Next Article in Journal
Natural Products Chemistry: Advances in Synthetic, Analytical and Bioactivity Studies
Next Article in Special Issue
Recent Advances of Constructing Metal/Semiconductor Catalysts Designing for Photocatalytic CO2 Hydrogenation
Previous Article in Journal
Agricultural Solid Wastes Based Adsorbent Materials in the Remediation of Heavy Metal Ions from Water and Wastewater by Adsorption: A Review
Previous Article in Special Issue
Armchair Janus WSSe Nanotube Designed with Selenium Vacancy as a Promising Photocatalyst for CO2 Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Cu2O-Based Composites in Photocatalysis: Synthesis, Modification, and Applications

College of Chemistry & Chemical and Environmental Engineering, Weifang University, Weifang 261061, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(14), 5576; https://doi.org/10.3390/molecules28145576
Submission received: 10 July 2023 / Revised: 20 July 2023 / Accepted: 21 July 2023 / Published: 22 July 2023

Abstract

:
Photocatalysis technology has the advantages of being green, clean, and environmentally friendly, and has been widely used in CO2 reduction, hydrolytic hydrogen production, and the degradation of pollutants in water. Cu2O has the advantages of abundant reserves, a low cost, and environmental friendliness. Based on the narrow bandgap and strong visible light absorption ability of Cu2O, Cu2O-based composite materials show infinite development potential in photocatalysis. However, in practical large-scale applications, Cu2O-based composites still pose some urgent problems that need to be solved, such as the high composite rate of photogenerated carriers, and poor photocatalytic activity. This paper introduces a series of Cu2O-based composites, based on recent reports, including pure Cu2O and Cu2O hybrid materials. The modification strategies of photocatalysts, critical physical and chemical parameters of photocatalytic reactions, and the mechanism for the synergistic improvement of photocatalytic performance are investigated and explored. In addition, the application and photocatalytic performance of Cu2O-based photocatalysts in CO2 photoreduction, hydrogen production, and water pollution treatment are discussed and evaluated. Finally, the current challenges and development prospects are pointed out, to provide guidance in applying Cu2O-based catalysts in renewable energy utilization and environmental protection.

1. Introduction

With the development of industrialization, the use of fossil fuels in industry has caused many problems, such as carbon dioxide emissions causing global warming, water pollution, and the destruction of surrounding biological habitats. Energy shortages and environmental pollution pose a serious threat to the development of industry and agriculture, and have become hot topics that need to be addressed [1,2]. Photocatalytic technology utilizes semiconductor materials to achieve the photoreduction of CO2, the photocatalytic decomposition of water, and the degradation of pollutants, and has the advantages of a low cost, simple operation, and no secondary pollution [3,4,5]. Photocatalysis is a green technology that fully utilizes solar energy, and is considered one of the most feasible and promising methods to solve environmental and energy problems.
Since Fukushima et al. [6] discovered in 1967 that TiO2 can decompose water to produce hydrogen under light, tremendous progress has been made in photocatalytic technology. Due to its stable structure, high efficiency, low cost, nontoxicity, and high optical stability, TiO2 has become widely studied in the past few decades [7,8]. However, TiO2 can only absorb 3–5% of total ultraviolet light, so its utilization of sunlight is not high, significantly limiting its practical application under sunlight [9,10]. To effectively utilize the maximum proportion of visible light covering the solar spectrum (λ > 400 nm), in addition to modifying TiO2, researchers have studied a series of novel photocatalysts with a visible light response, such as simple oxides (ZnO [11] and Cu2O [12]), sulfides (CdS [13] and MoS2 [14]), Bi-based materials (Bi2WO6 [15] and BiVO4 [16]), and nitrides (C3N4 [17]).
Compared with other semiconductors, copper(I) oxide (Cu2O) has the advantages of nontoxicity, a favorable environmental acceptability, low cost, and high activity. It has been widely used in solar cells [18], carbon monoxide oxidation [19], photocatalysts [20], electrocatalysts [21], and sensors [22]. As a p-type semiconductor, Cu2O has a bandgap width of 2.17 eV, and a broad response range to the solar spectrum. A Cu2O material is currently one of the most promising visible light photocatalysts, and has become a research hotspot in photocatalysis. Li et al. [23] researched cubic c-Cu2O with the main exposure surface of (100), and tested its photocatalytic degradation performance on methyl orange (MO). It could completely decompose MO in an aqueous solution within 80 min under visible light, and almost remained unchanged in five consecutive cycles, showing a satisfactory stability. However, the application prospects of Cu2O in photocatalysis have been limited due to its poor stability, susceptibility to photocorrosion, and low quantum yield. To further improve the photocatalytic performance and stability of Cu2O, researchers have focused on a series of studies on morphology control, heteroatom doping, and the construction of semiconductor heterojunctions. With the increasing maturity in preparation and detection methods, the research on improving the photocatalytic activity of Cu2O has become more in-depth and diversified.
Unlike other reports in the literature [24,25,26], this paper reviews the preparation methods and applications of different Cu2O-based composites reported in recent years. By analyzing the roles of the different components in the photocatalytic process, we explain the reasons behind the improvement of photocatalytic performance, and point out the future direction that the industrial application of Cu2O-based composites could take.
According to data from the Web of Science platform by Clarivate Analytics (Figure 1), research on the subject of the photocatalysis of Cu2O and Cu2O-based materials is increasing year by year, indicating that Cu2O-based materials are becoming ideal candidates for a variety of energy and environmental photocatalysis applications. Based on the research direction of Cu2O photocatalysts, in this paper, the main preparation methods are introduced, including their merits and demerits, and the current main research focus and progress are reviewed. The research ideas and framework of this review are shown in Figure 2. The frequently used methods for improving their photocatalytic performance are reviewed, including morphology and size improvement, doping, metal loading, and semiconductor hybridization. In addition, several representative Cu2O-based composite photocatalysts are introduced, including metal/Cu2O composite, Cu2O semiconductor composite, Cu2O/carbon material composite, and ternary composite photocatalysts. Their applications in the photocatalytic reduction of CO2, photocatalytic hydrolysis of hydrogen, photocatalytic degradation of pollutants, and photocatalytic reduction of metal atoms are discussed through the combination of the experimental numbers and reaction principles. Finally, the future development of Cu2O-based composite photocatalysts is considered.

2. Synthetic Methods

The low utilization of visible light by single CuO materials, and the easy complexation of electron–hole pairs generated by CuO under photoexcitation limit the application of CuO materials in photocatalysis. There is currently more research on doping Cu2O with high Li or Na metal concentrations. The bonding network of the off-domain two- and three-electron centers is disrupted, effectively localizing the electrons in the limited space.

2.1. Preparation Methods

Different preparation methods to prepare effective Cu2O materials for photocatalytic experiments have been reported in the literature. The thermal oxidation of metals is a widely used method for synthesizing high-quality oxides. The final desired thickness of the Cu2O layer is prepared based on the oxidation of the high-purity copper foil. The temperature range is between 1000 and 1500 °C under a pure oxygen atmosphere or mixed gas atmosphere (e.g., Ar + O2). The obtained Cu2O is polycrystalline, with different grain structures depending on the chosen experimental conditions. In general, a mixture of CuO and Cu2O appears in copper foil at the end of oxidation. The Cu2O appears first, at the beginning of the oxidation process, while the CuO takes a long time to appear in the oxidation process. Electrodeposition is one of the methods for the production of high-quality Cu2O. The advantages of this method are that it is cheap, can efficiently work on different substrates, and allows for adjustments in the material properties and morphology according to the following parameters: the applied potential, current, and temperature, and the pH of the tank solution. The first electrochemical synthesis of Cu2O was by Stareck [27]. Subsequently, many other scholars developed different synthesis methods, using copper precursors, electrolytes, and electrochemistry. Two types of Cu2O nanoparticles were successfully synthesized by adjusting the pH of the electrolyte [28]. Firstly, the pH of the electrolyte was adjusted to 10, and a pyramid-shaped p-type Cu2O crystal was grown on the FTO substrate. Subsequently, the pH was adjusted to 4.9, and an ultra-thin layer of n-Cu2O deposition product was obtained (Figure 3). The p-Cu2O nanoparticles and the n-Cu2O protective layer on the surface formed a p/n heterojunction. The modified p/n-Cu2O had a bandgap of 2–2.2 eV, and could be excited by visible light with a wavelength less than 600 nm. Its photocurrent response was significantly improved, increasing the charge transfer rate and the stability of the catalyst. Thus, the modified p/n-Cu2O catalyst exhibited much higher activity than the original p-Cu2O.
Magnetron sputtering is a process that uses high-energy particles to bombard a solid target, so that atoms or molecules sputtered from the surface of the target form a thin film in a specific region. The CuO films prepared using this method exhibit nanometer-sized columnar structures, and the crystallinity, grain size, and film thickness of the Cu2O films can be controlled by varying the sputtering parameters (e.g., the sputtering power, oxygen content, oxygen concentration, sputter deposition time, and annealing temperature). Cu2O–CuO films with an excellent photocatalytic performance have been deposited on glass substrates using RF magnetron sputtering (Figure 4) [29]. It has been observed that with the prolongation of the sputtering deposition time, the size of the Cu2O–CuO nanoparticles has increased from 7 nm to 13 nm, and the thickness of the thin films from 7 nm to 50 nm, resulting in a rougher surface, reduced bandgap, and decreased PL strength. The results indicate that the structure, morphology, and optical and photocatalytic properties of prepared Cu2O–CuO films are strongly dependent on the deposition thickness. Under sunlight exposure, Cu2O–CuO films can completely degrade pollutants (methylene blue and methyl orange) from water within only 60 min.

2.2. Other Methods

In addition to the above methods, different surfactants [30,31,32,33,34,35] and micelles [36] have been used, mainly to control the morphology of the prepared Cu-based catalyst particles. Cu2O nanocrystals and nanoarray with cubic [37,38,39], octahedral [40], and multipod structures [41] have been prepared using these methods. Yang et al. [42] have proposed a metal-induced thermal reduction (MITR) method for the in-situ growth of Cu2O crystals on a copper substrate. The corresponding scheme is shown in Figure 5, and the operation is divided into two steps: (a) under alkaline conditions, the Cu(OH)2 nanorod array is in-situ grown by impregnating copper foil with a mixed solution of (NH4)2S2O8 and ammonia; and (b) the Cu(OH)2 on copper foil is directly thermally reduced to Cu2O nanorod array films in a N2 atmosphere at 500 °C. The average diameter of a nanorod was 400 ± 100 nm, with a length of several micrometers. The method is simple and efficient, and the preparation process has a low energy consumption and is controllable. In addition, the introduction of the substrate metal Cu can significantly reduce the reduction temperature, by changing the Gibbs free energy of the reaction. Surfactant-free synthesis has also been developed to reduce the interference of these surfactants [43,44,45]. Solvothermal [46,47] and sol–gel [48] methods have also been tested. The wet chemistry route [49,50], thermal evaporation [51,52], chemical vapor deposition [53,54], and hydrothermal route [55,56,57,58] are also common methods for synthesizing such semiconductors. In addition, the corresponding properties of Cu2O-based materials synthesized by different methods are detailed in the Section 4, including their morphology, structure, band gap, and photocatalytic applications.

3. Modification Strategies

Although noble metals have been used in photocatalytic organic waste degradation and CO2 reduction, their efficiency is still high. However, the cost is also high (e.g., Pt and Au), making them unsuitable for future industrial development. In contrast, Cu2O is inexpensive to use. It also has excellent CO2 capture ability and photochemical and structural properties, and shows unlimited development potential in CO2 reduction. However, the high electron–hole complexation rate and the low optical quantum efficiency limit the application of Cu2O in photocatalysis. To improve the photocatalytic efficiency of Cu2O, the structure of Cu2O needs to be modified. The modified structures are mainly divided into binary and ternary Cu2O heterostructure structures, and the addition of co-catalysts, in this section.

3.1. Binary Cu2O-Based Heterojunctions

3.1.1. Cu2O/Noble Metal Heterojunction

The Fermi energy level of the noble metal material is relatively low compared to that of the catalyst in the photocatalytic reduction of CO2, which has a higher work function than that of the catalyst. The mutual contact between the two will form a Schottky barrier at the metal–semiconductor interface, which can effectively inhibit the complexation of photoexcited electron–hole pairs, thus promoting the catalytic process, and improving the catalytic efficiency of the catalyst. The currently synthesized Cu2O/noble metal composites are Cu2O/Ag [59], Cu2O/Au [60], and Cu2O/Pt [61]. These materials show more than 90% photocatalytic efficiency for modified Cu2O.
Cu2O/Au nanostructures have been extensively investigated in recent years. Kuo et al. [62] reported the synthesis of Au@Cu2O core–shell nanocrystals using a chemical reduction method. The nanocrystals exhibited high activity in degrading methyl orange. Ag is relatively inexpensive, and has a higher electron transfer efficiency than metallic Au. Therefore, Ag/Cu2O catalysts have been more widely studied. Yang et al. [63] prepared Cu2O/Ag spherical microstructures by depositing silver nanoparticles on the surface of Cu2O through the thermal decomposition of silver acetate.

3.1.2. Cu2O/Graphene (GO) Heterojunction

From amorphous carbon black to crystalline structured natural layered graphite, and from zero-dimensional nanostructured fullerenes to two-dimensional structured graphene, carbon materials have been the most widely used and endlessly promising materials on earth. In recent decades, carbon nanomaterials have attracted much attention. The discovery of graphene self-assembled hydrogels with three-dimensional mesh structures has dramatically enriched the carbon material family, and provided a new growth point for new materials. Due to their unique nanostructure and properties, they have also shown significant scientific significance and experimental results. Thus, they provide a new target and direction when it comes to researching carbon-based materials. Graphene has been compounded with semiconductor photocatalysts, using its regular two-dimensional planar structure as a photocatalyst carrier. On the one hand, this could improve the dispersion of the catalyst. On the other hand, it could accelerate the photogenerated charge migration rate, and improve the photocatalytic activity of the composites.
Huang et al. [64] used the hydrothermal method to add graphene with the mass fractions of 0.1, 0.5, and 1 to Cu2O, which were noted as Cu2O/GO-0.1, Cu2O/GO-0.5, and Cu2O/GO-1, respectively (Figure 6). The experimental results showed that the highest hydrogen yield of Cu2O modified with graphene (118.3 mmol) was more than twice that of pure Cu2O (44.6 mmol). During the formation of Cu2O/GO composites, many negatively charged functional groups in graphene can recombine with positively charged copper ions by electrostatic adsorption, thus forming Cu2O/graphene composite structures directly during the reduction process. This principle has been used to synthesize cubic and octahedral Cu2O/GO composites. This structure could improve the efficiency of electron–hole separation. It could also improve the stability of the prepared catalysts. The experimental results showed that the cubic and octahedral Cu2O/GO composites degraded methyl orange with more than 90% efficiency. After six replicate tests, the efficiency remained above 70%, indicating that the prepared catalysts had excellent stability [65].
Graphene has properties such as the half-integer Hall effect, a unique quantum tunneling effect, and the bipolar electric field effect. In particular, its excellent electrical conductivity and huge specific surface area provide a feasible way to solve the bottleneck problem in the photocatalytic reaction of Cu2O-based composites.

3.2. Ternary Cu2O-Based Heterojunctions

In recent years, binary photocatalytic composites of Cu2O have achieved high achievements in the treatment of organic matter form wastewater and CO2 reduction. However, it will be a long time before binary photocatalytic composites can be used in society and daily life. Therefore, the development of ternary photocatalytic composites has become inevitable.
Yang et al. [66] used ternary Ag-CuO/GO as a photocatalytic material in the photocatalytic degradation of methyl orange, and the degradation efficiency of Ag-CuO/GO on the methyl orange was 90% after 60 min of visible light irradiation. Fu et al. [67] prepared TiO2-Ag-Cu2O composite catalysts for enhanced photocatalytic hydrogen production. The experimental results showed that the synergistic effect of Ag and Cu2O improved the photocatalytic efficiency of the reaction. In addition, the prepared composite catalysts had a double Z-scheme charge transfer pathway, which reduced the electron–hole complexation probability. The weak oxidation holes and weak reduction electrons in the charge transfer process were directly quenched, and the photogenerated carrier separation efficiency and catalyst reduction capacity were significantly enhanced.

3.3. Co-Catalyst Addition

In addition to constructing heterojunction structures, the photocatalytic efficiency can be improved by adding co-catalysts. Suitable co-catalysts are often present on the photocatalyst surface as active centers for oxidation or reduction, which can reduce the oxidation or reduction overpotential, and thus contribute to the photocatalytic reaction. In general, co-catalysts have three primary roles: (1) promoting the separation of the photoexcited electron–hole pairs; (2) inhibiting side reactions; and (3) improving the selectivity of the target products. Yu et al. [68] reported that adding the co-catalyst Cl to Cu2O nanorods led to a strong CO2 reduction ability. The experimental results showed that the addition of co-catalyst Cl mainly reduced its direct energy band, and also achieved an increase in the carrier density and conductivity. Zhang et al. [69] doped Zn in Cu2O microcubes, and the hydrogen production rate of Cu2O was six times higher than that of pure Cu2O when the Zn content was 0.1 wt.%. Kalubowila et al. [70] proposed a new method for introducing cocatalysts. They used ascorbic acid (AA) to reduce the prepared Cu2O/GO, where Cu2O was partially converted to Cu, and GO was fully converted to rGO. Cu nanoparticles with tens of nanometers have acted as co-catalysts in Cu2O/Cu/rGO composites, providing centers for effective charge transfer, and enhancing the performance of photocatalytic degradation.

4. Photocatalytic Applications

Semiconductor photocatalytic reactions are based on the solid energy band theory. Under the light, the available photogenerated electrons (e) and holes (h+) in the conduction band (CB) and valence band (VB) of the semiconductor migrate to the surface, to participate in the redox reaction. Therefore, the appropriate match between the CB/VB position of the photocatalyst and the redox potential determines whether the reaction can occur. In general, the CB position of the photocatalyst should be more negative to the reduction potential of the reaction, to promote the transfer of e from CB to the reactant; at the same time, the VB position should be corrected to the oxidation potential of the reaction, to ensure that holes can be transported from VB to the reactant. The bandgap of Cu2O-based materials is shown in Figure 7 [71]. It has been proven that they can be used as photocatalysts to achieve CO2 reduction (CO2RR), hydrogen production from water, pollutant degradation, and the reduction reaction of Cr. This section summarizes and discusses the latest progress in applications of Cu2O-based photocatalysts.

4.1. Photocatalytic CO2 Reduction

Photocatalytic technology can convert CO2 into CO and hydrocarbon fuels, achieving carbon recycling, and reducing greenhouse gas emissions. The application of Cu2O has been hampered largely by its inherent photocorrosion, ultra-fast charge recombination rate, and slow charge transport dynamics. In recent years, researchers have conducted and developed a series of novel Cu2O-based photocatalysts, making significant progress.
As is well known, semiconductors with different morphologies often expose different crystal faces, and exhibit varying photocatalytic activity. Celaya et al. [72] calculated by density-functional theory (DFT) that the (110) and (111) crystal faces of Cu2O have the potential of photocatalytic reduction of CO2 to produce hydrocarbon derivatives. To further determine the catalytic mechanism and active site, Wu and his colleagues [73] successfully prepared Cu2O nanocrystals with (110) and (100) crystal faces through colloidal synthesis, and carried out photocatalytic reactions using CO2 and H2O. gas chromatography–mass spectrometry (GC-MS), confirming that methanol was the only product of photoreduction, and the internal quantum yield was approximately 72%. In photocatalytic reactions, the (110) surface of a single Cu2O particle showed photocatalytic activity, while the (100) surface was inert. The electronic density of the Cu active site on the (110) surface moved from Cu (i) to Cu (ii), and the oxidation state of the Cu changed from Cu (ii) to Cu (i) after CO2 conversion under light. In 2022, Sahu et al. [74] synthesized and characterized Cu2O photocatalysts with cubic and truncated cubic structures. Their correspondingly exposed crystal faces were different (Figure 8). Due to the selective accumulation of e and h+ on different crystal planes, the photocatalytic activity in selectively reducing CO2 to methanol on cubic Cu2O with anisotropic {100} and {110} crystal planes was nearly 5.5 times higher than that on cubic Cu2O with only {100} crystal planes.
Meanwhile, researchers have adopted various modification methods to optimize the structure and performance of the photocatalyst. Element doping is a commonly used method to effectively change the physical properties of semiconductors, to improve their catalytic activity. Cl doping has been shown to optimize the catalytic activity of Cu2O [75]. At 400 nm, the apparent quantum yield (AQE) of Cl-doped Cu2O photocatalytic reduction of CO2 to CO and CH4 increased, with 1.13% and 1.07% for CO and CH4, respectively. The reason behind the enhanced performance of CO2RR was not only that the Cl doping optimized the energy band structure and conductivity of Cu2O, and improved the adsorption capacity of CO2 and the separation efficiency of the photogenerated carriers, but also that the Cl-doped Cu2O was conducive to the conversion of CO2 into the intermediates of *COOH, *CO, and *CH3O, thus improving the yield and selectivity of CO and CH4.
Constructing heterojunction structures is also an effective method for band reconstruction. In heterostructures, the internal electric field is formed at the contact interface of two or more semiconductors with the movement of the Fermi level, which drives the directional migration and separation of photogenerated electrons and holes. Common heterojunctions include the traditional (Type-Ⅰ, Ⅱ, and Ⅲ), p–n, Z-scheme, and S-scheme. The p–n heterojunction of Cu2O and n-type semiconductors can effectively delay the recombination of photogenerated carriers, and promote electron transfer [76]. The yield of the photocatalytic reduction of CO2 to CH3OH from the Cu2O/TiO2 heterojunction after 6 h of UV–Vis irradiation has been 21.0–70.6 μmol/gcat. At the p–n heterojunction, the photogenerated electrons and holes are separated and transferred to the CB/VB with lower potential energy, respectively, resulting in a redox ability closer to the lower of the two semiconductors. The Z-scheme heterostructure solves this problem perfectly. Electrons and holes in the CB/VB with lower energy recombine, and cancel each other out in the Z-scheme heterojunction, thus retaining the higher conduction and valence band values in the two semiconductors, and enhancing the redox ability of the photocatalyst. For example, the Ag-Cu2O/ZnO nanorods (NRs) reported by Zhang and his team showed an enhanced photocatalytic CO2 reduction performance [77]. Under UV–vis light, the yield of CO significantly increased, which was seven times higher that of pure ZnO or Cu2O NRs. The results showed that the deposited Cu2O can enhance the chemical adsorption of CO2 on the catalyst surface, and the Z-scheme charge transfer pathway formed between the ZnO and Cu2O can promote effective charge separation, thereby improving the photocatalysis performance.
Due to the small bandgap energy and high conduction band value of Cu2O-based materials, the products of photocatalytic CO2RR are complex, mainly including CO and various organic compounds (CH4, CH3OH, HCOOH). According to the different reaction products, the application of Cu2O-based materials in photocatalytic CO2 reduction is summarized in Table 1.

4.2. Photocatalytic H2 Production

Hydrogen energy is abundant and renewable, which can effectively avoid energy exhaustion, and the products of hydrogen energy combustion will not cause pollution. Photocatalytic hydrogen production has the advantages of high efficiency, low cost, and environmental friendliness, and has great potential in high-efficiency hydrogen evolution. Common semiconductor photocatalysts (such as TiO2, ZnO, and g-C3N4.) have the disadvantage of a low utilization of sunlight, and the photocatalytic hydrogen evolution efficiency is not ideal. In recent years, Cu2O has become a research hotspot in photocatalytic hydrogen evolution because of its excellent photoresponsiveness. However, the poor charge separation ability of pure Cu2O lowers its hydrogen evolution performance. It is essential to modify and adjust Cu2O-based catalysts to meet the practical need to increase the hydrogen production yield.
Hybridizing Cu2O with other semiconductor materials to construct heterojunctions can achieve the effective separation of photo-induced charge carriers, which is an effective method to enhance photocatalytic activity, and has been validated in numerous studies on photocatalytic hydrogen production. NiFe2O4/Cu2O with different mass percentages has been synthesized by impregnation and thermal annealing methods to construct p–n heterojunctions [94]. The photocatalytic hydrogen production rate of all heterojunctions was significantly higher than that of the original material. The 50/50 mass ratio was the most effective, and the hydrogen production rate within 24 h was 102.4 mmol∙g−1, while NiFe2O4 and Cu2O only obtained 1.35 and 0.85 mmol∙g−1, respectively. The increase in activity came from the enhanced charge separation at the heterojunction, which increased the concentration of charge carriers (Figure 9). Cu2O/CaTiO3 series samples were synthesized using the hydrothermal method and NaBH4 reduction treatment [95]. The photocatalytic hydrogen production effect of the 50Ca10Cu sample was the best (8.268 mmol∙g−1·h−1), about 344.5 times that of the CaTiO3 sample. It also exhibited perfect stability after multiple cyclic tests.
The above p–n heterojunctions are typical type-Ⅱ heterojunctions, which often impair the redox capacity of photogenerated electrons and holes. Researchers have recently designed and constructed Z-scheme and step-scheme (S-scheme) heterojunctions for photocatalytic hydrogen production. For example, dendritic branched Cu2O was synthesized hydrothermally, and Cu2O/TiO2 composites were prepared via surface charge modulation [96]. The hydrogen production rate of the optimized CT-70 (Cu2O coupled with 70 wt.% TiO2) photocatalyst reached 14.020 mmol−1 within six hours, which was 264 and 44 times higher than that of pure Cu2O and TiO2, respectively. The electron transfer mechanism of the Z-scheme was proposed and verified via DFT calculation and EPR analysis. Under simulated sunlight, photoexcited electrons migrate from the CB of TiO2 to the VB of Cu2O, and then recombine with photogenerated holes in the VB of Cu2O, thereby retaining highly reducing electrons and highly oxidizing holes (Figure 10). Therefore, under the conditions of sensitive photosensitivity and the effective separation of photogenerated electrons and holes, the performance of photocatalysts in hydrogen evolution under visible light is significantly improved. The S-scheme heterojunction photocatalyst has a similar efficient carrier separation performance and enhanced redox capacity. Cu2O/g-C3N4 composites were successfully synthesized using a simple wet chemical method, and applied in the field of photocatalytic energy production. Cu2O/g-C3N4 series samples showed high catalytic activity. In particular, 1-Cu2O/g-C3N4 showed the highest hydrogen evolution rate of 480.6 μmol∙g−1·h−1 under visible light irradiation, 12.0 times that of the original Cu2O sample. Based on the analysis of the experimental and simulation results, the ideal catalytic performance of the Cu2O/g-C3N4 photocatalyst was derived from the efficient interfacial charge separation and transfer of the S-scheme heterostructure [97].
Furthermore, photocorrosion is currently an urgent problem for Cu2O photocatalysts, and finding effective strategies to suppress photocorrosion in photocatalysts is still an enormous challenge. To overcome this challenge, Liu et al. [98] proposed a core–shell model: the Cu2O/PyTTA-TPA COF nanocube photocatalyst was constructed using an energy level matching the Cu2O and 2D PyTTA-TPA COF. It exhibited an excellent photocatalytic hydrogen evolution rate of 12.5 mmol∙g−1·h−1, approximately 8.0 and 20.0 times higher than the PyTTA TPA COF and Cu2O, respectively. Most importantly, under the protection of the stable PyTTATPA-COF shell, the Cu2O nanocube core was protected from photocorrosion, and did not show noticeable morphological or crystal structure changes after 1000 light excitations, thus significantly improving the photocorrosion resistance stability of the catalyst. Table 2 shows the recently reported Cu2O-based materials for photocatalytic hydrogen production.

4.3. Photocatalytic Degradation of Pollutants

With the rapid development of the global economy, industrial and agricultural waste is produced in large quantities, and continues to enter the environment. Many organic pollutants also enter the environment, and some show persistent pollution, which is difficult to remove through microbial action and hydrolysis. The long-term existence and accumulation of refractory pollutants leads to environmental pollution and ecological imbalance, and even threatens human survival and development. Research and development around pollutant degradation technology are critical. Photocatalytic technology has shown promising prospects for treating refractory pollutants, such as the photocatalytic processes that mineralize organic pollutants into water and CO2, and which essentially eliminate secondary pollution, rather than concentrating these pollutants and their by-products into the waste stream. In the past few decades, extensive research has been conducted on Cu2O-based photocatalysts to purify the environment. Table 3 summarizes the recent reports of Cu2O-based photocatalysts in pollutant degradation.
Among all the types of pollutants, organic dyes have become an important source of water pollution. As refractory organic pollutants, dyes cause severe damage to human health and the ecological balance. Traditional adsorption methods only transfer toxic organic molecules to the solid surface, without eliminating them, and still run the risk of desorption. MBC@Cu2O composites have been prepared by loading porous spherical Cu2O onto wood biochar carriers, with a liquid-phase synthesis strategy, at room temperature [109]. As a bi-functional adsorption-based photocatalytic composite, MBC@Cu2O showed great potential in removing anionic dye methyl orange (MO) from water. Under visible light irradiation, the photocatalytic degradation efficiency of MO reached 94.5%, and remained above 80% after five cycles. In another work, Sehrawat and his team prepared MoS2/Cu2O composites with different weight ratios via precipitation, using MoS2 nanosheets and Cu2O nanospheres [110]. The photocatalytic degradation of indigo carmine (IC) dye was carried out under simulated visible light. Compared to the original MoS2, the optimized MC-3 sample showed the best degradation performance, with a degradation rate of 99.59% for IC within 90 min, and no significant change in performance after five cycles. Experiments regarding the capture of active species showed that the photocatalytic reaction relied on the production of the superoxide radical (•O2), and further verified the Z-scheme mechanism of the MoS2/Cu2O photocatalyst. In the same year, Li et al. synthesized the core–shell WO3-Cu2O Z-scheme heterojunction via hydrothermal and electrochemical deposition methods for the photocatalytic degradation of methylene blue (MB) under visible light [111]. The Cu2O nanoparticles deposited on the surface of WO3 enhanced the visible light absorption ability. The Z-scheme heterojunction achieved the effective spatial separation of the charges, and retained the strong redox ability of the photogenerated electrons and holes. The WO3-Cu2O-120s photocatalyst showed the highest reaction rate, almost twice that of the original WO3.
As a typical persistent organic pollutant, antibiotics are difficult to degrade and remove, due to their low biodegradability, which has become a thorny problem in water pollution control. Research has shown that the defect states and vacancies caused by element doping significantly impact the catalytic performance of semiconductor materials. Doping semiconductor functional materials with specific elements provides a feasible way to overcome the obstacles in applications for photocatalytic degradation. Nie et al. synthesized Cl-doped Cu2O microcrystals using a simple hydrothermal method, and used them to treat levofloxacin contaminants (LVX) under mild reaction conditions [112]. Compared with other reaction systems, the synthesis of Cl-doped Cu2O has a higher degradation efficiency for levofloxacin. After 240 min of photocatalytic reaction, the maximum degradation rate of LVX was 85.8% and 80.3% after eight cycles, indicating the stability and reusability of the photocatalyst. Based on the theoretical calculation and test results, it can be concluded that introducing hybrid orbitals and oxygen vacancy defects into Cu2O crystal cells by doping Cl reduces the band gap of Cu2O, resulting in a red shift in the absorption edge. Compared with pure Cu2O microcrystals, the prepared Cl-doped Cu2O single crystals with oxygen vacancy had a narrower band gap, and higher photogenerated electron–hole separation and transport efficiency. Considering the close relationship between the morphology and electronic structure, surface energy, and chemical reactivity of nanocrystals, it is of great significance to explore the influence of the morphology/exposed crystal surface of Cu2O on the synthesis process and the photocatalytic performance. Wu et al. developed a series of Cu2O@HKUST-1 core–shell structures via self-constrained strategies, using Cu2O nanocrystals with different morphologies as templates [113]. The characterization results indicated that the (111) surface of Cu2O was more favorable for the growth of HKUST-1 than the (100) surface. Comparing the photocatalytic degradation performance of tetracycline hydrochloride (TC-HCl), it was found that Cu2O@HKUST-1 had the best photocatalytic performance among the three types of composite material, with a degradation efficiency of 95.35% for TC-HCl. It was attributed to the excellent photoresponse, and the most effective interfacial charge transfer and separation in the Cu2O@HKUST-1 cubes.
In addition to organic dyes and antibiotics, solar-powered Cu2O-based photocatalysts can degrade heavy metal pollutants in wastewater, mainly toxic hexavalent chromium (Cr (Ⅵ)). Xiong et al. [114] constructed a Cu2O/LDH photocatalyst by grafting Cu2O-NP, and embedding it into the LDH host layer through an in-situ reduction strategy. CuZnTi LDH is valuable in two aspects: (a) as a source of Cu2O, and (b) as a support bracket to avoid the self-oxidation of Cu2O-NPs. The optimized photocatalyst showed a high degradation efficiency for difficult-to-degrade pollutants under visible light conditions, with a reduction rate of 95.5% for Cr (Ⅵ) by Cu2O/LDH0.10, and a degradation rate of 71.6% for TC. The excellent photocatalytic efficiency was attributed to the charge transfer mechanism of the Cu2O/ZnTiLDH p–n heterojunction, effectively promoting the separation and migration of the photogenerated electron–hole. Recently, Zhu et al. [115] used the Si and Cu of waste serpentine tailings and WPCB to prepare low-cost waste-based Cu-Cu2O/SiO2 photocatalysts. Due to the dispersion of Cu-Cu2O3 on the surface of the SiO2 carrier, the composite material obtained a higher specific surface area. The photocatalytic reduction of Cr (Ⅵ) using waste-based catalysts was the best at a loading rate of 9% Cu and 7g∙L−1 SiO2, and the photocatalytic activity decreased by only 4.93% after five cycles. The mechanism of Cr (Ⅵ) reduction by the waste Cu-Cu2O/SiO2 photocatalyst is to excite the waste Cu2O to produce photoelectron–hole pairs. The electrons in the waste group Cu2O CB reduce Cr (Ⅵ) adsorbed on the surface to Cr (Ⅲ), and the surface Cu drives the electrons to the surface of the Cu metal, without returning the waste group Cu2O.
Moreover, the accumulation in soil and water of herbicides, insecticides used in the agriculture and food industries, and phenolic compounds emitted from industry, such as petrochemicals and pharmaceuticals, can have significant harmful effects on humans and aquaculture systems. The use of metal oxide photocatalysts has been proven to be an effective, low-cost, and green method for treating such wastewater. In 2021, Alp [116] successfully synthesized hybrid Cu2O-Cu cubes by reducing D(+)-glucose in an alkaline solution using a one-step aqueous solution synthesis method, without any toxic reagents or surfactants. The Cu2O-Cu exhibited excellent photocatalytic properties for dyes and herbicides, due to the effective separation of photogenerated electron–holes and the enhanced charge transfer mechanism at heterojunctions. In particular, when dealing with 2,4-Dichlorophenoxyacetic acid (2,4-D), one of the widely used herbicides in agriculture and urban landscaping, the degradation effect of the Cu2O-Cu heterojunction was outstanding. It photodegraded all of the 2,4-D in the medium within 40 min, while the original Cu2O cube photodegraded 85% within 60 min. In the same year, Mkhalid et al. [117] prepared a Cu2O photocatalyst loaded with Cu nanoparticles via sol–gel and photo-assisted deposition technology. The structure and optical and photoelectric properties of the prepared photocatalyst were improved by adjusting the Cu content. The results showed that the band gap of the Cu2O loaded with 15% Cu was reduced to 1.95 eV, significantly enhancing the visible light absorption ability. The optimized Cu@Cu2O photocatalyst completely photodecomposed atrazine (AZ, a commonly used triazine herbicide) within 30 min, and demonstrated excellent durability. In recent years, effectively solving the problem of phenolic pollutants in livable environments has also been a major challenge faced by humanity, and has received a high level of attention from many researchers. A low-cost but highly efficient phosphate-doped carbon/Cu2O composite (HKUST-1-P-300) was reported by Dubai et al. [118]. The catalyst was derived from the modification of HKUST-1 with triphenylphosphine and conditioned calcination. Under visible light irradiation, the degradation efficiency of HKUST-1-P-300 for phenol was 99.8%, the hydrogen evolution rate was 1208 µmol, and the external quantum efficiency was 48.6% (at 425 nm) within 90 min, and the high performance could still be maintained after four cycles. Mechanism studies showed that the excellent photocatalytic activity of HKUST-1-P-300 came from multiple synergistic effects: an enhanced visible light absorption efficiency, a larger surface area, the effective separation of photogenerated carriers, a reduced aggregation of Cu2O, and the P-doped carbon/Cu2O structure. These novel Cu2O-based materials, as highly efficient photocatalysts, have potential applications in removing environmental pollutants, and generating clean energy, to promote sustainable environmental construction.
Table 3. Recently reported Cu2O-based materials for the photocatalytic removal of pollutants.
Table 3. Recently reported Cu2O-based materials for the photocatalytic removal of pollutants.
PhotocatalystSynthesis MethodMorphology and StructureSizeBandgap (Eg)Light ResourceTarge Pollutant/Concentration/VolumeEfficiencyCycleRefs.
Ag-Cu2OElectrochemical deposition and redox reactionComposite film2.02 eV500 W halogen lampMB/30 mg∙L−1/50 mL92%3[119]
Cr-doped Cu2OHydrothermal methodOctahedrons800–1200 nm2.06 eV500 W tungsten halogen lamp (400–1100 nm)LVX/40 mg∙L−1/50 mL79.6–72.4%1–8[120]
BiOCl/Cu2OSolvothermal methodSpherical shape3–5 μm2.00 eV500 W Xenon lampMoxifloxacin/20 mg∙L−1/50 mL72.3%5[121]
C-dots/Cu2O/SrTiO3Hydrothermal and two-step methodChocolate ball with sesame on the surface~2.16 μmEgSrTiO3: 3.19 eV;
EgCu2O: 2.10 eV
500 W Xenon lamp (λ > 420 nm)CTC.HCl/15 mg∙L−1/50 mL92.6%4[122]
CuO-Cu2OChemical–thermal oxidationNanorods60 nm1.90 eV150 W metal halide lamp (λ > 400 nm)MB/5 mg∙L−1/50 mL80%3[123]
Cotton fabrics/Cu2O-NCImpregnation and HH reductionOctahedron Cu2O attached to cotton fibers20–40 nm of diameter of Cu2OEgCu2O: 2.20 eV350 W Xenon lamp (λ > 400 nm)MB/200 ppm/200 mL98.32–85%1–5[124]
Cu2O@HKUST-1In-situ converted strategyOctahedron structureEgCu2O: 1.95 eV;
EgHKUST-1: 2.59 eV
Tungsten lamp (>420 nm, 500 W)TC-HCl/20 mg∙L−1/100 mL93.40–90.02%1–4[125]
Fe3O4/Cu2O-AgSolvothermal and liquid deposition methodsDouble six peak structure~5 nm2.23 eVPAHs/5 mg∙L−1/100 mL95–90%1–8[126]
Cu2O/ZnO@PETElectroless template depositionRectangular-shaped~13 ± 4.5 nm3.2–3.4 eVUltra-Vitalux 300WCzm/1.0 mg∙L−1/100 mL98–26%1–6[127]
Cu2O-Au-TiO2Two-step photocatalytic depositionCore–shell structure~50 nm1.4–1.7 eVXenon lamp (λ > 422 nm)Cr(Ⅵ)/10 mg∙L−1/50 mL100% (3h)3[128]
Cu2O/N-CQD/ZIF-8Reduction precipitationSpherical structure~80–100 nm2.6 eV,300 W Xenon lamp (λ > 420 nm)Cr(Ⅵ)/20 mg∙L−1/50 mL98.99–97.13%1–5[129]
Cu2O/rGO/BiOBrTwo-step strategyHierarchical microspheres500 nm–1 μmEgBiOBr: 2.7 eV;
EgCu2O: 1.9 eV
300 W Xenon lamp (λ ≥ 420 nm)Cr(Ⅵ)/20 mg∙L−1/50 mL100% (40 min)5[130]
Cu-TiO2-Cu2OPhotodepositionThe triple junction structure~20 nm300 W Xenon lamp (200–2400 nm)2,4,5-T/50 ppm/100 mL93%3[131]
Ag-Cu2O/rGOTwo-step reduction processSpherical AgNPs deposited on the Cu2O situated on the surface of rGO sheets~60 nm60 W tungsten filament lamp (500–700 nm, 0.24 W/cm2)MO/40 mg∙L−1/50 mL
Phenol/20 mg∙L−1/50 mL
90% (60min);
Rate constant of phenol degradation: 0.09732
3[132]

4.4. DFT Study Applied in the Photocatalysis

At the present time, there are fewer studies revealing the reaction mechanism of Cu2O through DFT simulations. Moreover, the catalytic microstructure and mechanism of Cu2O-based composites are still unclear. Designing Cu2O-based photocatalysts, and investigating the mechanism of improving photocatalytic activity at the molecular level require the introduction of theoretical calculations. In future studies, DFT simulations and experiments are needed, to reveal the relationship between the establishment of the microstructure and the catalytic activity of the photocatalysts, which will provide the theoretical basis for future photocatalytic industrial applications.
Lv et al. [133] analyzed the electronic structure and photocatalytic properties of Cu2O doped with different contents of Mn, using first-principle calculations. The simulation results showed that the visible light absorption intensity and photocatalytic efficiency were enhanced with the increase in doping concentration, and varied with the doping configuration, compared to pure Cu2O. The enhanced light absorption was mainly attributed to the in-band leaps of the electrons in the three-dimensional state of Mn. The enhancement of light absorption was mainly due to the in-band leaps of electrons in the three-dimensional state of Mn, which gave the semiconductor material certain metallic properties, and increased the absorbance of the visible light. Therefore, Cu2O applied to the future industrialization of photocatalysis could be doped with a small amount of Mn in the semiconductor, to improve the photocatalytic efficiency.

5. Conclusions

In recent years, the practical photocatalytic applications of Cu2O-based materials in scientific fields such as solar energy conversion and environmental remediation have attracted great interest. As a transition metal oxide, Cu2O has the advantages of a narrow band gap, strong visible light response, suitable conduction band position, low cost, and great potential as a photocatalyst. This paper introduces the basic properties, synthesis methods, and modification strategies of Cu2O-based materials. Recently reported Cu2O-based photocatalysts and their recent advances in photocatalysis, such as photocatalytic CO2 reduction, photocatalytic hydrogen production, and pollutant degradation, are reviewed. However, the research on Cu2O-based materials is still in its early stages, and there is room for improvement in their photocatalytic performance.
  • Currently, most Cu2O-based composites and sacrificial agents are synthesized from noble metal materials, which have high costs and significantly limit their large-scale applications. The development of non-precious metal catalysts, such as graphene, is vital to future development. More importantly, the catalytic efficiency of most Cu2O-based composites is very low, and the catalytic performance needs to be improved to meet the requirements of practical applications.
  • Although many experimental studies on the photocatalysis of Cu2O-based composites are introduced in this paper, these works are still in their infancy. In addition, the large-scale production of high-quality Cu2O-based photocatalysts faces numerous difficulties, considering the secondary hazards of nanomaterials. Therefore, it is urgent that we further study the photocatalytic mechanism of Cu2O-based composites from the above perspectives, and promote the industrial application process of Cu2O-based composite catalysts.
  • The photocorrosion of Cu2O still deserves attention. Although the current method of constructing heterojunctions to suppress photocorrosion has achieved certain results, the photocorrosion phenomenon of Cu2O still exists, and affects its long-term use. Establishing a core–shell structure is a good governance measure but, when synthesizing photocatalysts, it is necessary to carefully handle the thickness of the shell layer, to ensure sufficient absorption of light by the Cu2O.
  • The structure of the catalyst determines the catalytic activity, while the catalytic microstructure and mechanism of Cu2O-based composites is still unclear. Theoretical calculations should be introduced when designing a Cu2O-based photocatalyst, and studying the mechanism of improving photocatalytic activity at the molecular level. In future research, DFT simulations and experiments are needed, to reveal the relationship between the establishment of the microstructure and the catalytic activity of photocatalysts.

Author Contributions

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

Funding

Financial support in 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. Ju, L.; Tang, X.; Zhang, Y.; Li, X.; Cui, X.; Yang, G. Single Selenium Atomic Vacancy Enabled Efficient Visible-Light-Response Photocatalytic NO Reduction to NH3 on Janus WSSe Monolayer. Molecules 2023, 28, 2959. [Google Scholar] [CrossRef]
  2. Duic, N.; Guzovic, Z.; Vyatcheslav, K.; Klemes, J.J.; Mathiessen, B.V.; Yan, J.Y. Sustainable development of energy, water and environment systems. Appl. Energ. 2013, 101, 3–5. [Google Scholar] [CrossRef]
  3. Ju, L.; Tang, X.; Li, J.; Dong, H.; Yang, S.; Gao, Y.; Liu, W. Armchair Janus WSSe Nanotube Designed with Selenium Vacancy as a Promising Photocatalyst for CO2 Reduction. Molecules 2023, 28, 4602. [Google Scholar] [CrossRef] [PubMed]
  4. 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]
  5. Albero, J.; Peng, Y.; Garcia, H. Photocatalytic CO2 reduction to C2+ products. ACS Catal. 2020, 10, 5734–5749. [Google Scholar] [CrossRef]
  6. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  7. Henderson, M.A. A surface science perspective on TiO2 photocatalysis. Surf. Sci. Ref. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  8. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photoch. Photobio. C 2012, 13, 169–189. [Google Scholar] [CrossRef]
  9. Ma, D.G.; Zhai, S.; Wang, Y.; Liu, A.A.; Chen, C.C. TiO2 photocatalysis for transfer hydrogenation. Molecules 2019, 24, 330. [Google Scholar] [CrossRef] [Green Version]
  10. Guo, Q.; Zhou, C.Y.; Ma, Z.B.; Yang, X.M. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef] [PubMed]
  11. Goktas, S.; Goktas, A. A comparative study on recent progress in efficient ZnO based nanocomposite and heterojunction photocatalysts: A review. J. Alloy Compd. 2021, 863, 158734. [Google Scholar] [CrossRef]
  12. Liu, X.Q.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.H.; Zhao, S.Q.; Li, Z.; Lin, Z.Q. Noble metal-metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  13. Liu, Y.P.; Shen, S.J.; Zhang, J.T.; Zhong, W.W.; Huang, X.H. Cu2−xSe/CdS composite photocatalyst with enhanced visible light photocatalysis activity. Appl. Surf. Sci. 2019, 478, 762–769. [Google Scholar] [CrossRef]
  14. Sun, Y.B.; Xiao, J.T.; Huang, X.S.; Mei, P.; Wang, H.H. Boosting photocatalytic efficiency of MoS2/CdS by modulating morphology. Environ. Sci. Pollut. Res. 2022, 29, 73282–73291. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, M.X.; Jin, H.M.; Wu, Z.K.; Guo, Y.N.; Shang, Q.K. Selective photocatalytic oxidation of aromatic alcohols using B-g-C3N4/Bi2WO6 composites. Sep. Purif. Technol. 2023, 317, 123915. [Google Scholar] [CrossRef]
  16. Qin, N.B.; Zhang, S.F.; He, J.Y.; Long, F.; Wang, L.L. In situ synthesis of BiVO4/BiOBr microsphere heterojunction with enhanced photocatalytic performance. J. Alloy Compd. 2022, 927, 166661. [Google Scholar] [CrossRef]
  17. Jiang, J.Z.; Xiong, Z.G.; Wang, H.T.; Liao, G.D.; Bai, S.S.; Zou, J.; Wu, P.X.; Zhang, P.; Li, X. Sulfur-doped g-C3N4/g-C3N4 isotype step-scheme heterojunction for photocatalytic H2 evolution. J. Mater. Sci. Technol. 2022, 118, 15–24. [Google Scholar] [CrossRef]
  18. Sekkat, A.; Bellet, D.; Chichignoud, G.; Munoz-Rojas, D.; Kaminski-Cachopo, A. Unveiling key limitations of ZnO/Cu2O all-oxide solar cells through numerical simulations. ACS Appl. Energy Mater. 2022, 5, 5423–5433. [Google Scholar] [CrossRef]
  19. Wu, L.K.; Ma, P.D.; Zhang, C.H.; Yi, X.K.; Hao, Q.L.; Dou, B.J.; Bin, F. Effects of Cu2O morphology on the performance of CO self-sustained catalytic combustion. Appl. Catal. A-Gen. 2023, 652, 119034. [Google Scholar] [CrossRef]
  20. Li, J.W.; Sun, Z.L.; He, M.Z.; Gao, D.; Li, Y.T.; Ma, J.J. Simple synthesis of Ag nanoparticles/Cu2O cube photocatalyst at room temperature: Efficient electron transfer improves photocatalytic performance. Inorg. Chem. Commun. 2022, 138, 109200. [Google Scholar] [CrossRef]
  21. Yang, X.; Cheng, J.; Yang, X.; Xu, Y.; Sun, W.F.; Zhou, J.H. MOF-derived Cu@Cu2O heterogeneous electrocatalyst with moderate intermediates adsorption for highly selective reduction of CO2 to methanol. Chem. Eng. J. 2022, 431, 134171. [Google Scholar] [CrossRef]
  22. Wang, N.; Tao, W.; Gong, X.Q.; Zhao, L.P.; Wang, T.S.; Zhao, L.J.; Liu, F.M.; Liu, X.M.; Sun, P.; Lu, G.Y. Highly sensitive and selective NO2 gas sensor fabricated from Cu2O-CuO microflowers. Sens. Actuators B-Chem. 2022, 362, 131803. [Google Scholar] [CrossRef]
  23. Li, J.W.; He, M.Z.; Yan, J.K.; Liu, J.H.; Zhang, J.X.; Ma, J.J. Room temperature engineering crystal facet of Cu2O for photocatalytic degradation of methyl orange. Nanomaterials 2022, 12, 1697. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Y.H.; Liu, M.M.; Chen, J.L.; Fang, S.M.; Zhou, P.P. Recent advances in Cu2O-based composites for photocatalysis: A review. Dalton Trans. 2021, 50, 4091–4111. [Google Scholar] [CrossRef] [PubMed]
  25. Li, C.F.; Guo, R.T.; Zhang, Z.R.; Wu, T.; Pan, W.G. Converting CO2 into Value-Added Products by Cu2O-Based Catalysts: From Photocatalysis, Electrocatalysis to Photoelectrocatalysis. Small 2023, 19, 2207875. [Google Scholar] [CrossRef]
  26. Yu, W.B.; Yi, M.; Fu, H.H.; Pei, M.J.; Liu, Y.; Xu, B.M.; Zhang, J. Dandelion-Like Nanostructured Cu/Cu2O Heterojunctions with Fast Diffusion Channels Enabling Rapid Photocatalytic Pollutant Removal. ACS Appl. Nano Mater. 2023, 6, 2928–2941. [Google Scholar] [CrossRef]
  27. Stareck, J.E. Decorating Metals. U.S. Patents 2,081,121A, 18 May 1937. [Google Scholar]
  28. Zhou, Q.Q.; Chen, Y.X.; Shi, H.Y.; Chen, R.; Ji, M.H.; Li, K.X.; Wang, H.L.; Jiang, X.; Lu, C.Z. The construction of p/n-Cu2O heterojunction catalysts for efficient CO2 photoelectric reduction. Catalysts 2023, 13, 857. [Google Scholar] [CrossRef]
  29. Sahu, K.; Bisht, A.; Khan, S.A.; Sulania, I.; Singhal, R.; Pandey, A.; Mohapatra, S. Thickness dependent optical, structural, morphological, photocatalytic and catalytic properties of radio frequency magnetron sputtered nanostructured Cu2O-CuO thin films. Ceram. Int. 2020, 46, 14902–14912. [Google Scholar] [CrossRef]
  30. Huang, L.; Peng, F.; Yu, H.; Wang, H. Synthesis of Cu2O nanoboxes, nanocubes and nanospheres by polyol process and their adsorption characteristic. Mater. Res. Bull. 2008, 43, 3047–3053. [Google Scholar] [CrossRef]
  31. Ma, L.L.; Li, J.L.; Sun, H.Z.; Qiu, M.Q.; Wang, J.B.; Chen, J.Y.; Yu, Y. Self-assembled Cu2O flowerlike architecture: Polyol synthesis, photocatalytic activity and stability under simulated solar light. Mater. Res. Bull. 2010, 45, 961–968. [Google Scholar] [CrossRef]
  32. Zhu, J.; Wang, Y.; Wang, X.; Yang, X.; Lu, L. A convenient method for preparing shape-controlled nanocrystalline Cu2O in a polyol or water/polyol system. Powder Technol. 2008, 181, 249–254. [Google Scholar] [CrossRef]
  33. Huang, X.W.; Liu, Z.J.; Zheng, Y.F. Synthesis of Cu2O nanobelts via surfactant-assisted polyol method. Chin. Chem. Lett. 2011, 22, 879–882. [Google Scholar] [CrossRef]
  34. Bai, Y.; Yang, T.; Gu, Q.; Cheng, G.; Zheng, R. Shape control mechanism of cuprous oxide nanoparticles in aqueous colloidal solutions. Powder Technol. 2012, 227, 35–42. [Google Scholar] [CrossRef]
  35. Shin, H.S.; Song, J.Y.; Yu, J. Template-assisted electrochemical synthesis of cuprous oxide nanowires. Mater. Lett. 2009, 63, 397–399. [Google Scholar] [CrossRef]
  36. Dodoo-Arhin, D.; Leoni, M.; Scardi, P.; Garnier, E.; Mittiga, A. Synthesis, characterisation and stability of Cu2O nanoparticles produced via reverse micelles microemulsion. Mater. Chem. Phys. 2010, 122, 602–608. [Google Scholar] [CrossRef]
  37. Wang, Y.Q.; Liang, W.S.; Satti, A.; Nikitin, K. Fabrication and microstructure of Cu2O nanocubes. J. Cryst. Growth 2010, 312, 1605–1609. [Google Scholar] [CrossRef]
  38. Zhang, H.; Cui, Z. Solution-phase synthesis of smaller cuprous oxide nanocubes. Mater. Res. Bull. 2008, 43, 1583–1589. [Google Scholar] [CrossRef]
  39. Zhu, J.; Bi, H.; Wang, Y.; Wang, X.; Yang, X.; Lu, L. Solution-phase synthesis of Cu2O cubes using CuO as a precursor. Mater. Lett. 2008, 62, 2081–2083. [Google Scholar] [CrossRef]
  40. Zhang, X.; Cui, Z. One-pot growth of Cu2O concave octahedron microcrystal in alkaline solution. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2009, 162, 82–86. [Google Scholar] [CrossRef]
  41. Liang, Z.H.; Zhu, Y.J. Synthesis of uniformly sized Cu2O crystals with star-like and flower-like morphologies. Mater. Lett. 2005, 59, 2423–2425. [Google Scholar] [CrossRef]
  42. Yang, Y.M.; Wang, K.; Yang, Z.H.; Zhang, Y.M.; Gu, H.Y.; Zhang, W.X.; Li, E.R.; Zhou, C. An efficient route to Cu2O nanorod array film for high-performance Li-ion batteries. Thin Solid Films 2016, 608, 79–87. [Google Scholar] [CrossRef]
  43. Luo, Y.; Tu, Y.; Ren, Q.; Dai, X.; Xing, L.; Li, J. Surfactant-free fabrication of Cu2O nanosheets from Cu colloids and their tunable optical properties. J. Solid State Chem. 2009, 182, 182–186. [Google Scholar] [CrossRef]
  44. Wang, H.; He, S.; Yu, S.; Shi, T.; Jiang, S. Template-free synthesis of Cu2O hollow nanospheres and their conversion into Cuhollow nanospheres. Powder Technol. 2009, 193, 182–186. [Google Scholar] [CrossRef]
  45. Sui, Y.; Zhang, Y.; Fu, W.; Yang, H.; Zhao, Q.; Sun, P.; Ma, D.; Yuan, M.; Li, Y.; Zou, G. Low-temperature template-free synthesis of Cu2O hollow spheres. J. Cryst. Growth 2009, 311, 2285–2290. [Google Scholar] [CrossRef]
  46. Wei, M.; Huo, J. Preparation of Cu2O nanorods by a simple solvothermal method. Mater. Chem. Phys. 2010, 121, 291–294. [Google Scholar] [CrossRef]
  47. Wei, M.; Lun, N.; Ma, X.; Wen, S. A simple solvothermal reduction route to copper and cuprous oxide. Mater. Lett. 2007, 61, 2147–2150. [Google Scholar] [CrossRef]
  48. Lim, Y.F.; Chua, C.S.; Lee, C.J.J.; Chi, D. Sol-gel deposited Cu2O and CuO thin films for photocatalytic water splitting. Phys. Chem. Chem. Phys. 2014, 16, 25928–25934. [Google Scholar] [CrossRef]
  49. Qu, Y.; Li, X.; Chen, G.; Zhang, H.; Chen, Y. Synthesis of Cu2O nano-whiskers by a novel wet-chemical route. Mater. Lett. 2008, 62, 886–888. [Google Scholar] [CrossRef]
  50. Susman, M.D.; Feldman, Y.; Vaskevich, A.; Rubinstein, I. Chemical Deposition of Cu2O Nanocrystals with Precise Morphology Control. ACS Nano 2014, 8, 162–174. [Google Scholar] [CrossRef] [PubMed]
  51. Balamurugan, B.; Mehta, B.R. Optical and structural properties of nanocrystalline copper oxide thin films prepared by activated reactive evaporation. Thin Solid Films 2001, 396, 90–96. [Google Scholar] [CrossRef]
  52. Al-Kuhaili, M.F. Characterization of copper oxide thin films deposited by the thermal evaporation of cuprous oxide (Cu2O). Vacuum 2008, 82, 623–629. [Google Scholar] [CrossRef]
  53. Barreca, D.; Comini, E.; Gasparotto, A.; Maccato, C.; Sada, C.; Sberveglieri, G.; Tondello, E. Chemical vapor deposition of copper oxide films and entangled quasi-1D nanoarchitectures as innovative gas sensors. Sens. Actuators B Chem. 2009, 141, 270–275. [Google Scholar] [CrossRef]
  54. Gomersall, D.E.; Flewitt, A.J. Plasma enhanced chemical vapor deposition of p-type Cu2O from metal organic precursors. J. Appl. Phys. 2022, 131, 215301. [Google Scholar] [CrossRef]
  55. Wang, S.; Zhang, X.; Pan, L.; Zhao, F.-M.; Zou, J.-J.; Zhang, T.; Wang, L. Controllable sonochemical synthesis of Cu2O/Cu2(OH)3NO3 composites toward synergy of adsorption and photocatalysis. Appl. Catal. B Environ. 2015, 164, 234–240. [Google Scholar] [CrossRef]
  56. Ma, D.; Liu, H.; Yang, H.; Fu, W.; Zhang, Y.; Yuan, M.; Sun, P.; Zhou, X. High pressure hydrothermal synthesis of cuprous oxide microstructures of novel morphologies. Mater. Chem. Phys. 2009, 116, 458–463. [Google Scholar] [CrossRef]
  57. Valodkar, M.; Pal, A.; Thakore, S. Synthesis and characterization of cuprous oxide dendrites: New simplified green hydrothermal route. J. Alloys Compd. 2011, 509, 523–528. [Google Scholar] [CrossRef]
  58. Togashi, T.; Hitaka, H.; Ohara, S.; Naka, T.; Takami, S.; Adschiri, T. Controlled reduction of Cu2+ to Cu+ with an N,O-type chelate under hydrothermal conditions to produce Cu2O nanoparticles. Mater. Lett. 2010, 64, 1049–1051. [Google Scholar] [CrossRef]
  59. Li, J.; Sun, L.; Yan, Y. One-step in-situ fabrication of silver-modified Cu2O crystals with enhanced visible photocatalytic activity. Micro-Nano Lett. 2016, 11, 363–365. [Google Scholar] [CrossRef]
  60. Wang, B.; Li, R.; Zhang, Z.; Zhang, W.; Yan, X.; Wu, X.; Cheng, G.; Zheng, R. Novel Au/CuO Multi-shelled Porous heterostructures forenhanced efficiency photoelectrochemical water splitting. J. Mater. Chem. A 2017, 5, 14415–14421. [Google Scholar] [CrossRef]
  61. Jin, J.Y.; Mei, H.; Wu, H.M.; Wang, S.F.; Xia, Q.H.; Ding, Y. Selective detection of dopamine based on Cu2O@Pt core-shell nanoparticles modified electrode in the presence of ascorbic acid and uric acid. J. Alloy Compd. 2016, 689, 174–181. [Google Scholar] [CrossRef]
  62. Kuo, M.Y.; Hsiao, C.F.; Chiu, Y.H.; Lai, T.H.; Fang, M.J.; Wu, J.Y.; Chen, J.W.; Wu, C.L.; Wei, K.H.; Lin, H.C.; et al. Au@Cu2O core@shell nanocrystals as dual-functional catalysts for sustainable environmental applications. Appl. Catal. B 2019, 242, 499–506. [Google Scholar] [CrossRef]
  63. Yang, Z.; Ma, C.; Wang, W.; Zhang, M.; Hao, X.S. Chen. Long-term antibacterial stable reduced graphene oxide nanocomposites loaded with cuprous oxide nanoparticles. J. Colloid Interface Sci. 2019, 557, 156–167. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, Y.; Yan, C.F.; Guo, C.Q.; Lu, Z.X.; Shi, Y.; Wang, Z.D. Synthesis of GO-modified Cu2O nanosphere and the photocatalytic mechanism of water splitting for hydrogen production. Int. J. Hydrogen Energy 2016, 7, 4007–4016. [Google Scholar] [CrossRef] [Green Version]
  65. 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]
  66. Yang, L.; Luo, S.; Li, Y. High Efficient Photocatalytic Degradation of p-Nitrophenol on a Unique Cu2O/TiO2 p-n Heterojunction Network Catalyst. Environ. Sci. technol. 2010, 44, 7641–7646. [Google Scholar] [CrossRef]
  67. Fu, J.; Cao, S.; Yu, J. Dual Z-scheme charge transfer in TiO2-Ag-Cu2O composite for enhanced photocatalytic hydrogen generation. J. Mater. 2015, 1, 124–133. [Google Scholar] [CrossRef] [Green Version]
  68. Yu, L.; Ba, X.; Qiu, M.; Li, Y.; Shuai, L.; Zhang, W.; Ren, Z.; Yu, Y. Visible-light driven CO2 reduction coupled with water oxidation on Cl-doped Cu2O nanorods. Nano Energy 2019, 60, 576–582. [Google Scholar] [CrossRef]
  69. Zhang, L.Z.; Jing, D.W.; Guo, L.J.; Yao, X.D. In situ photochemical synthesis of Zn-Doped Cu2O hollow microcubes for high efficient photocatalytic H2 production. ACS Sustain. Chem. Eng. 2014, 2, 1446–1452. [Google Scholar] [CrossRef]
  70. Kalubowila, K.D.R.N.; Gunewardene, M.S.; Jayasingha, J.L.K.; Dissanayake, D.; Jayathilaka, C.; Jayasundara, J.M.D.; Gao, Y.; Jayanetti, J.K.D.S. Reduction-induced synthesis of reduced graphene oxide-wrapped Cu2O/Cu nanoparticles for photodegradation of methylene blue. ACS Appl. Nano Mater. 2021, 4, 2673–2681. [Google Scholar] [CrossRef]
  71. Wang, W.L.; Wang, L.; Su, W.; Xing, Y. Photocatalytic CO2 reduction over copper-based materials: A review. J. CO2 Util. 2022, 61, 102056. [Google Scholar] [CrossRef]
  72. Celaya, C.A.; Delesma, C.; Torres-Arellano, S.; Sebastian, P.J.; Muniz, J. Understanding CO2 conversion into hydrocarbons via a photoreductive process supported on the Cu2O(100), (110) and (111) surface facets: A first principles study. Fuel 2021, 306, 121643. [Google Scholar] [CrossRef]
  73. Wu, Y.A.; McNulty, I.; Liu, C.; Lau, K.C.; Liu, Q.; Paulikas, A.P.; Sun, C.J.; Cai, Z.H.; Guest, J.R.; Ren, Y.; et al. Facet-dependent active sites of a single Cu2O particle photocatalyst for CO2 reduction to methanol. Nat. Energy 2019, 4, 957–968. [Google Scholar] [CrossRef]
  74. Sahu, A.K.; Pokhriyal, M.; Upadhyayula, S.; Zhao, X.S. Modulating charge carrier dynamics among anisotropic crystal facets of Cu2O for enhanced CO2 photoreduction. J. Phys. Chem. C 2022, 126, 13094–13104. [Google Scholar] [CrossRef]
  75. Zhao, Z.Y.; Yi, J.; Zhou, D.C. Electronic structures of halogen-doped Cu2O based on DFT calculations. Chinese. Phys. B. 2014, 23, 017401. [Google Scholar] [CrossRef]
  76. Cheng, S.P.; Wei, L.W.; Wang, H.P. Photocatalytic reduction of CO2 to methanol by Cu2O/TiO2 heterojunctions. Sustainability 2022, 14, 374. [Google Scholar] [CrossRef]
  77. Zhang, F.; Li, Y.H.; Qi, M.Y.; Tang, Z.R.; Xu, Y.J. Boosting the activity and stability of Ag-Cu2O/ZnO nanorods for photocatalytic CO2 reduction. Appl. Catal. B-Environ. 2020, 268, 118380. [Google Scholar] [CrossRef]
  78. Song, Y.Y.; Zhao, X.J.; Feng, X.Y.; Chen, L.M.; Yuan, T.C.; Zhang, F.Q. Z-scheme Cu2O/Cu/Cu3V2O7(OH)2·2H2O heterostructures for efficient visible-light photocatalytic CO2 reduction. ACS Appl. Energy Mater. 2022, 5, 10542–10552. [Google Scholar] [CrossRef]
  79. Cui, L.K.; Hu, L.Q.; Shen, Q.Q.; Liu, X.G.; Jia, H.S.; Xue, J.B. Three-dimensional porous Cu2O with dendrite for efficient photocatalytic reduction of CO2 under visible light. Appl. Catal. B-Environ. 2022, 581, 152343. [Google Scholar] [CrossRef]
  80. Zheng, Y.K.; Duan, Z.T.; Liang, R.X.; Lv, R.Q.; Wang, C.; Zhang, Z.X.; Wan, S.L.; Wang, S.; Xiong, H.F.; Ngaw, C.K. Shape-dependent performance of Cu/Cu2O for photocatalytic reduction of CO2. ChemSusChem 2022, 15, e202200216. [Google Scholar] [CrossRef]
  81. Zhang, X.J.; Zhao, X.; Chen, K.; Fan, Y.Y.; Wei, S.L.; Zhang, W.S.; Han, D.X.; Niu, L. Palladium-modified cuprous(i) oxide with {100} facets for photocatalytic CO2 reduction. Nanoscale 2021, 13, 2883–2890. [Google Scholar] [CrossRef]
  82. Zhao, X.X.; Sun, L.L.; Jin, X.; Xu, M.Y.; Yin, S.K.; Li, J.Z.; Li, X.; Shen, D.; Yan, Y.; Huo, P.W. Cu media constructed Z-scheme heterojunction of UiO-66-NH2/Cu2O/Cu for enhanced photocatalytic induction of CO2. Appl. Surf. Sci. 2021, 545, 148967. [Google Scholar] [CrossRef]
  83. Deng, Y.; Wan, C.A.; Li, C.; Wang, Y.Y.; Mu, X.Y.; Liu, W.; Huang, Y.P.; Wong, P.K.; Ye, L.Q. Synergy effect between facet and zero-valent copper for selectivity photocatalytic methane formation from CO2. ACS Catal. 2022, 12, 4526–4533. [Google Scholar] [CrossRef]
  84. Tang, Z.L.; He, W.J.; Wang, Y.L.; Wei, Y.C.; Yu, X.L.; Xiong, J.; Wang, X.; Zhang, X.; Zhao, Z.; Liu, J. Ternary heterojunction in rGO-coated Ag/Cu2O catalysts for boosting selective photocatalytic CO2 reduction into CH4. Appl. Catal. B-Environ. 2022, 311, 121371. [Google Scholar] [CrossRef]
  85. Zhou, J.Q.; Li, Y.F.; Yu, L.; Li, Z.P.; Xie, D.F.; Zhao, Y.Y.; Yu, Y. Facile in situ fabrication of Cu2O@Cu metal-semiconductor heterostructured nanorods for efficient visible-light driven CO2 reduction. Chem. Eng. J. 2020, 385, 123940. [Google Scholar] [CrossRef]
  86. Ali, S.; Lee, J.; Kim, H.; Hwang, Y.; Razzaq, A.; Jung, J.W.; Cho, C.H.; In, S.I. Sustained, photocatalytic CO2 reduction to CH4 in a continuous flow reactor by earth-abundant materials: Reduced titania-Cu2O Z-scheme heterostructures. Appl. Catal. B-Environ. 2020, 279, 119344. [Google Scholar] [CrossRef]
  87. Yao, S.; Sun, B.Q.; Zhang, P.; Tian, Z.Y.; Yin, H.Q.; Zhang, Z.M. Anchoring ultrafine Cu2O nanocluster on PCN for CO2 photoreduction in water vapor with much improved stability. Appl. Catal. B-Environ. 2022, 317, 121702. [Google Scholar] [CrossRef]
  88. Movahed, S.K.; Najinasab, A.; Nikbakht, R.; Dabiri, M. Visible light assisted photocatalytic reduction of CO2 to methanol using Fe3O4@N-C/Cu2O nanostructure photocatalyst. J. Photochem. Photobiol. A 2020, 401, 112763. [Google Scholar] [CrossRef]
  89. Liu, S.H.; Lu, J.S.; Pu, Y.C.; Fan, H.C. Enhanced photoreduction of CO2 into methanol by facet-dependent Cu2O/reduce graphene oxide. J. CO2 Util. 2019, 33, 171–178. [Google Scholar] [CrossRef]
  90. Li, H.T.; Deng, Y.D.; Liu, Y.D.; Zeng, X.; Wiley, D.; Huang, J. Carbon quantum dots and carbon layer double protected cuprous oxide for efficient visible light CO2 reduction. Chem. Commun. 2019, 55, 4419–4422. [Google Scholar] [CrossRef]
  91. Zeng, Z.P.; Yan, Y.B.; Chen, J.; Zan, P.; Tian, Q.H.; Chen, P. Boosting the photocatalytic ability of Cu2O nanowires for CO2 conversion by mxene quantum dots. Adv. Funct. Mater. 2018, 29, 1806500. [Google Scholar] [CrossRef]
  92. Wang, H.; Cheng, S.J.; Cai, X.; Cheng, L.H.; Zhou, R.J.; Hou, T.T.; Li, Y.W. Photocatalytic CO2 reduction to HCOOH over core-shell Cu@Cu2O catalysts. Catal. Commun. 2022, 162, 106372. [Google Scholar] [CrossRef]
  93. Zhu, Q.; Cao, Y.N.; Tao, Y.; Li, T.; Zhang, Y.; Shang, H.; Song, J.X.; Li, G.S. CO2 reduction to formic acid via NH2-C@Cu2O photocatalyst in situ derived from amino modified Cu-MOF. J. CO2 Util. 2021, 54, 101781. [Google Scholar] [CrossRef]
  94. Dominguez-Arvizu, J.L.; Jimenez-Miramontes, J.A.; Hernandez-Majalca, B.C.; Valenzuela-Castro, G.E.; Gaxiola-Cebreros, F.A.; Salinas-Gutierrez, J.M.; Collins-Martinez, V.; Lopez-Ortiz, A. Study of NiFe2O4/Cu2O p-n heterojunctions for hydrogen production by photocatalytic water splitting with visible light. J. Mater. Res. Technol. 2023, 21, 4184–4199. [Google Scholar] [CrossRef]
  95. Yang, J.F.; Yang, H.Y.; Dong, Y.H.; Cui, H.; Sun, H.; Yin, S.Y. Fabrication of Cu2O/MTiO3 (M = Ca, Sr and Ba) p-n heterojunction for highly enhanced photocatalytic hydrogen generation. J. Alloy Compd. 2023, 930, 167333. [Google Scholar] [CrossRef]
  96. Zhang, Y.H.; Liu, M.M.; Chen, J.L.; Xie, K.F.; Fang, S.M. Dendritic branching Z-scheme Cu2O/TiO2 heterostructure photocatalysts for boosting H2 production. J. Phys. Chem. Solids 2021, 152, 109948. [Google Scholar] [CrossRef]
  97. Dai, B.L.; Li, Y.Y.; Xu, J.M.; Sun, C.; Li, S.J.; Zhao, W. Photocatalytic oxidation of tetracycline, reduction of hexavalent chromium and hydrogen evolution by Cu2O/g-C3N4 S-scheme photocatalyst: Performance and mechanism insight. Appl. Surf. Sci. 2022, 592, 153309. [Google Scholar] [CrossRef]
  98. Liu, Y.X.; Tan, H.; Wei, Y.A.; Liu, M.H.; Hong, J.X.; Gao, W.Q.; Zhao, S.Q.; Zhang, S.P.; Guo, S.J. Cu2O/2D COFs core/shell nanocubes with antiphotocorrosion ability for efficient evolution. ACS Nano 2023, 17, 5994–6001. [Google Scholar] [CrossRef]
  99. Muscetta, M.; Al Jitan, S.; Palmisano, G.; Andreozzi, R.; Marotta, R.; Cimino, S.; Di Somma, I. Visible light-driven photocatalytic hydrogen production using Cu2O/TiO2 composites prepared by facile mechanochemical synthesis. J. Environ. Chem. Eng. 2022, 10, 107735. [Google Scholar] [CrossRef]
  100. Chang, Y.C.; Syu, S.Y.; Lu, M.Y. Fabrication of In(OH)3-In2S3-Cu2O nanofiber for highly efficient photocatalytic hydrogen evolution under blue light LED excitation. Int. J. Hydrogen Energy 2023, 48, 9318–9332. [Google Scholar] [CrossRef]
  101. Lai, T.H.; Tsao, C.W.; Fang, M.J.; Wu, J.Y.; Chang, Y.P.; Chiu, Y.H.; Hsieh, P.Y.; Kuo, M.Y.; Chang, K.D.; Hsu, Y.J. Au@Cu2O core-shell and Au@Cu2Se yolk-shell nanocrystals as promising photocatalysts in photoelectrochemical water splitting and photocatalytic hydrogen production. ACS Appl. Mater. Interfaces 2022, 14, 40771–40783. [Google Scholar] [CrossRef]
  102. Ibarra-Rodriguez, L.I.; Huerta-Flores, A.M.; Torres-Martinez, L.M. Development of Na2Ti6O13/CuO/Cu2O heterostructures for solar photocatalytic production of low-carbon fuels. Mater. Res. Bull. 2020, 122, 110679. [Google Scholar] [CrossRef]
  103. Dubale, A.A.; Ahmed, I.N.; Zhang, Y.J.; Yang, X.L.; Xie, M.H. A facile strategy for fabricating C@Cu2O/CuO composite for efficient photochemical hydrogen production with high external quantum efficiency. Appl. Surf. Sci. 2020, 534, 147582. [Google Scholar] [CrossRef]
  104. Fan, Z.B.; Zhang, X.J.; Li, Y.J.; Guo, X.; Jin, Z.L. Construct 3D NiCo-LDH/Cu2O p-n heterojunction via electrostatic self-assembly for enhanced photocatalytic hydrogen evolution. J. Ind. Eng. Chem. 2022, 110, 491–502. [Google Scholar] [CrossRef]
  105. Mohite, S.V.; Kim, S.; Lee, C.S.; Bae, J.; Kim, Y. Z-scheme heterojunction photocatalyst: Deep eutectic solvents-assisted synthesis of Cu2O nanocluster improved hydrogen production of TiO2. J. Alloy Compd. 2022, 928, 167168. [Google Scholar] [CrossRef]
  106. Qiu, P.; Xiong, J.Y.; Lu, M.J.; Liu, L.J.; Li, W.; Wen, Z.P.; Li, W.J.; Chen, R.; Cheng, G. Integrated p-n/Schottky junctions for efficient photocatalytic hydrogen evolution upon Cu@TiO2-Cu2O ternary hybrids with steering charge transfer. J. Colloid Interface Sci. 2022, 622, 924–937. [Google Scholar] [CrossRef] [PubMed]
  107. Becerra-Paniagua, D.K.; Torres-Arellano, S.; Martinez-Alonso, C.; Luevano-Hipolito, E.; Sebastian, P.J. Facile and green synthesis of Cu/Cu2O composite for photocatalytic H2 generation. Mater. Sci. Semicond. Process. 2023, 162, 107485. [Google Scholar] [CrossRef]
  108. Zhu, H.; Xi, M.Y.; Huang, G.P.; Qin, L.X.; Zhang, T.Y.; Kang, S.Z.; Li, X.Q. Cuprous oxide core-shell heterostructure facilely encapsulated by cadmium metal organic frameworks for enhanced photocatalytic hydrogen generation. J. Phys. Chem. Solids 2023, 181, 111476. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Li, X.J.; Chen, J.F.; Wang, Y.A.; Cheng, Z.Y.; Chen, X.Q.; Gao, X.; Guo, M.H. Porous spherical Cu2O supported by wood-based biochar skeleton for the adsorption-photocatalytic degradation of methyl orange. Appl. Surf. Sci. 2023, 611, 155744. [Google Scholar] [CrossRef]
  110. Sehrawat, P.; Rana, S.; Mehta, S.K.; Kansal, S.K. Optimal synthesis of MoS2/Cu2O nanocomposite to enhance photocatalytic performance towards indigo carmine dye degradation. Appl. Surf. Sci. 2022, 604, 154482. [Google Scholar] [CrossRef]
  111. Li, J.J.; Guo, C.P.; Li, L.H.; Gu, Y.J.; BoK-Hee, K.; Huang, J.L. Construction of Z-scheme WO3-Cu2O nanorods array heterojunction for efficient photocatalytic degradation of methylene blue. Inorg. Chem. Commun. 2022, 138, 109248. [Google Scholar] [CrossRef]
  112. Nie, J.K.; Yu, X.J.; Liu, Z.B.; Zhang, J.; Ma, Y.; Chen, Y.Y.; Ji, Q.G.; Zhao, N.N.; Chang, Z. Energy band reconstruction mechanism of Cl-doped Cu2O and photocatalytic degradation pathway for levofloxacin. J. Clean. Prod. 2022, 363, 132593. [Google Scholar] [CrossRef]
  113. Wu, Y.; Li, Y.Q.; Li, H.; Guo, H.; Yang, Q.; Li, X.M. Tunning heterostructures interface of Cu2O@HKUST-1 for enhanced photocatalytic degradation of tetracycline hydrochloride. Sep. Purif. Technil. 2022, 303, 122106. [Google Scholar] [CrossRef]
  114. Xiong, J.; Zeng, H.Y.; Peng, J.F.; Peng, D.Y.; Liu, F.Y.; Xu, S.; Yang, Z.L. abrication of Cu2O/ZnTi-LDH p-n heterostructure by grafting Cu2O NPs onto the LDH host layers from Cu-doped ZnTi-LDH and insight into the photocatalytic mechanism. Compos. Part. B-Eng. 2023, 250, 110447. [Google Scholar] [CrossRef]
  115. Zhu, P.; Li, Y.; Ma, Y.; Ruan, X.X.; Zhang, Q.Z. Preparation of waste-based Cu-Cu2O/SiO2 photocatalyst from serpentine tailings and waste printed circuit boards and photoreduction of Cr(Ⅵ). Ceram. Int. 2023, 49, 12518–12528. [Google Scholar] [CrossRef]
  116. Alp, E. The facile synthesis of Cu2O-Cu hybrid cubes as efficient visible-light-driven photocatalysts for water remediation processes. Powder Technol. 2021, 394, 1111–1120. [Google Scholar] [CrossRef]
  117. Mkhalid, I.A.; Shawky, A. Cu-supported Cu2O nanoparticles: Optimized photodeposition enhances the visible light photodestruction of atrazine. J. Alloy Compd. 2021, 853, 157040. [Google Scholar] [CrossRef]
  118. Dubale, A.A.; Ahmed, I.N.; Chen, X.H.; Ding, C.; Hou, G.H.; Guan, R.F.; Meng, X.M.; Yang, X.L.; Xie, M.H. A highly stable metal-organic framework derived phosphorus doped carbon/Cu2O structure for efficient photocatalytic phenol degradation and hydrogen production. J. Mater. Chem. A 2019, 7, 6062–6079. [Google Scholar] [CrossRef]
  119. Yu, X.J.; Zhang, J.; Chen, Y.Y.; Ji, Q.G.; Wei, Y.C.; Niu, J.N.; Yu, Z.; Yao, B.H. Ag-Cu2O composite films with enhanced photocatalytic activities for methylene blue degradation: Analysis of the mechanism and the degradation pathways. J. Environ. Chem. Eng. 2021, 9, 106161. [Google Scholar] [CrossRef]
  120. Nie, J.K.; Yu, X.J.; Liu, Z.B.; Wei, Y.C.; Zhang, J.; Zhao, N.N.; Yu, Z.; Yao, B.H. Boosting principles for the photocatalytic performance of Cr-doped Cu2O crystallites and mechanisms of photocatalytic oxidation for levofloxacin. Appl. Surf. Sci. 2021, 576, 151842. [Google Scholar] [CrossRef]
  121. Liu, Z.B.; Yu, X.J.; Gao, P.H.; Nie, J.K.; Yang, F.; Guo, B.Q.; Zhang, J. Preparation of BiOCl/Cu2O composite particles and its photocatalytic degradation of moxifloxacin. Opt. Mater. 2022, 128, 112432. [Google Scholar] [CrossRef]
  122. Zhang, Y.Y.; Li, Y.; Ruan, Z.H.; Yuan, Y.; Lin, K.F. Extensive solar light utilizing by ternary C-dots/Cu2O/SrTiO3: Highly enhanced photocatalytic degradation of antibiotics and inactivation of E. coli. Chemosphere 2022, 290, 133340. [Google Scholar] [CrossRef]
  123. Bayat, F.; Sheibani, S. Enhancement of photocatalytic activity of CuO-Cu2O heterostructures through the controlled content of Cu2O. Mater. Res. Bull. 2021, 145, 111561. [Google Scholar] [CrossRef]
  124. Su, X.P.; Chen, W.; Han, Y.N.; Wang, D.C.; Yao, J.M. In-situ synthesis of Cu2O on cotton fibers with antibacterial properties and reusable photocatalytic degradation of dyes. Appl. Surf. Sci. 2021, 536, 147945. [Google Scholar] [CrossRef]
  125. Wu, Y.; Li, X.M.; Zhao, H.; Yao, F.B.; Cao, J.; Chen, Z.; Wang, D.B.; Yang, Q. Core-shell structured Cu2O@HKUST-1 heterojunction photocatalyst with robust stability for highly efficient tetracycline hydrochloride degradation under visible light. Chem. Eng. J. 2021, 426, 131255. [Google Scholar] [CrossRef]
  126. Huang, J.; Zhou, T.X.; Zhao, W.S.; Cui, S.C.; Guo, R.; Li, D.; Kadasala, N.R.; Han, D.L.; Jiang, Y.H.; Liu, Y. Multifunctional magnetic Fe3O4/Cu2O-Ag nanocomposites with high sensitivity for SERS detection and efficient visible light-driven photocatalytic degradation of polycyclic aromatic hydrocarbons (PAHs). J. Colloid Interf. Sci. 2022, 628, 315–326. [Google Scholar] [CrossRef]
  127. Altynbaeva, L.S.; Barsbay, M.; Aimanova, N.A.; Jakupova, Z.Y.; Nurpeisova, D.T.; Zdorovets, M.V.; Mashentseva, A.A. A Novel Cu2O/ZnO@PET Composite Membrane for the Photocatalytic Degradation of Carbendazim. Nanomaterials 2022, 12, 1724. [Google Scholar] [CrossRef]
  128. Yanagida, S.; Yajima, T.; Takei, T.; Kumada, N. Removal of hexavalent chromium from water by Z-scheme photocatalysis using TiO2 (rutile) nanorods loaded with Au core-Cu2O shell particles. J. Environ. Sci. 2022, 115, 173–189. [Google Scholar] [CrossRef] [PubMed]
  129. Qiang, T.T.; Wang, S.T.; Ren, L.F.; Gao, X.D. Novel 3D Cu2O/N-CQD/ZIF-8 composite photocatalyst with Z-scheme heterojunction for the efficient photocatalytic reduction of Cr(VI). J. Environ. Chem. Eng. 2022, 10, 108784. [Google Scholar] [CrossRef]
  130. Ma, J.; Liang, C.J.; Yu, C.J.; Li, H.M.; Xu, H.; Hua, Y.J.; Wang, C.T. BiOBr microspheres anchored with Cu2O nanoparticles and rGO: A Z-scheme heterojunction photocatalyst for efficient reduction of Cr(VI) under visible light irradiation. Appl. Surf. Sci. 2023, 609, 155247. [Google Scholar] [CrossRef]
  131. An, X.Q.; Liu, H.J.; Qu, J.H.; Moniz, S.J.A.; Tang, J.W. Photocatalytic mineralisation of herbicide 2,4,5-trichlorophenoxyacetic acid: Enhanced performance by triple junction Cu-TiO2-Cu2O and the underlying reaction mechanism. New J. Chem. 2015, 39, 314–320. [Google Scholar] [CrossRef] [Green Version]
  132. Sharma, K.; Maiti, K.; Kim, N.H.; Hui, D.; Lee, J.H. Green synthesis of glucose-reduced graphene oxide supported Ag-Cu2O nanocomposites for the enhanced visible-light photocatalytic activity. Compos. Part. B-Eng. 2018, 138, 35–44. [Google Scholar] [CrossRef]
  133. Lv, Q.Y.; Li, L.L.; Li, Y.F.; Mao, J.H.; Chen, T.; Shao, D.W.; Li, M.M.; Tan, R.S.; Zhao, J.Q.; Shi, S.H. A DFT Study of Electronic Structures and Photocatalytic Properties of Mn-Cu2O. Russ. J. Phys. Chem. A 2020, 94, 641–646. [Google Scholar]
Figure 1. The annual number of publications using “Cu2O” as a topic keyword since 2010 (data taken from Web of Science on 1 January 2023).
Figure 1. The annual number of publications using “Cu2O” as a topic keyword since 2010 (data taken from Web of Science on 1 January 2023).
Molecules 28 05576 g001
Figure 2. The research framework and basis of thinking for this review.
Figure 2. The research framework and basis of thinking for this review.
Molecules 28 05576 g002
Figure 3. Schematic diagram of the procedure for the electrodeposition of p/n Cu2O on the FTO substrate [28].
Figure 3. Schematic diagram of the procedure for the electrodeposition of p/n Cu2O on the FTO substrate [28].
Molecules 28 05576 g003
Figure 4. Schematic diagram of a Cu2O–CuO film prepared using the magnetron sputtering method [29]. Copyright 2020, Elsevier.
Figure 4. Schematic diagram of a Cu2O–CuO film prepared using the magnetron sputtering method [29]. Copyright 2020, Elsevier.
Molecules 28 05576 g004
Figure 5. The schematic illustration of the synthesis process of Cu2O nanorod array films [42]. Copyright 2016, Elsevier.
Figure 5. The schematic illustration of the synthesis process of Cu2O nanorod array films [42]. Copyright 2016, Elsevier.
Molecules 28 05576 g005
Figure 6. SEM images of (a) pure Cu2O, (b) GO/Cu2O-0.1, (c) GO/Cu2O-0.5, and (d) GO/Cu2O-1, and (e) a TEM image of GO/Cu2O-0.5 [64]. Copyright 2017, Elsevier.
Figure 6. SEM images of (a) pure Cu2O, (b) GO/Cu2O-0.1, (c) GO/Cu2O-0.5, and (d) GO/Cu2O-1, and (e) a TEM image of GO/Cu2O-0.5 [64]. Copyright 2017, Elsevier.
Molecules 28 05576 g006
Figure 7. Diagram of the bandgap of copper-oxide-based photocatalysts [71]. Copyright 2022, Elsevier.
Figure 7. Diagram of the bandgap of copper-oxide-based photocatalysts [71]. Copyright 2022, Elsevier.
Molecules 28 05576 g007
Figure 8. TEM images of (a,b) cubic Cu2O and (c,d) edge-truncated cubic Cu2O; simulated images of (e) 3D structure of edge-truncated cubic Cu2O and (f) 2D crystal orientation [74]. Copyright 2022 American Chemical Society.
Figure 8. TEM images of (a,b) cubic Cu2O and (c,d) edge-truncated cubic Cu2O; simulated images of (e) 3D structure of edge-truncated cubic Cu2O and (f) 2D crystal orientation [74]. Copyright 2022 American Chemical Society.
Molecules 28 05576 g008
Figure 9. The p–n heterostructures in the NiFe2O4/Cu2O photocatalyst [94].
Figure 9. The p–n heterostructures in the NiFe2O4/Cu2O photocatalyst [94].
Molecules 28 05576 g009
Figure 10. (a) Type-Ⅱ and (b) Z-scheme electron transfer mechanism in Cu2O/TiO2 photocatalyst [96]. Copyright 2021, Elsevier.
Figure 10. (a) Type-Ⅱ and (b) Z-scheme electron transfer mechanism in Cu2O/TiO2 photocatalyst [96]. Copyright 2021, Elsevier.
Molecules 28 05576 g010
Table 1. The application of Cu2O-based photocatalysts for CO2RR.
Table 1. The application of Cu2O-based photocatalysts for CO2RR.
PhotocatalystSynthesis MethodMorphology and StructureSizeBandgap (Eg)Light ResourceProductYield (μmol∙g−1 h−1)Energy Conversion Efficiency/SelectivityRefs.
Cu2O/Cu/CVOHydrothermal and wet chemical reduction methodsCu2O nanoclusters and Cu NPs cover the surface of elliptic CVO NPs~100 nmEgCVO: 2.34 eV;
EgCu2O: 1.87 eV
300 W Xe lamp (λ > 400 nm)CO and CH46.97 and 1.62Selectivity: 51.3% for CO[78]
3D porous Cu2OElectrodeposition and thermal oxidation.3D porous structure23–25 μm2.0 eV300 W Xe lamp (λ > 420 nm)CO, CH4, and C2H426.8, 4.04, and 0.66[79]
Spherical Cu/Cu2OSolution chemical methodSpherical structure1 μm300 W Xe lamp (λ > 420 nm)CO, CH3OH, and H287.7, 10.2, and 5.4[80]
Cu2O-PdAA reduction and in situ methodsCube~2 μm1.90 eV300 W Xe lamp (λ > 420 nm)CO0.13[81]
Uio-66-NH2/Cu2O/CuHydrothermal methodOctahedron UiO-66-NH2 and Cu attached to the surface of polyhedron Cu2O1.5 μm2.79 eV300 W Xe lampCO4.54[82]
Cu2O-111-Cu0One-pot methodOctahedral structureside length of ~1 μm1.98 eV300 W Xe lampCH478.497%[83]
Ag4/Cu2O@rGOWater bath combining with gas-bubbling-assisted membrane reductionUltrathin rGO nanosheet and Ag NPs supported on Cu2O octahedral nanocrystalsCu2O: 300 nm;
Ag: 10.7 nm;
rGO: 1.0 nm (thickness)
EgAg/Cu2O: 1.92 eV;
EgCu2O: 2.0 eV
300 W Xe lamp (λ > 380 nm)CH482.6AQE: 1.26%. Selectivity: 95.4%[84]
1D Cu2O@Cu NRsIn situ reduction methodOne-dimensional nanorod arrays<100 nm2.03 eV350 W Xe lamp (λ > 420 nm)CH4 and C2H4AQE: 2.4%[85]
RT-Cu0.75Low temperature thermochemical reduction and photo-deposition2.72 eV100 W solar simulator with an AM 1.5 filterCH477 nmol·g−1 h−1AQE: 0.012%[86]
U-Cu2O-LTH@PCN-XIn situ reductionUltrafine nanoclusters<3 nmEgPCN: 2.62 eV;
EgU-Cu2O-LTH: 2.07 eV
300 W Xe lamp (λ > 400 nm)CH3OH51.22AQE: 1.01%[87]
Fe3O4@N-C/Cu2OAA reduction and aerobic oxidationRod-shaped core–shell nanostructure5 nm (thickness of NC shell layer)5 W Xe HID lampCH3OH146.7[88]
Dodeca-Cu2O/rGOSolution-chemistryRhombic dodecahedra400–700 nm2.16 eV300 W Xe lamp (λ > 420 nm)CH3OH17.765[89]
Carbon layer@CQDs/Cu2OHydrothermal methodNearly spherical structure~2 µm diameter2.09 eV300 W Xe lampCH3OH99.6[90]
Ti3C2 QDs/Cu2O NWs/CuSelf-assembly strategyQDs incorporated onto NWs~500 nm (diameter of NWs)2.02 eVAM 1.5,
300 W Xe lamp
CH3OH78.50[91]
Cu@Cu2OThermal treatmentCore–shell nanoparticles~70 nm diameterXe lamp (420–780 nm)HCOOH67.35AQE: 0.12% at 560 nm[92]
NH2-C@Cu2OLow temperature annealingOctahedral structure1.79 eV300 W Xe lamp (λ > 420 nm)HCOOH138.65Selectivity: 92%[93]
Table 2. The reported applications of Cu2O-based materials in photocatalytic hydrogen production in recent years.
Table 2. The reported applications of Cu2O-based materials in photocatalytic hydrogen production in recent years.
PhotocatalystSynthesis MethodMorphology and StructureSizeBandgap (Eg)Light ResourceYield (μmol∙g−1 h−1)Energy Conversion Efficiency/SelectivityRefs.
Cu2O/TiO2Ball-millingIrregular shapesAnatase: 16.2 nm
Rutile: 30.5 nm
3.08 eVHigh-pressure Hg lamp (125 W)200AQE: 1.51%
Light-to-chemical energy efficiency: 0.6%
[99]
In(OH)3-In2S3-Cu2OHydrothermal, wet chemical and electrospinning processNanofiber100–200 nm of diameterEgIn(OH)3: 5.15 eV
EgIn2S3: 1.98 eV
EgCu2O: 2.17 eV
5 W blue light LED (λmax = 420 nm)1786.5[100]
Au@Cu2OSequential ion-exchange reactionCore–shell architectures54.4 ± 4.8 nmEgCu2O: 2.40 eVXenon lamp and AM 1.5G filter55.5AQE: 0.29% at 420 nm[101]
Na2Ti6O13/CuO/Cu2OSolid-state and impregnation methodBelt morphology1 μm3.61 eVUV/vis lamp (254 nm, 4400 μW/cm2)33[102]
C@Cu2O/CuOCalcinationChrysanthemum-like crystalline2.0 eV350 W Xe lamp (40 mW/cm2)26,700External quantum efficiency (EQE): 52.4%[103]
NiCo-LDH/Cu2OElectrostatic self-assembly3D flower clusterEgNiCo-LDH: 1.78 eV
EgCu2O: 1.89 eV
5 W LED (λ ≥ 420 nm)3666[104]
Cu2O/TiO2DES-assisted synthesisCu2O nanoclusters on TiO2 surfaces1.5 nm of Cu2O nanoclusters and 25.8 nm of TiO2 particlesEgTiO2: 3.12 eV
EgCu2O: 2.13 eV
300 W Xe lamp24,210[105]
Cu@TiO2-Cu2OHydrothermal and NaBH4 treatmentUrchin-like hierarchical spheresEgTiO2: 3.18 eV
EgCu2O: 2.05 eV
300 W Xenon lamp12,000.6AQE: 8.26%[106]
Cu/Cu2OMicrowave-assisted heatingHollow spherical morphology430 ± 1.2 nm in diameter2.0 eVLED light (20 W)141[107]
Cu2O/SiO2/CdIFReactive depositionCore–shell structureEgCdIF: 5.09 eV
EgCu2O: 2.22 eV
300 W xenon lamp (340–780 nm)2879.09AQE: 0.040% at 420 nm[108]
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

Su, Q.; Zuo, C.; Liu, M.; Tai, X. A Review on Cu2O-Based Composites in Photocatalysis: Synthesis, Modification, and Applications. Molecules 2023, 28, 5576. https://doi.org/10.3390/molecules28145576

AMA Style

Su Q, Zuo C, Liu M, Tai X. A Review on Cu2O-Based Composites in Photocatalysis: Synthesis, Modification, and Applications. Molecules. 2023; 28(14):5576. https://doi.org/10.3390/molecules28145576

Chicago/Turabian Style

Su, Qian, Cheng Zuo, Meifang Liu, and Xishi Tai. 2023. "A Review on Cu2O-Based Composites in Photocatalysis: Synthesis, Modification, and Applications" Molecules 28, no. 14: 5576. https://doi.org/10.3390/molecules28145576

Article Metrics

Back to TopTop