in the photocatalytic reduction of CO2. The mechanisms behind the influence of defect types, concentrations, and distributions on CO2 reduction efficiency, selectivity, and stability are revealed. Additionally, this paper surveys the key research hotspots and advancements in this field. The paper also surveys the most prevalent methods to enhance photocatalytic performance, including precise defect concentration control, defect distribution optimization strategies, and the synergistic effects of composite defects. The aim is to provide theoretical guidance and an experimental basis for the design and preparation of highly efficient and stable photocatalysts. Finally, the paper concludes with an overview of the future of defect engineering in photocatalytic CO2 reduction.

2. Basic Principles of the Photocatalytic Reduction of CO2

The photocatalytic CO2 reaction mechanism primarily comprises three stages, exemplified here through conventional semiconductors. As depicted in Figure 1, the initial stage involves the semiconductor catalyst being irradiated by photons exceeding the band gap, triggering electron jumps. They leap to form holes at the top of the valence band and photogenerated electrons at the bottom of the conduction band [39]. Immediately following the second stage is the transfer of the photogenerated electrons and holes; this process must occur swiftly due to the short lifespan of the photogenerated electrons and holes, which is only nanoseconds. To catalyze the reaction, the photogenerated holes and electrons must be promptly transferred to the active surface. In the third stage, the photogenerated electrons on the active surface combine with CO2, breaking the C=O bond and reacting with H2O to form organic fuel. Concurrently, the photogenerated holes on the active surface react with H2O to produce O2.
Hence, two prerequisites must be met for the photocatalytic reduction of CO2: (i) the photon energy must be equal to or exceed the band gap energy; and (ii) the CB potential must be more negative than the surface electron acceptor potential, while the VB potential must be more positive than the surface electron donor potential. This enables the reaction process of the photocatalytic reduction of CO2 to occur [40]. Defect engineering plays an important role in understanding the mechanism of the CO2 reduction reaction and improving the performance of CO2 photoreduction. Single-component photocatalysts with defects provide excellent prototypes for theoretical calculations. This facilitates an in-depth study of the CO2 conversion mechanism and the constitutive relationship. However, some reported that defect engineering strategies for photocatalysts are quite complex and time-consuming. As a result, the conversion efficiency of carbon dioxide is still very low (not just in defective engineered photocatalysts), and there is still a long way to go to achieve industrialization in the future [40].

3. Defect Types and the Properties of Photocatalysts

3.1. Metallic and Metal-Free Vacancy Defects

ZnS (VZn-ZnS) with metal cation (Zn) vacancies was prepared using an acid-etching strategy by Pang et al. [41] The experimental results show that VZn-ZnS had an excellent selectivity for HCOOH generation in the absence of co-catalysts, averaging above 85%. An examination of the experimental and simulation outcomes reveals that the generation of VZn led to a reduction in the energy potential barrier, an elevation in surface energy, and a faster charge separation process, ultimately leading to a higher yield of HOOH. A ZnIn2S4 catalyst was prepared by Jiao et al. [42]. Their findings suggested that the introduction of VZn was beneficial for elevating the charge density of adjacent S atoms, thus accelerating carrier transfer and separation. This modification significantly shortened the migration time of photogenerated electrons to ~15 ps. The catalyst with VZn exhibited a higher average CO yield of 33.2 μmol g−1 h−1, representing a 3.6-fold increase compared to the catalyst without VZn. Huang et al. [43] presented a novel three-phase photocatalytic CO2 reduction system by depositing Ag-TiO2 nanoparticles at the gas–water interface. This innovation led to a substantial boost in the utilization of visible light and an impressive eight-fold surge in CH4 production. Wang et al. [44] alternatively deposited Pt-Cu alloy nanoparticles with varying Pt/Cu ratios onto the TiO2 surface. This approach dramatically enhanced visible light absorption through the localized plasma resonance effect of the PtCu alloy, while also elevating the catalytic activity for the photoreduction of CO2 to CH4. Furthermore, it increased the selectivity of CH4 to 100%. Thus, the localized plasma resonance effect of the noble metal could significantly enhance the visible light utilization of the catalyst and obtain ultra-high activity. Fu et al. [45] significantly enhanced CO2 adsorption and activation and improved photocatalyst activity after the introduction of Co monoatomic sites in g-C3N4.
Carbon nitride (g-C3N4), a metal-free polymer, has garnered significant attention for its photocatalytic CO2 reduction capabilities, due to its simplicity in preparation, remarkable chemical stability, and notably potent reduction potential (ECB ≈ −1.0 eV) [46,47,48,49,50]. However, achieving efficient photocatalytic performance for CO2 reduction with g-C3N4 remains challenging, due to its limited number of active sites. Recent studies have demonstrated that introducing non-metallic defects into g-C3N4 can generate more active sites, broaden the visible light response, and enhance electron capture, thereby significantly boosting its photocatalytic performance [51,52,53,54]. Figure 2A,B illustrates the optimal reaction pathways for CO2 conversion on g-C3N4 and S-doped g-C3N4 [51]. Shen et al. [55] successfully prepared g-C3N4 with carbon vacancies using NH3 thermal treatment. This approach enhances catalytic activity for the photocatalytic reduction of CO2 by increasing the presence of basic amino groups, elongating the C=O bond lengths, and obtaining stronger CO2 adsorption energy.

3.2. Metal and Metal-Free Doping Defects

The ideal semiconductor catalyst exhibits a band gap ranging from 1.5 eV to 2.5 eV, while many semiconductors represented by TiO2 have band gaps in excess of 2.5 eV. Consequently, the utilization of visible light is limited, necessitating the narrowing of the band gap. This adjustment is commonly achieved through doping techniques, including metallic element doping, non-metallic element doping, and self-doping. These methods introduce impurity levels within the wider band gap, effectively narrowing it and enhancing the utilization of visible and infrared light. Zhang et al. [56] reduced the band gap of TiO2 from 3.20 eV to 2.06 eV by doping Cu in TiO2. They used the prepared Cu-TiO2 for the photoreduction of CO2, and the yield of CH4 was increased by a factor of 21 in comparison with that of commercial TiO2 (P25). Similarly, Xiong et al. [57] prepared N-doped TiO2, achieving a reduced band gap of 2.34 eV, significantly improving visible light utilization. Zhang et al. [58] employed a surface hydrogenation strategy to synthesize Ti3+ self-doped TiO2 mesoporous nanorods, narrowing the band gap to 2.11 eV and enhancing visible light utilization. Collectively, these strategies demonstrate that adjusting the band gap can lead to the improved utilization of visible and infrared light, ultimately resulting in a more efficient photocatalytic reduction of CO2.
Abdellah et al. [59] successfully composited the phosphine-phosphated molecular catalyst [ReIBr(bpy)(CO)3]0 with the wide band-gap semiconductor TiO2. This integration broadened the light absorption range of TiO2 and significantly improved its performance in reducing CO2 to CO in visible light. This enhancement was achieved through the rapid electron injection provided by the phosphine-phosphine-phosphated molecular catalyst. To summarize, defect engineering introduces specific defects to the surface and bulk phase of the photocatalyst, such as vacancy, lattice distortion, and crystal surface defects, to regulate its electronic structure and energy level structure, thus significantly affecting the performance of the photocatalyst, including its light absorption capacity, photogenerated carrier separation, and migration efficiency, along with CO2 adsorption capacity (Table 2).

4. The Effect of Defect Modulation

Defect modulation plays a crucial role in the performance of the photocatalytic reduction of carbon dioxide (CO2). Experimental results in recent years have shown that the efficiency, selectivity, and stability of CO2 reduction can be significantly improved by precisely controlling the type, concentration, and distribution of defects, as well as their relationship with photocatalytic activity.

4.1. Effect of Defect Types on CO2 Reduction Efficiency

The choice of defect type has a significant effect on the electronic structure and the formation of active sites of the photocatalyst. Metallic and non-metallic defects can promote CO2 reduction by changing the energy band structure and electronic properties of the photocatalysts. Yang et al. [61] showed that oxygen vacancy-modified titanium dioxide, OV-TiO2, exhibits little photocurrent under visible light (λ > 420 nm) irradiation. When the oxygen vacancy sites were filled with hydrogen atoms by hydrogenation treatment, it showed greatly enhanced photocurrent under visible light (λ > 420 nm) irradiation, which improved the catalytic activity for the reduction of CO2. Graphitic carbon nitride (GCN) can be used as a two-dimensional photocatalyst, and inter-band light absorption can usually be achieved by doping or vacancy engineering [62]. The creation of nitrogen vacancies and oxygen doping in its structure usually leads to a redshift of its absorption edge, while in the case of carbon vacancies, this leads to a blueshift of its absorption edge. However, both the redshift and blueshift in the light absorption of defect-modified GCN showed enhanced photocatalytic performance for CO2 reduction [60,63,64].

4.2. Effect of Defect Concentration on the Selectivity of CO2 Reduction

The precise control of defect concentration is critical to the selectivity of CO2 reduction products. Appropriate defect concentrations could provide sufficient active sites while avoiding excessive carrier complexation, thus improving the selectivity of a particular product. Pang et al. [65] reported the production of Ni-doped ZnS (ZnS:Ni) nanocrystals, which were mainly used to enhance the yield of HCOOH. Figure 3a–d shows the changes in the reduction properties of different CO2 with different Ni additions. Excessive Ni doping did not guarantee excellent CO2 reduction performance. As can be seen in Figure 3b–d, the electron spin resonance (ESR), photoluminescence (PL), and DFT results show that a small amount of Ni-doping produces abundant VS, which limits the recombination of photogenerated electrons and holes, while excessive Ni doping can reduce the number of VS, which is extremely unfavorable for photocatalytic CO2 reduction. Therefore, an appropriate amount of introduced defect concentration is very important for CO2 reduction.

4.3. Effect of Defect Distribution on the Stability of CO2 Reduction

The distribution of defects has an important effect on the stability and long-term activity of photocatalysts. Uniformly distributed defects help to maintain the structural stability of the photocatalyst and reduce the local stress concentration, thus improving the durability of the catalysts. Fu et al. [60] prepared ultra-thin g-C3N4 nanosheets, which improved the inherent CO2 adsorption and activation ability with a larger specific surface area. This significantly enhanced CO2 adsorption and activation and improved the photocatalyst activity after the introduction of O-doping in g-C3N4. In addition, the optimization of defect distribution avoids catalyst deactivation by reducing the aggregation of active sites.

4.4. Relationship between Defect Modulation and Photocatalytic Activity

There is a close relationship between defect modulation and photocatalytic activity. Defects could not only participate in CO2 reduction as active sites but also regulate photocatalytic activity by affecting the separation and migration of photogenerated carriers. Shi et al. [66] prepared two-dimensional graphitic carbon nitride nanosheets with nitrogen vacancies (DCN-x, where x denotes the mass of tartaric acid in the preparation process) using the tartaric acid-assisted one-step thermal polymerization of dicyandiamide. The visible light absorption of all the DCN nanosheets containing nitrogen vacancies was enhanced compared with the pristine DCN nanosheets. The change in band gap structure was confirmed by calculating the Kubelka–Munk function. The reduction of the band gap enhances the visible light absorption during photocatalysis. The presence of defects can also promote the effective separation of electron-hole pairs through the formation of intermediate energy bands or the modification of the electronic structure, further improving photocatalytic efficiency. The creation of oxygen vacancies on the WO3 atomic layer can introduce intermediate energy band states and extend their photoresponse to visible and even infrared light [67].
The performance of photocatalysts in the CO2 reduction reaction can be significantly improved by the rational design and precise control of defect type, concentration, and distribution. These strategies provide an important theoretical basis and experimental guidance for the development of new and efficient photocatalysts, which is of great significance for achieving sustainable energy conversion and environmental purification.

5. Optimization Methods for Defect Modulation Strategies

In the research field of the photocatalytic reduction of CO2, the optimization of defect modulation strategies is an important way to improve the performance of catalysts. In recent years, researchers have made remarkable progress by designing different types of defects, precisely controlling the defect concentration, optimizing defect distribution, and exploiting the synergistic effect of composite defects.

5.1. Optimization Strategies for Defect Distribution

The distribution of defects is equally important in the performance of photocatalysts. Uniformly distributed defects contribute to the separation efficiency of photogenerated carriers and the stability of the photocatalysts. For example, uniformly distributed defects could be formed on the surface of photocatalysts by specific synthesis methods, such as via solution or template methods. These methods help to improve the activity and selectivity of the photocatalyst while maintaining long-term stability. Tan et al. [68] used the template method to synthesize multilayered MnS/In2S3 p-n heterojunction nanosheets using Mn2+-doped MIL-68(In) as a template and a significant improvement in the activity of the catalysts was obtained, compared to pristine MnS and In2S3 (Figure 4A,B,D). In addition, the photocatalytic stability of the MnS/In2S3 heterojunction was tested by performing the reaction four times in succession. After 12 h of testing, no significant decrease in activity was observed (Figure 4C), indicating that the activity of the prepared catalysts was stable. Zhao et al. [69] prepared a Bi@Bi2MoO6 photocatalyst with excellent catalytic performance, wherein Bi nanoparticles were grown on the surface of Bi2MoO6 nanosheets with oxygen vacancies. The abundant oxygen vacancies on the surface of Bi2MoO6 nanosheets not only altered the chemical coordination of the Bi3+ ions to form the metal nanoparticles but also facilitated the CO2 molecules in their chemisorption on the catalyst surface. The resulting catalyst had strong light absorption efficiency, abundant active sites, and effective spatial separation of photogenerated carriers, and exhibited higher reactivity than the pure Bi2MoO6 catalyst.

5.2. Synergistic Effects of Compound Defects

The introduction of composite defects could realize a synergistic effect between different defects and further enhance the performance of photocatalysts. For example, composite defects can be formed by introducing both metallic and nonmetallic defects in semiconductor materials, and the interactions among these defects can help to improve the separation efficiency of photogenerated carriers and the reduction activity of CO2. In addition, the composite defects could also improve the photocatalytic efficiency by modulating the energy band structure of the photocatalysts for better absorption of sunlight. TiO2 is one of the most common catalysts in photocatalytic reactions. Liu et al. [70] successfully prepared an Fe-N co-doped TiO2 heterojunction photocatalyst (FeTi@C). The results show that the catalyst has a small band gap and excellent visible light absorption. Since the doped N comes from the original MOF, and the radii of Fe3+ and Ti+ are similar, the Ti-O-Fe tetrahedron and more oxygen vacancies are provided, which improves the photocatalytic efficiency of the CO2 reduction of CH4. In particular, Fe0.8Ti@C has 7.8 and 10.2 times higher CH4 yields than the original TiMOF template and the undoped Ti@C, respectively. This synthesis method is a new attempt to harness the synergistic effect of complex defects.

5.3. Precise Control of Defect Concentration

The precise control of defect concentration is essential to optimize the performance of the photocatalyst. Defects at too high or low a concentration may lead to the compounding of photogenerated carriers or insufficient active sites. In order to improve the performance of photocatalytic CO2 reduction materials, researchers have employed chemical vapor deposition, thermal annealing, and heavy ion irradiation, which are used to control different defect types and concentrations on the catalyst surface. Guo et al. [71] prepared porous ZnO nanosheets (PNS-ZnO) by calcining ZnS nanosheets in air and produced coated g-C3N4 (g-CN) nanofilms on the surface of PNS-ZnO by thermal deposition. The results show that due to the formation of a heterojunction between PNS-ZnO and g-CN, the defect concentration is uniformly distributed. This leads to significant inhibition of the photogenerated electron and hole complexes, g-CN loading on light absorption, and photocatalytic activity, due to its significant effect on light absorption and photocatalytic activity, prompting all the PNS-ZnO@g-CN nanocomposites to exhibit excellent photocatalytic activity. Among these catalysts, the catalytic activity of [email protected] appeared to be particularly outstanding. Ding et al. [72] employed single-atom In to modify g-C3N4 by a single-atom-assisted thermal polymerization process. Single-atom In-bonded N atoms (Inδ+__N4) were constructed on the (002) crystalline surface of g-C3N4. The results show that the (Inδ+__N4) structure modulates the g-C3N4 layer spacing and reduces the defect concentration, as well as changing the reaction path. The experimental result was a CO yield of 398.87 μmol g−1 h−1 with a selectivity close to 100%.
In general, such defect engineering could result in the formation of a differential charge distribution of polarized atom pairs. The polarized atom pairs have the potential to induce the adsorption of intermediates, such as methane and methanol, due to charge distribution differentiation, thereby enabling distinctive carbon-carbon bond formation processes, as reported in previous studies [73,74]. Figure 5 shows that the reaction pathways and mechanisms for the preparation of ethanol from CO2 and C2H6 [73]. Figure 6 depicts the CO2 synthesis CO2RR reaction pathway and the optimized catalyst model [74]. Therefore, a variety of catalysts need to be developed for the production of valuable products such as ethanol, ethane, or two-carbon hydrocarbons. Currently, some scholars have already converted CO2 to ethanol [75]. It is believed that in future research, new catalysts will be developed for the production of substances such as C3 and C4.

6. Conclusions

This paper aims to investigate the role and impact of defect modulation in the photocatalytic reduction of CO2. By analyzing the effects of different types of defects (e.g., metallic and non-metallic defects) on the performance of photocatalysts, the mechanisms of the influence of defect type, concentrations, and distribution on CO2 reduction efficiency, selectivity, and stability were revealed. In addition, the optimization methods of defect regulation strategies were discussed in this paper, with a view to providing theoretical guidance and experimental basis for the design and preparation of efficient and stable photocatalysts.
  • Employing advanced analytical techniques such as electron paramagnetic resonance (EPR), X-ray absorption near-edge structures (XANES), and extended X-ray absorption fine structures (EXAFS), the electronic structure and the intricate local coordination environment surrounding the defects can be meticulously delineated. This fine-level analysis facilitates careful study of the reaction process at the microscopic scale, revealing the complex ways in which defects affect the performance of photocatalysts. The subtle nuances of how defects can either enhance or impede photocatalytic activity can be uncovered by dissecting the reaction mechanisms at such a fundamental level. This profound understanding is crucial for the strategic manipulation of defects to optimize the efficiency of photocatalytic systems. It can achieve the customization of defect characteristics to achieve the desired photocatalytic performance, thereby driving innovation in the development of advanced photocatalytic materials.
  • With the aid of advanced computational methods and simulation software, the existence and distribution of defects in different photocatalysts can be simulated, and their impact on photocatalytic reaction activity can be predicted. This can not only reveal the influence of defects on the electronic structure and band structure of photocatalysts but also predict the impact of defects on the performance of photocatalysts, such as in the light absorption, separation, and transport of photogenerated carriers. Theoretical simulations allow for the identification of potential photocatalyst candidates, which can then be experimentally verified. By comparing the simulation results with experimental data, a more accurate understanding of how defects affect photocatalytic performance is obtained, thereby guiding the design and optimization of novel efficient photocatalysts. Simultaneously, based on the simulation and experimental findings, improvements in composition, structure, and preparation methods can be made to enhance the overall photocatalytic performance.
  • The application of the photocatalytic reduction of CO2 into high-value fuels is very promising, but unfavorable factors such as low spectral utilization and the short lifetime of photogenerated charges have been restricting the industrial application process. Therefore, in order to meet production requirements, it is necessary to develop a low-cost and simple synthesis method with a simple preparation process to prepare highly efficient and stable photocatalysts, which will make it possible to achieve practical applications and energy sustainability in the field of photocatalysis. The distribution of defects is important in the performance of photocatalysts. Uniformly distributed defects contribute to the separation efficiency of photogenerated carriers and the stability of the photocatalysts. For example, uniformly distributed defects could be formed on the surface of photocatalysts by specific synthesis methods, such as via solution or template methods. These methods help to improve the activity and selectivity of the photocatalyst while maintaining long-term stability.

Author Contributions

C.Z.: Writing and revising the manuscript. X.T.: Conceptualization, methodology, software, investigation, writing—original draft. H.W. and Q.S.: 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 Natural Science of Shandong Province, China (ZR2023QB086).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Omer, A.M. Energy, environment and sustainable development. Renew. Sust. Energ. Rev. 2008, 12, 2265–2300. [Google Scholar] [CrossRef]
  2. Zhu, X.L.; Zong, H.B.; Yuan, Z.M.; Wang, S.H.; Zeng, G.X.; Xu, H.; Jiang, Z.Y.; Ozin, G.A. Supercharged CO2 photothermal catalytic methanation: High conversion, rate, and selectivity. Angew. Chem. Int. Edit. 2023, 62, e202218694. [Google Scholar] [CrossRef] [PubMed]
  3. Low, J.X.; Yu, J.G.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef] [PubMed]
  4. Tai, X.S.; Yan, X.H.; Wang, L.H. Synthesis, structural characterization, Hirschfeld surface analysis, density functional theory, and photocatalytic CO2 reduction activity of a new Ca(II) complex with a Bis-Schiff base ligand. Molecules 2024, 29, 1047. [Google Scholar] [CrossRef] [PubMed]
  5. Albero, J.; Peng, Y.; Garcia, H. Photocatalytic CO2 reduction to C2+ products. ACS Catal. 2020, 10, 5734–5749. [Google Scholar] [CrossRef]
  6. 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]
  7. 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]
  8. Wang, L.H.; Tai, X.S. Synthesis, structural characterization, hirschfeld surface analysis and photocatalytic CO2 reduction activity of a new dinuclear Gd (III) complex with 6-phenylpyridine-2-carboxylic acid and 1, 10-phenanthroline ligands. Molecules 2023, 28, 7595. [Google Scholar] [CrossRef] [PubMed]
  9. 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]
  10. 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]
  11. 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]
  12. Duic, N.; Guzovic, Z.; Vyatcheslav, K.; Klemes, J.J.; Mathiessen, B.V.; Yan, J.Y. Sustainable development of energy, water and environment systems. Appl. Energy 2013, 101, 3–5. [Google Scholar] [CrossRef]
  13. Yang, X.G.; Wang, D.W. Photocatalysis: From fundamental principles to materials and applications. ACS Appl. Energy Mater. 2018, 1, 6657–6693. [Google Scholar] [CrossRef]
  14. Baur, E.; Perret, A. The action of light on dissolved silver salts in the presence of zinc oxide. Hel. Chim. Acta 1924, 7, 910–915. [Google Scholar] [CrossRef]
  15. Goodeve, C.F.; Kitchener, J.A. The mechanism of photosensitisation by solids. Trans. Faraday Soc. 1938, 34, 902–908. [Google Scholar] [CrossRef]
  16. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  17. Halmann, M. Photoelectrochemical reduction of aqueous carbon dioxide on p-type gallium phosphide in liquid junction solar cells. Nature 1978, 275, 115–116. [Google Scholar] [CrossRef]
  18. Luo, C.; Zhao, J.; Li, Y.; Zhao, W.; Zeng, Y.; Wang, C. Photocatalytic CO2 reduction over SrTiO3:Correlation between surface structure and activity. Appl. Surf. Sci. 2018, 447, 627–635. [Google Scholar] [CrossRef]
  19. Hao, X.Q.; Zhou, J.; Cui, Z.W.; Wang, Y.C.; Wang, Y.; Zou, Z. Zn-vacancy mediated electron-hole separation in ZnS/g-C3N4 heterojunction for efficient visible-light photocatalytic hydrogen production. Appl. Catal. B-Environ. 2018, 229, 41–51. [Google Scholar]
  20. Lai, C.; Zhang, M.; Li, B.; Huang, D.; Zeng, G.; Qin, L.; Liu, X.; Yi, H.; Cheng, M.; Li, L.; et al. Fabrication of CuS/BiVO4 (040) binary heterojunction photocatalysts with enhanced photocatalytic activity for ciprofloxacin degradation and mechanism insight. Chem. Eng. J. 2019, 358, 891–902. [Google Scholar] [CrossRef]
  21. Fajrina, N.; Tahir, M. Engineering approach in stimulating photocatalytic H2 production in a slurry and monolithic photoreactor systems using Ag-bridged Z-scheme pCN/TiO2 nanocomposite. Chem. Eng. J. 2019, 374, 1076–1095. [Google Scholar] [CrossRef]
  22. Fajrina, N.; Tahir, M. 2D-montmorillonite-dispersed g-C3N4/TiO2 2D/0D nanocomposite for enhanced photo-induced H2 evolution from glycerol-water mixture. Appl. Surf. Sci. 2019, 471, 1053–1064. [Google Scholar] [CrossRef]
  23. Chen, C.; Jin, J.; Chen, S.; Wang, T.; Xiao, J.; Peng, T. In-situ growth of ultrafine ZnO on g-C3N4 layer for highly active and selective CO2 photoreduction to CH4 under visible light. Mater. Res. Bull. 2021, 137, 111177. [Google Scholar] [CrossRef]
  24. Han, P.; Mih, A.A.; Ferre-borrull, J. Interplay between morphology, optical properties, and electronic structure of solution-processed Bi2S3 colloidal nanocrystals. J. Phys. Chem. C 2015, 119, 10693–10699. [Google Scholar] [CrossRef]
  25. Abdel-Mageed, A.M.; Rungtaweevoranit, B.; BParlinska-Wojtan, M.; Pei, X.; Yaghi, O.M.; Behm, R.J. Highly active and stable single-atom Cu catalysts supported by a metal—Organic framework. J. Am. Chem. Soc. 2019, 141, 5201–5210. [Google Scholar] [CrossRef]
  26. Niu, P.; Yang, Y.; Jimmy, C.Y.; Liu, G.; Cheng, H.-M. Switching the selectivity of the photoreduction reaction of carbon dioxide by controlling the band structure of a g-C3N4 photocatalyst. Chem. Commun. 2014, 50, 10837–10840. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. 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]
  29. 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]
  30. 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]
  31. 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]
  32. 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]
  33. 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]
  34. Humphrey, V.; Zscheischler, J.; Ciais, P.; Gudmundsson, L.; Sitch, S.; Seneviratne, S.I. Sensitivity of atmospheric CO2 growth rate to observed changes in terrestrial water storage. Nature 2018, 560, 628–631. [Google Scholar] [CrossRef] [PubMed]
  35. Soden, B.J.; Collins, W.D.; Feldman, D.R. Reducing uncertainties in climate models. Science 2018, 361, 326–327. [Google Scholar] [CrossRef]
  36. Xu, Q.L.; Xia, Z.H.; Zhang, J.M.; Wei, Z.Y.; Guo, Q.; Jin, H.; Tang, H.; Li, S.; Pan, X.; Su, Z.; et al. Recent advances in solar-driven CO2 reduction over g-C3N4-based photocatalysts. Carbon Energy 2023, 5, e205. [Google Scholar] [CrossRef]
  37. Zhao, W.G.; Wang, F.; Zhao, K.Y.; Liu, K.Y.; Zhu, X.T.; Yan, L.; Yin, Y.; Xu, Q.; Yin, D.L. Recent advances in the catalytic production of bio-based diol 2,5-bis (hydroxymethyl) furan. Carbon Resour. Convers. 2023, 6, 116–131. [Google Scholar] [CrossRef]
  38. Lo, A.Y.; Chung, Y.C.; Xie, P.J.; Delbari, H.; Yang, H.; Taghipour, H. Effect of Ag-doping strategies on the Lewis acid/base behavior of mesoporous TiO2 photocatalyst and its performance in CO2 photoreduction. Appl. Mater. Today 2023, 32, 101811. [Google Scholar] [CrossRef]
  39. Li, X.; Yu, G.J.; Jaroniec, M.; Chen, X. Cocatalysts for selective photoreduction of CO2 into solar fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef]
  40. Shen, M.; Zhang, L.X.; Shi, J.L. Defect Engineering of Photocatalysts towards Elevated CO2 Reduction Performance. ChemSusChem 2021, 14, 2635–2654. [Google Scholar] [CrossRef] [PubMed]
  41. Pang, H.; Meng, X.; Li, P.; Chang, K.; Zhou, W.; Wang, X.; Zhang, X.; Jevasuwan, W.; Fukata, N.; Wang, D.; et al. Cation vacancy-initiated CO2 photoreduction over ZnS for efficient formate production. ACS Energy Lett. 2019, 4, 1387–1393. [Google Scholar] [CrossRef]
  42. Jiao, X.; Chen, Z.; Li, X.; Sun, Y.; Gao, S.; Yan, W.; Wang, C.; Zhang, Q.; Lin, Y.; Luo, Y.; et al. Defect-mediated electron-hole separation in one-unit-cell ZnIn2S4 layers for boosted solar-driven CO2 reduction. J. Am. Chem. Soc. 2017, 139, 7586–7594. [Google Scholar] [CrossRef]
  43. Wang, Q.L.; Dong, P.M.; Huang, Z.F.; Zhang, X.W. Synthesis of Ag or Pt nanoparticle-deposited TiO2 nanorods for the highly efficient photoreduction of CO2 to CH4. Chem. Phys. Lett. 2015, 639, 11–16. [Google Scholar] [CrossRef]
  44. Wang, J.Y.; Li, Y.Z.; Zhao, J.T.; Xiong, Z.; Zhao, Y.C.; Zhang, J.Y. PtCu alloy cocatalysts for efficient photocatalytic CO2 reduction into CH4 with 100% selectivity. Catal. Sci. Technol. 2022, 12, 3454–3463. [Google Scholar] [CrossRef]
  45. Fu, J.; Zhu, L.; Jiang, K.; Liu, K.; Wang, Z.; Qiu, X.; Li, H.; Hu, J.; Pan, H.; Lu, Y.-R.; et al. Activation of CO2 on graphitic carbon nitride supported single-atom cobalt sites. Chem. Eng. J. 2021, 415, 128982. [Google Scholar] [CrossRef]
  46. Li, Y.; Gong, F.; Zhou, Q.; Feng, X.H.; Fan, J.J.; Xiang, Q.J. Crystalline isotype heptazine-/triazine-based carbon nitride heterojunctions for an improved hydrogen evolution. Appl. Catal. B Environ. 2020, 268, 118381. [Google Scholar] [CrossRef]
  47. Zhou, G.; Shan, Y.; Hu, Y.Y.; Xu, X.Y.; Long, L.Y.; Zhang, J.L.; Dai, J.; Guo, J.H.; Shen, J.C.; Li, S.; et al. Half-metallic carbon nitride nanosheets with microgrid mode resonance structure for efficient photocatalytic hydrogen evolution. Nat. Commun. 2018, 9, 3366. [Google Scholar] [CrossRef] [PubMed]
  48. Liang, M.F.; Borjigin, T.; Zhang, Y.H.; Liu, H.; Liu, B.H.; Guo, H. Z-scheme Au@Void@ g-C3N4/SnS yolk-shell heterostructures for superior photocatalytic CO2 reduction under visible light. ACS Appl. Mater. Interfaces 2018, 10, 34123–34131. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, Q.X.; Ai, L.H.; Jiang, J. MXene-derived TiO2@C/g-C3N4 heterojunctions for highly efficient nitrogen photofixation. J. Mater. Chem. A 2018, 6, 4102–4110. [Google Scholar] [CrossRef]
  50. Shen, Y.; Han, Q.T.; Hu, J.Q.; Gao, W.; Wang, L.; Yang, L.Q.; Gao, C.; Shen, Q.; Wu, C.P.; Wang, X.Y.; et al. Artificial trees for artificial photosynthesis: Construction of dendrite-structured α-Fe2O3/g-C3N4 Z-Scheme system for efficient CO2 reduction into solar fuels. ACS Appl. Energy Mater. 2018, 3, 6561–6572. [Google Scholar] [CrossRef]
  51. Wang, K.; Li, Q.; Liu, B.; Cheng, B.; Ho, W.; Yu, J. Sulfur-doped g-C3N4 with enhanced photocatalytic CO2- reduction performance. Appl. Catal. B-Environ. 2015, 176, 44–52. [Google Scholar] [CrossRef]
  52. Ruan, D.; Kim, S.; Fujitsuka, M.; Majima, T. Defects rich g-C3N4 with mesoporous structure for efficient photocatalytic H2 production under visible light irradiation. Appl. Catal. B Environ. 2018, 238, 638–646. [Google Scholar] [CrossRef]
  53. Tay, Q.; Kanhere, P.; Ng, C.F.; Chen, S.; Chakraborty, S.; Huan, A.C.H.; Sum, T.C.; Ahuja, R.; Chen, Z. Defect engineered g-C3N4 for efficient visible light photocatalytic hydrogen production. Chem. Mater. 2015, 27, 4930–4933. [Google Scholar] [CrossRef]
  54. Sun, S.Q.; Wu, Y.C.; Zhu, J.F.; Lu, C.J.; Sun, Y.; Wang, Z.; Chen, J. Stabilizing plasmainduced highly nitrogen-deficient g-C3N4 by heteroatom-refilling for excellent lithium-ion battery anodes. Chem. Eng. J. 2022, 427, 131032. [Google Scholar] [CrossRef]
  55. Shen, M.; Zhang, L.; Wang, M.; Tian, J.; Jin, X.; Guo, L.; Wang, L.; Shi, J. Carbon-vacancy modified graphitic carbon nitride: Enhanced CO2 photocatalytic reduction performance and mechanism probing. J. Mater. Chem. A 2019, 7, 1556–1563. [Google Scholar] [CrossRef]
  56. Zhang, T.; Low, J.X.; Huang, X.X.; Al-Sharab, J.F.; Yu, J.; Asefa, T. Copper-Decorated Microsized Nanoporous Titanium Dioxide Photocatalysts for Carbon Dioxide Reduction by Water. Chemcatchem 2017, 9, 3054–3062. [Google Scholar] [CrossRef]
  57. Xiong, C.; Zhang, S.S.; Zhao, Y.; Zheng, M.; Hou, D.D.; Xiao, J. Enhancing visible light photocatalytic performance with N-doped TiO2 nanotube arrays assisted by H2O2. Int. J. Mod. Phys. B 2019, 33, 1940025. [Google Scholar] [CrossRef]
  58. Zhang, Y.G.; Zhang, Y.H.; Zhang, H.F.; Bai, L.Q.; Hao, L.; Ma, T.Y.; Huang, H.W. Defect engineering in metal sulfides for energy conversion and storage. Coordin. Chem. Rev. 2021, 448, 214147. [Google Scholar] [CrossRef]
  59. Abdellah, M.; El-Zohry, A.M.; Antila, L.J.; Windle, C.D.; Reisner, E.; Hammarström, L. Time-Resolved IR Spectroscopy Reveals a. Mechanism with TiO2 as a Reversible Electron Acceptor in a TiO2-Re Catalyst System for CO2 Photoreduction. J. Am. Chem. Soc. 2017, 139, 1226–1232. [Google Scholar] [CrossRef] [PubMed]
  60. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical Porous O-Doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef]
  61. Yang, Y.; Yin, L.C.; Gong, Y.; Niu, P.; Wang, J.Q.; Gu, L.; Chen, X.; Liu, G.; Wang, L.; Cheng, H.M. An Unusual Strong Visible-Light Absorption Band in Red Anatase TiO2 Photocatalyst Induced by Atomic Hydrogen-Occupied Oxygen Vacancies. Adv. Mater. 2018, 30, 1704479. [Google Scholar] [CrossRef]
  62. Li, F.; Yue, X.Y.; Zhang, D.N.; Fan, J.J.; Xiang, Q.J. Targeted regulation of exciton dissociation in graphitic carbon nitride by vacancy modification for efficient photocatalytic CO2 reduction. Appl. Catal. B-Environ. 2021, 292, 120179. [Google Scholar]
  63. Ikreedeegh, R.R. Recent developments of Fe-based metal-organic frameworks and their composites in photocatalytic applications: Fundamentals, synthesis and challenges. Russ. Chem. Rev. 2022, 91, RCR5064. [Google Scholar] [CrossRef]
  64. Li, Y.; Ho, W.; Lv, K.; Zhu, B.; Lee, S.C. Carbon vacancy-induced enhancement of the visible light-driven photocatalytic oxidation of NO over g-C3N4 nanosheets. Appl. Surf. Sci. 2018, 430, 380–389. [Google Scholar] [CrossRef]
  65. Pang, H.; Meng, X.; Song, H.; Zhou, W.; Yang, G.; Zhang, H.; Izumi, Y.; Takei, T.; Jewasuwan, W.; Fukata, N.; et al. Probing the role of nickel dopant in aqueous colloidal ZnS nanocrystals for efficient solar-driven CO2 reduction. Appl. Catal. B 2019, 244, 1013–1020. [Google Scholar] [CrossRef]
  66. Shi, H.N.; Long, S.R.; Hou, J.G.; Ye, L.; Sun, Y.; Ni, W.; Song, C.; Li, K.; Gurzadyan, G.G.; Guo, X. Defects Promote ultrafast charge separation in graphitic carbon nitride for enhanced visible-light-driven CO2 reduction activity. Chem.-Eur. J. 2018, 25, 5028–5035. [Google Scholar] [CrossRef] [PubMed]
  67. Pan, R.R.; Liu, J.; Zhang, J.T. Defect Engineering in 2D Photocatalytic Materials for CO2 Reduction. ChemNanoMat 2021, 7, 737–747. [Google Scholar] [CrossRef]
  68. Tan, J.; Yu, M.; Cai, Z.; Lou, X.; Wang, J.; Li, Z. MOF-derived synthesis of MnS/In2S3 p-n heterojunctions with hierarchical structures for efficient photocatalytic CO2 reduction. J. Colloid Interface Sci. 2021, 588, 547–556. [Google Scholar] [CrossRef] [PubMed]
  69. 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]
  70. Liu, T.; Qileng, A.; Wubulikasimu, N.; Liu, Z.X.; Liu, W.P.; Liu, Y.J. Self-sacrificial templated tynthesis of Fe/N co-doping TiO2 for enhanced CO2 photocatalytic reduction. ChemNanoMat 2023, 9, e202300343. [Google Scholar] [CrossRef]
  71. Guo, Q.; Fu, L.M.; Yan, T.; Tian, W.; Ma, D.F.; Li, J.M.; Jiang, Y.N.; Wang, X.T. Improved photocatalytic activity of porous ZnO nanosheets by thermal deposition graphene-like g-C3N4 for CO2 reduction with H2O vapor. Appl. Surf. Sci. 2020, 509, 144773. [Google Scholar] [CrossRef]
  72. Ding, C.; Lu, X.X.; Tao, B.; Yang, L.Q. Interlayer spacing regulation by single-atom indium Inδ+__N4 on carbon nitride for boosting CO2/CO photo-conversion. Adv. Funct. Mater. 2023, 33, 2302824. [Google Scholar] [CrossRef]
  73. Paulista, L.O.; Albero, J.; Martins, R.J.E.; Boaventura, R.A.R.; Vilar, V.J.P.; Silva, T.F.C.V.; García, H. Turning carbon dioxide and ethane into ethanol by solar-driven heterogeneous photocatalysis over RuO2- and NiO-co-Doped SrTiO3. Catalysts 2021, 11, 461. [Google Scholar] [CrossRef]
  74. Cheng, H.; Wu, X.; Li, X.; Nie, X.; Fan, S.; Feng, M.; Fan, Z.; Tan, M.; Chen, Y.; He, G. Construction of atomically dispersed Cu-N4 sites via engineered coordination environment for high-efficient CO2 electroreduction. Chem. Eng. J. 2021, 407, 126842. [Google Scholar] [CrossRef]
  75. Ikreedeegh, R.R.; Tahir, M. A critical review in recent developments of metal-organic-frameworks (MOFs) with band engineering alteration for photocatalytic CO2 reduction to solar fuels. J. CO2 Utili. 2021, 43, 101381. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the photocatalytic reaction mechanism [39].
Figure 1. Schematic diagram of the photocatalytic reaction mechanism [39].
Molecules 29 02308 g001
Figure 2. (A) Optimal reaction pathways for CO2 conversion on g-C3N4 and (B) S-doped g-C3N4 [51]. Copyright 2015, Elsevier (Amsterdam, The Netherlands).
Figure 2. (A) Optimal reaction pathways for CO2 conversion on g-C3N4 and (B) S-doped g-C3N4 [51]. Copyright 2015, Elsevier (Amsterdam, The Netherlands).
Molecules 29 02308 g002
Figure 3. (a) ESR spectra and (b) PL spectra of undoped ZnS and ZnS:Ni nanocrystals at room temperature. The 425 nm emission that is indicated represents the sulfur vacancy signature. High-resolution XPS spectra of (c) Zn 2p and (d) S 2p, respectively, of the undoped, ZnS:Ni (0.1%), ZnS:Ni (0.2%), and ZnS:Ni (4.0%) nanocrystals [65]. Copyright 2019, Elsevier.
Figure 3. (a) ESR spectra and (b) PL spectra of undoped ZnS and ZnS:Ni nanocrystals at room temperature. The 425 nm emission that is indicated represents the sulfur vacancy signature. High-resolution XPS spectra of (c) Zn 2p and (d) S 2p, respectively, of the undoped, ZnS:Ni (0.1%), ZnS:Ni (0.2%), and ZnS:Ni (4.0%) nanocrystals [65]. Copyright 2019, Elsevier.
Molecules 29 02308 g003
Figure 4. (A) M-S plot of the MnS/In2S3 heterojunction, (B) Nyquist plots, (C) transient photocurrents, and (D) PL spectra of different samples [68]. Copyright 2021, Elsevier.
Figure 4. (A) M-S plot of the MnS/In2S3 heterojunction, (B) Nyquist plots, (C) transient photocurrents, and (D) PL spectra of different samples [68]. Copyright 2021, Elsevier.
Molecules 29 02308 g004
Figure 5. Reaction pathways and mechanisms for the preparation of ethanol, using CO2 and C2H6 as the reactants [73].
Figure 5. Reaction pathways and mechanisms for the preparation of ethanol, using CO2 and C2H6 as the reactants [73].
Molecules 29 02308 g005
Figure 6. (a) The CO2 synthesis CO2RR reaction pathway and the optimized catalyst model. The gray, blue, and orange shapes represent the C, N, and Cu atoms, respectively. (b) Free energy diagrams for CO2RR [74]. Copyright 2021, Elsevier. * Active sites on the catalyst surface.
Figure 6. (a) The CO2 synthesis CO2RR reaction pathway and the optimized catalyst model. The gray, blue, and orange shapes represent the C, N, and Cu atoms, respectively. (b) Free energy diagrams for CO2RR [74]. Copyright 2021, Elsevier. * Active sites on the catalyst surface.
Molecules 29 02308 g006
Table 1. Standard redox potentials for some specific reactions in the reduction of CO2 [36].
Table 1. Standard redox potentials for some specific reactions in the reduction of CO2 [36].
Specific ReactionRedox Potentials E° (V) versus SHE at pH = 7
CO + e → CO−1.9
CO2 + 2H+ + 2e → HCOOH−0.61
CO2 + 2H+ + 2e → CO + H2O−0.53
CO2 + 4H+ + 4e → HCHO + H2O-0.48
CO2 + 6H+ + 6e → CH3OH + H2O−0.38
CO2 + 8H+ + 8e → CH4 + 2H2O−0.24
2CO2 + 12H+ + 12e → C2H4 + 4H2O−0.34
2CO2 + 12H+ + 12e → C2H5OH + 3H2O−0.33
2CO2 + 14H+ + 14e → C2H6 + 4H2O−0.27
3CO2 + 18H+ + 18e→ C3H7OH + 5H2O−0.32
3CO2 + 20H+ + 20e → C3H8 + 6H2O−0.33
2H+ + 2e → H2−0.41
2H2O + 4h+ → 4H+ + O20.82
Table 2. The role of defects in CO2 conversion and product selectivity.
Table 2. The role of defects in CO2 conversion and product selectivity.
Type of DefectDefective PhotocatalystProductYield and SelectivityRoles of Defects on PhotocatalysisMechanismRef.
Cation vacancyZnSHCOOHSelectivity up to 86.6%Lowering the barrier of CO2RR
and suppressing proton adsorption
Changing the electronic states of density[41]
Cation vacancyZnIn2S4CO33.2 μmol g−1 h−1Increasing the light absorption, improving the CO2 adsorption capacity, and enhancing surface hydrophilicityIncreased charge density[42]
Metal vacancy Ag-TiO2CH416.0 ppm/g hPromoting the separation of photo-induced electron-hole pairsForming a Schottky barrier and surface plasmon resonance[43]
Metal vacancyPtCu/TiO2CH4Selectivity up to 100%Enhancing the adsorption/activation of CO2/CO and the further hydrogenation of COReducing the activation energy barriers of *CO2 and *CHO and inhibiting the desorption of *CO[44]
Metal vacancyCo-CNCO94.9 umol/g/hReducing the energy barrier of CO2 adsorption/activation and promoting theStrong interaction between electrons[45]
Carbon vacancyGCNCO4.18 mmol g−1 h−1Enhancing CO2 adsorption/activation, upshifting the conduction band and elevating the charge carrier concentration and lifetimeAttenuating the exciton effect and facilitating charge carrier generation[55]
DopingCu-TiO2CH48.04 μmol g−1 h−1Increasing visible-light absorption in the materials, and suppressing photogenerated electron-hole recombinationServing as electron traps[56]
DopingO-doped g-C3N4CH3OH0.88 µmol g-1 h-1Improving light utilization efficiency and CO2 affinity, and separation efficiency of photogenerated charge carriersOptimizing the band structure[60]
* Active sites on the catalyst surface.
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.; Tang, X.; Wang, H.; Su, Q. A Review of the Effect of Defect Modulation on the Photocatalytic Reduction Performance of Carbon Dioxide. Molecules 2024, 29, 2308. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules29102308

AMA Style

Zuo C, Tang X, Wang H, Su Q. A Review of the Effect of Defect Modulation on the Photocatalytic Reduction Performance of Carbon Dioxide. Molecules. 2024; 29(10):2308. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules29102308

Chicago/Turabian Style

Zuo, Cheng, Xiao Tang, Haiquan Wang, and Qian Su. 2024. "A Review of the Effect of Defect Modulation on the Photocatalytic Reduction Performance of Carbon Dioxide" Molecules 29, no. 10: 2308. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules29102308

Article Metrics

Back to TopTop