Next Article in Journal
Impact of Oxygen-Containing Groups on Pd/C in the Catalytic Hydrogenation of Acetophenone and Phenylacetylene
Previous Article in Journal
Influence of UV-A Light Modulation on Phenol Mineralization by TiO2 Photocatalytic Process Coadjuvated with H2O2
Previous Article in Special Issue
Facile Preparation of Attapulgite-Supported Ag-AgCl Composite Photocatalysts for Enhanced Degradation of Tetracycline
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight

1
School of Chemistry and Chemical Engineering, Institute for Advanced Materials, Jiangsu University, Zhenjiang 212013, China
2
School of Chemistry and Chemical Engineering, Liaoning Normal University, Dalian 116029, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(8), 546; https://doi.org/10.3390/catal14080546
Submission received: 23 July 2024 / Revised: 13 August 2024 / Accepted: 16 August 2024 / Published: 20 August 2024
(This article belongs to the Special Issue Mineral-Based Composite Catalytic Materials)

Abstract

:
In this work, the successful synthesis of a Cu@g-C3N4/CeO2-like Z-scheme heterojunction through hydrothermal and photo-deposition methods represents high CO2 reduction activity with remarkable CO selectivity, as evidenced by the impressive CO yield of 33.8 μmol/g for Cu@g-C3N4/CeO2, which is over 10 times higher than that of g-C3N4 and CeO2 individually. The characterization and control experimental results indicate that the formation of heterojunctions and the introduction of Cu sites promote charge separation and the transfer of hot electrons, as well as the photothermal effect, which are the essential reasons for the improved CO2 reduction activity. Remarkably, Cu@g-C3N4/CeO2 still exhibits about 92% performance even after multiple cycles. In situ FTIR was utilized to confirm the production of COOH* at 1472 cm−1 and to elucidate the mechanism behind the high selectivity for CO production. The study’s investigation into the wide-ranging applicability of the Cu@g-C3N4/CeO2-like Z-scheme heterojunction catalysts is noteworthy, and the exploration of potential reaction mechanisms for CO2 reduction adds valuable insights to the field of catalysis.

Graphical Abstract

1. Introduction

In the face of the escalating energy crisis and environmental degradation spurred by the rapid pace of global industrialization, the quest for sustainable and clean energy solutions has become more critical than ever [1,2,3]. Solar energy stands out as a beacon of hope, offering a pollution-free, inexhaustible, and renewable source of power that is poised to play a pivotal role in the future landscape of energy production [4]. The harnessing of this abundant energy through semiconductor photocatalytic technology reduction of carbon dioxide (CO2) into useful chemicals or fuel presents a dual-edged sword in tackling not only energy scarcity but also the pressing issue of the environmental greenhouse effect [5].
Graphitic carbon nitride (g-C3N4) has emerged as a promising candidate due to its remarkable stability, non-toxic nature, ready availability from abundant precursors, and economic feasibility [6]. However, the inherent limitations of pure g-C3N4, such as its low solar energy utilization efficiency, modest surface area, and high recombination rates of photogenerated charge carriers, have hindered its full potential as an efficient photocatalyst [7]. To overcome these challenges, researchers have been exploring the construction of g-C3N4-based photocatalysts with enhanced performance [8,9,10,11,12]. One such approach involves the design and fabrication of g-C3N4-based heterojunctions [13], which, through bandgap engineering, can significantly boost the response to visible light [14,15,16], charge separation efficiency, and overall photocatalytic activity of g-C3N4 [17]. Especially, the Z-scheme heterojunction with high redox potential exhibits excellent photocatalytic performance, which can reduce CO2 to high-value-added hydrocarbon fuels (It refers to hydrocarbons with high market value and application value, which are usually produced through specific chemical or biological processes and have high energy density and/or specific chemical properties, making them widely used in energy, chemical, transportation and other fields).
Cerium dioxide (CeO2) has garnered significant attention due to its exceptional properties, such as low cost, high stability, high oxygen storage capacity (as shown in Figure 1a), rich array of oxygen vacancies, reversible Ce3+/Ce4+ redox couples, and robust resistance to photocorrosion [18,19,20]. The well-matched band structures to g-C3N4 to form CeO2/g-C3N4 Z-scheme heterojunctions has attracted widespread attention [21,22,23]. As far as we know the CeO2/g-C3N4 Z-scheme has been extensively reported on CO2 reduction, but there are still some issues: 1. The synthesis process require multiple steps and high temperatures, which increases production costs and complicates scalability. 2. The selectivity towards a specific product such as methane, ethylene, or methanol is often not optimal. Unwanted side reactions or competing pathways can lead to a mix of products, reducing the overall efficiency and economic viability. 3. The catalytic activity of the CeO2/g-C3N4 composite under visible light or solar irradiation is not ideal. Therefore, it is necessary to further modify the CeO2/g-C3N4 Z-scheme and systematically study the CO2 reduction performance and mechanism.
Through literature research, introducing heteroatoms into the g-C3N4 matrix or CeO2 lattice seems to be a very effective method. The structure of g-C3N4 is shown in Figure 1b. The doping of transition metals such as Cu, Co, Ce, etc. [24,25,26] within g-C3N4 or CeO2 could adjust the electronic structure, band gap and augment the charge separation process of the composite catalyst, as it allows for a sustained supply of electrons and holes that participate in the reduction and oxidation reactions, respectively [27]. Among these transition metals, Cu has attracted much attention because Cu 2p orbitals can hold more electrons, effectively facilitating the charge carrier transfer and further improving the CO2 photoreduction capability of the catalyst. Moreover, the Cu can also be excited and generate the heat to supply electrons, making it easier for these hot electrons to participate in the CO2 reduction reaction [28]. Therefore, the addition of Cu not only has the potential to establish a more effective barrier against electron-hole recombination [29,30,31], but also offers a novel approach to CO2 reduction by involving thermal electrons in the reaction, thereby further improving the photocatalytic efficiency under visible light irradiation [32].
Thus, a novel Cu@CeO2/g-C3N4-like Z-scheme heterojunction photocatalyst was prepared; by adding Cu sites (as shown in Figure 1c) to increase the thermal reaction pathway, the separation efficiency of electrons and holes can be further improved to optimize CO2 conversion. By combining materials science strategies (TEM, electrochemical) with advanced spectroscopic techniques (such as the in-situ FT-IR and FL), researchers strive to gain a deeper understanding of the complex reaction dynamics at play, with the ultimate goal of advancing sustainable technologies for CO2 utilization and contributing to the global effort against climate change.

2. Results and Discussion

2.1. Photocatalyst Characterization

The microstructures of g-C3N4, CeO2, and Cu@g-C3N4/CeO2 were evaluated using SEM and TEM. The results show that the g-C3N4 exhibits a uniform two-dimensional micro-sheet morphology with a diameter ranging from approximately 1.2–1.5 μm, as depicted in Figure 2a,d. The CeO2 also possess a two-dimensional ultra-thin micro-sheet structure with a smooth surface, as illustrated in Figure 2b,e. Upon the formation of the Cu@g-C3N4/CeO2, the surface of the CeO2 micro-sheets exhibited a noticeable increase in roughness and thickness. The SEM and TEM images in Figure 2c,f demonstrate the tight binding between the CeO2 and g-C3N4, along with the successful loading copper (Cu) [33,34]. The integration of the two-dimensional g-C3N4 with the two-dimensional CeO2 and the loading of Cu nanoparticles is anticipated to enhance the transfer and separation of photogenerated charges. This synergistic effect is expected to augment the number of active sites available for photocatalysis, thereby improving the overall catalytic activity of the material. Additionally, the elemental distribution and composition within Cu@g-C3N4/CeO2 were investigated using EDX-mapping. As shown in Figure 2(g1–g5), the mapping reveals a uniform distribution of Cu, C, N, Ce, and O elements across the surface of Cu@g-C3N4/CeO2. In summary, the comprehensive characterization of the photocatalyst using SEM, TEM, and EDX mapping analysis confirms the successful fabrication of the Cu@g-C3N4/CeO2. The unique structural features and elemental composition of the composite are expected to contribute to its enhanced photocatalytic performance.

2.2. Surface Composition and Photoelectric Analysis

Figure 3a shows the XRD patterns of the as-prepared samples. For CeO2, the diffraction peaks at 2θ = 28.5°, 33°, 47.4°, 56.3°, 59.3°, 69.3°, 76.6° and 79°; these peaks are consistent with the material’s cubic fluorite crystal structure, which belongs to the Fm-3m space group and can be indexed to the (111), (200), (220), (311), (222), (400), (331) and (420) planes [35]. g-C3N4 displays a diffraction peak at 2θ = 27.3°, corresponding to the (002) plane and reflecting the interlayer spacing [14]. The Cu@g-C3N4/CeO2 exhibits diffraction peaks at 28.5°, 33.0°, 47.4°, and 56.3°, which coincide with the known peaks of CeO2 [36]. This alignment suggests the successful formation of a hybrid material through the integration of CeO2 with g-C3N4. The preservation of CeO2’s crystallographic planes in the hybrid is evidenced by the correspondence of these peaks to the (111), (200), (220), and (311) planes of CeO2. The characteristic peak of g-C3N4 at 27.30° may be either influenced or superimposed upon by the hybridization process with CeO2 [37,38]. The minimal presence of Cu peaks should be attributed to the metal’s uniform dispersion and low concentration within Cu@g-C3N4/CeO2. Figure 3b displays the survey spectrum of CeO2, g-C3N4, Cu@g-C3N4/CeO2, which proved Cu@g-C3N4/CeO2 was made up of C, N, O, Ce and Cu. And C, N, O, Ce and Cu had atomic percentages of 26.12%, 8.11%, 34.24%, 30.33%, and 1.21%, respectively. The presence of C and N elements were found in g-C3N4, Ce and O elements in CeO2, and C, N, Ce, and O elements in Cu@ g-C3N4/CeO2; Cu was not effectively detected due to its low concentration. Figure 3c shows the C1s spectra of g-C3N4 and Cu@g-C3N4/CeO2, featured peaks at 288.4 eV and 284.5 eV, corresponding to sp2 hybridized carbon (C=N–C) and hydroxylated carbon species with C–O and C–OH, or specific out-of-plane sp3 C–N species [13,39,40]. Concurrently, the O1s high-resolution XPS spectrum of g-C3N4 is also displayed in Figure 3d, where the peaks at 529.5 eV and 532.2 eV are assigned to the O=C and O–C groups, respectively. For Cu@g-C3N4/CeO2, these peaks exhibited a slight shift towards higher binding energy relative to pure g-C3N4 [41], suggesting a significant interaction between CeO2 and g-C3N4. As depicted in Figure 3e, the N1s spectra of g-C3N4 and Cu@g-C3N4/CeO2 featured peaks at 395.1 eV and 398.7 eV, corresponding to sp2 hybridized carbon (C=N–C). Figure 3f presents the high-resolution XPS spectra of the Ce 3d levels for pure CeO2 and Cu@g-C3N4/CeO2. The Ce 3d energy level spectrum of the CeO2 sample can be deconvoluted into six contributions at 882.5 eV, 888.9 eV, 898.4 eV, 901.0 eV, 907.9 eV, and 916.7 eV, labeled as ν, ν′, ν″, μ, μ′, and μ″, where ν and μ represent the spin-orbit split states of Ce 3d5/2 and Ce 3d3/2, respectively. The peaks labeled as ν, ν′, ν″, and μ are characteristic of Ce4+, while the peaks labeled as ν″ and μ″ indicate the presence of Ce3+ ions. The coexistence of Ce3+/Ce4+ on the surface of Cu@g-C3N4/CeO2 suggests that under oxygen-deficient conditions, CeO2 can release lattice oxygen, which has an active effect on surface adsorption [42]. Some Ce4+ ions are reduced to Ce3+, forming oxygen vacancies. Under oxygen-rich conditions, the adsorbed oxygen is stored in the form of lattice oxygen, allowing Ce3+ to be re-oxidized to Ce4+, which helps improve charge separation efficiency and enhances the photocatalytic activity. Additionally, the characteristic peaks of N 1s for Cu@g-C3N4/CeO2 also show a slight shift towards higher binding energy compared to pure g-C3N4, further confirming the strong interaction between CeO2 and g-C3N4.
The decrease in photoluminescence (PL) intensity and extended fluorescence (FL) lifetime of Cu@g-C3N4/CeO2 compared to pure g-C3N4 under excitation light at 375 nm, as illustrated in Figure 4a,b, indicates a reduction in recombination of charge carriers within the heterojunction structure. This phenomenon suggests that the Cu@g-C3N4/CeO2 Z-scheme catalyst effectively promotes the separation and migration of photoexcited electrons and holes, leading to enhanced photocatalytic performance, which are crucial factors in achieving high catalytic activity and selectivity in CO2 reduction processes [43]. The UV–Vis diffuse reflectance spectra (DRS) analysis depicted in Figure 4c reveals that there is no significant change in the absorption edge and strength of CeO2, g-C3N4, and Cu@g-C3N4/CeO2 [44]. This observation suggests that the introduction of Cu into the g-C3N4/CeO2 heterojunction structure does not notably alter the light-absorption properties of the materials. The consistent absorption characteristics across the different samples indicate that the enhanced photocatalytic performance of Cu@g-C3N4/CeO2 is likely attributed to factors other than changes in light-absorption capacity. Figure 4d presents a detailed depiction of the photocurrent response characteristics of three distinct materials: pristine CeO2, g-C3N4, and the novel Cu@g-C3N4/CeO2 composite. Observations indicate that both pristine CeO2 and g-C3N4 exhibit relatively stable photocurrent responses under prolonged illumination, with no significant decay in photocurrent over time. This phenomenon reveals their superior photochemical stability, conducive to maintaining high efficiency under continuous light exposure [45,46]. Moreover, the Cu@g-C3N4/CeO2 shows a higher photocurrent response compared to individual CeO2 and g-C3N4. This finding not only confirms the significant enhancement in the photochemical activity but also implies Cu@g-C3N4/CeO2 possesses superior charge carrier separation capabilities. Such improvements in performance are likely attributed to the synergistic effects between CeO2 and g-C3N4 within the composite, as well as the optimization of charge transfer and transport properties brought about by the incorporation of Cu elements [13,14].
As shown in Figure 4e, CeO2 and g-C3N4 exhibit higher impedance values, and the Cu@g-C3N4/CeO2 exhibits the lowest charge transfer impedance; the low impedance characteristic is likely due to the synergistic effect between CeO2 and g-C3N4, as well as the modification effect of Cu, which collectively promote efficient charge transfer, thereby enhancing the electrochemical activity of the material [47,48]. Additionally, the Mott–Schottky (M-S) plots in Figure 4f, at different frequencies of 2000 Hz, 2500 Hz, and 3000 Hz, of Cu@g-C3N4/CeO2 consistently show a positive slope, indicating that the material is an n-type semiconductor with electrons as the primary charge carriers. Moreover, the flat-band potential (Efb) is determined to be −0.35 V and remains stable across different frequencies, indicating the absence of defect or trap states within Cu@g-C3N4/CeO2, implying minimal loss of charge carriers during their transport.

2.3. Photocatalytic Activity

This experiment used gas chromatography to detect the performance of catalysts in CO2 reduction under a xenon lamp (1000 mW/cm2), stirring and adsorption for 30 min to achieve the adsorption and desorption equilibria. CO is our only CO2 reduction product in this system. As shown in Figure 5a, in the presence of condensed water (6 °C), the CO yields of g-C3N4, CeO2 and Cu@g-C3N4/CeO2 are 0.56, 1.68, 11.44 μmol/g; the CO yield of Cu@g-C3N4/CeO2 has been significantly improved; the reason should be the formation of the like-Z-scheme and the introducing of Cu reaction sites, allowing charge carriers to be quickly transferred and participate in CO2 reduction reactions. In order to verify the effect of hot electrons on CO2 reduction, the CO2 reduction reaction without temperature control was conducted and it was found that the CO yields of g-C3N4, CeO2 and Cu@g-C3N4/CeO2 were 1.96, 3.28 and 33.82 μmol/g, respectively [49]; the result is shown in Figure 5b. Through data comparison, it was found that the CO performance without condensate water was more than three times higher than that with condensate water; this is because the addition of Cu makes better use of heat, converting some photo-generated charge carriers into hot electrons. These hot electrons not only participate in the reduction reaction of CO2, but also bombard g-C3N4 at high speed due to the large amount of energy they carry, causing more charge carriers to participate in the chain reaction. This explains why higher CO yields can be achieved without condensation. Cycling experiment without condensate (Figure 5c) was carried out and it was found that after 5 cycles, the performance of Cu@g-C3N4/CeO2 slightly decreased. XRD (X-ray diffraction) before and after the cycles shows that the crystal of Cu@g-C3N4/CeO2 has not changed (Figure 6a). In order to determine that CO comes from CO2, the different blank control experiments results are shown in Figure 6b. (a) Under a N2 atmosphere for 4 h of illumination, no significant CO generation was detected. (b) Under a CO2 atmosphere but without light, no significant CO generation was detected after 4 h. (c) Under normal testing conditions. (d) Under a CO2 atmosphere illumination but without a catalyst, no significant CO generation was detected. Therefore, the prepared Cu@g-C3N4/CeO2 can be identified as a stable photocatalyst.

2.4. Mechanism of CO2 Photoreduction

The in situ FTIR analysis was used to detect the intermediate products, in an attempt to clarify the mechanism of the reaction process of CO2 photoreduction. As shown in Figure 7, a peak at 1398 cm−1 is b-CO32−, and the inverted peak at 1672 cm−1 was attributed to the typical bending vibrations of molecular water. These peaks indicate the water molecules participate in the reaction process as the electron donors. With the LED illumination, two distinct positive peaks appeared, which belonged to the intermediates of the CO2 photoreaction process; the peak at 1472 cm−1 belonged to the COOH* [13], which is the critical intermediate from CO2 to CO; moreover, the peaks at 1123 and 1063 cm−1 belonged to HCO3, and the peaks at 1398 cm−1 were considered to be the CO32− from the hydrolysis of CO2. Based on the in situ FTIR results, the CO2 photoreduction process can be inferred as follows:
The CO2 reduction mechanism of Cu@g-C3N4/CeO2 is shown in Figure 8. The light energy generated under xenon lamp irradiation not only stimulates the generation of e/h+ pairs but also causes intramolecular motion, resulting in an undeniable thermal effect. The thermal effect of this part further promotes the generation of electron hole pairs. At the same time, this part of electrons is generated by receiving the thermal energy of intramolecular motion, so we call them hot electrons. The addition of Cu reaction sites not only builds a bridge for high-speed electron conduction between the two, but also serves as a high-speed channel for heat conduction. This allows photo-generated electrons to be converted into hot electrons, which are always in a high-speed-motion state, making it difficult to recombine with holes. These electrons are difficult to recombine with holes, and the electron transport channel constructed by Cu@g-C3N4/CeO2 makes it easier for them to be delivered to the catalyst surface and react with CO2.
Cu@g-C3N4/CeO2 + Photothermal → Cu@g-C3N4/CeO2 (e + h+)
CO2 → CO2*
H2O + h+ → ·OH + H+
H2O + ·OH + 3h+ → O2 + 3H+
2H+ + 2e → H2
CO2 + H2O → H2CO3
H2CO3 → H+ + HCO3
HCO3 → H+ + CO32−
CO2* + e + H+ → COOH*
COOH* + e + H+ → CO* + H2O
CO* → CO

3. Experimental Section

3.1. Methods

3.1.1. Materials

The precursors used to prepare Cu@g-C3N4/CeO2-like Z-scheme heterojunctions are Cupric Nitrate Trihydrate (Cu (NO3)2·3H2O, ≥99.0%), bought from Macklin (Shanghai, China). Cerium nitrate hexahydrate (Ce (NO3)3·6H2O ≥ 99.95%) was bought from Aladdin Bio-Chem Technology Co. (Shanghai, China). Ammonium bicarbonate (NH4HCO3 ≥ 97.0%), Urea (CO(NH2)2 ≥ 97.0%), Polyvinylpyrrolidone (PVP ≥ 97.0%) and Methyl alcohol (MeOH ≥ 97.0%) were bought from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). All experiments were performed with deionized water, and all reagents did not require further purification.

3.1.2. Synthesis of CeO2 Micro-Sheets

Ce(NO3)3·6H2O (0.013 mmol) and NH4HCO3 (0.037 mmol) were dissolved into 200 mL deionized water and stirred thoroughly. Then the mix was statically aged at 30 °C for 24 h and the white precipitate was obtained. After the sediment was separated by filtration, it was washed three times with deionized water and then dried under 60 °C for 8 h. Finally, the dried sample was transferred to a tube furnace and calcined at 500 °C for 4 h to obtain the CeO2 sample.

3.1.3. Synthesis of g-C3N4 Micro-Sheets

The single-step pyrolysis method was employed to create g-C3N4 by using CO(NH2)2 as a precursor. In a muffle furnace, 10 g of urea were heated to 550 °C for 240 min to obtain g-C3N4 micro-sheets [13].

3.1.4. Preparation of g-C3N4/CeO2 Photocatalysts

CeO2 (0.2 g) and g-C3N4 (0.8 g) were dissolved in a 10 mL methanol solution, stirred thoroughly at 60 °C for 24 h, followed by drying to obtain the solid powder of g-C3N4/CeO2 [14].

3.1.5. Preparation of Cu@g-C3N4/CeO2 Photocatalysts

A total of 0.1 g of the obtained solid powder from the above, along with 0.01 g of Cu(NO3)2, was dispersed in 20 mL of deionized water and sonicated for 10 min. Subsequently, the dispersion was subjected to photo-deposition under UV light (365 nm) for 30 min. Finally, the resulting Cu@g-C3N4/CeO2 was obtained after centrifugation and drying.

3.2. Characterizations

The crystal structure was examined by X-ray diffraction (XRD model, MAC Science, Yokohama, Japan) using Ni-filtered Cu-K radiation. At 2θ, the scanning range was 10–80° and the scanning speed 10°/min. The measurements of X-ray photoelectron spectroscopy (XPS) were analyzed on a PerkinElmer PHI 5300 (PerkinElmer, Shanghai, China) instrument to check the element’s status on the sample surface tested with a monochrome Mg K source. The High0 Resolution Transmission Electron Microscope (HRTEM) images were taken on a field-emission electron microscope (Tecnai G2 F20, Hitachi, Tokyo, Japan) with an acceleration voltage. The XRD patterns were performed on a Shimazu-6100 powder X-ray (Shimadzu, Tokyo, Japan) diffractometer with Cu Kα radiation at a scan rate of 6° min−1. The XPS data were obtained using a Shimadzu/Krayos AXIS Ultra DLD (Hitachi, Tokyo, Japan). The scanning electron microscopy (SEM) images were characterized on a field-emission scanning electron microscope (FE-SEM, JSM-7001F, Hitachi, Tokyo, Japan). The ultraviolet–visible diffuse reflectance spectroscopy (UV–Vis DRS) was carried out on a Shimadzu UV-3600 spectrometer. The photoluminescence (PL) spectra for solid power were investigated by an F4500 (Hitachi, Tokyo, Japan) and the xenon (Xe) lamp with an excitation wavelength of 375 nm. In situ Fourier transform infrared spectroscopy (in situ FTIR) was carried out to research the CO2 photoreduction process (Thermo Fisher Nicolet iS-10, Thermo Fisher Scientific, Waltham, MA USA).

3.3. CO2 Photoreduction Experiments

CO2 photoreduction was carried out in a sealed self-made 50 mL quartz reactor with a 300 W Xe lamp (1000 mW/cm2, Education Au-light Co., Ltd., Beijing, China) as the white-light source. In a typical procedure, 10 mg of catalyst was dispersed in 10 mL of deionized water. Then, it was subjected to ultrasound for 10 min to make it completely disperse. CO2 was then introduced into the reactor for 20 min to completely remove air. Gas products were detected by a gas chromatography (GC 5890N, Education Au-light Co., Ltd., Beijing, China) equipped with hydrogen flame ionization detector (FID), and the carrier gas was N2.

3.4. Photo-Electrochemical Measurements

The corresponding electrode was prepared as follows: 0.05 g of the prepared photocatalyst was dispersed in 3.0 mL of ethanol and 0.03 mL of oleic acid, and then 0.01 g of polyvinylpyrrolidone (PVP) was added. The above mixture was spin-coated on a 1 × 1 cm2 fluorine-doped tin oxide glass electrode, and then dried at room temperature. The measurement system includes a standard three-electrode quartz cells with 0.5 M Na2SO4 electrolyte solution; a Pt wire and a saturated calomel electrode (SCE) were used as the counter electrode and the reference electrodes, respectively. The amplitude of applied sine wave potential in each case was 5 mV, which was carried out using the Chen Hua electrochemical workstation, and all electrochemical signals were recorded by a CHI660 B electrochemical analyzer (Chen Hua Instruments, Shanghai, China).

4. Conclusions

The Cu@g-C3N4/CeO2-like Z-scheme heterojunction was meticulously fabricated through hydrothermal and photo-deposition techniques. It represents charming CO2 reducing activity and product selectivity (CO selectivity 100%), attributed to the internal electric field between g-C3N4 and CeO2, promoting the photogenerated charge transfer, as well as the introduction of Cu, which generates high-energy hot electrons, triggering a cascade of charge carrier generation. The findings expand the photocatalytic mechanism beyond the conventional photoreduction induced by photogenerated carriers to include photothermal reactions. The enhanced charge separation, the introduction of high-energy hot electrons, and the robust stability of the material are key factors that make this heterojunction a promising candidate for CO2 photoreduction under white-light irradiation. This work not only advances the understanding of photocatalytic mechanisms but also opens new avenues for the development of efficient and sustainable photocatalytic systems.

Author Contributions

Conceptualization, Y.Y. and Z.Z.; methodology, Y.Z. (Yiying Zhou); software, B.H., J.C. and Y.Z. (Yue Zhang); validation, Y.Z.(Yiying Zhou), Y.S., S.J. and Z.L.; formal analysis, Y.Z. (Yiying Zhou), Y.S., S.J. and Z.L.; investigation, Y.Z. (Yue Zhang); resources, B.H., J.C, X.T. and Y.Z. (Yue Zhang); data curation, Y.Y. and Z.Z.; writing—original draft preparation, Y.Z. (Yiying Zhou) and Z.Z.; writing—review and editing, Y.Z. (Yiying Zhou); supervision, Z.Z.; project administration, Z.Z.; funding acquisition, Y.Y. and Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (No. 22208127), the Senior Talent Research Foundation of Jiangsu University (No.22JDG017, No.23JDG030), the RGC Postdoctoral Fellowship Scheme of Hong Kong (RGC-PDFS-2324-2S04), and the Postgraduate Research and Practice Innovation Program of Jiangsu Province (KYCX24_3952, SJCX24_2419, KYCX24_4009). This work was financially supported by the Research project approval of Jiangsu University (23A097, Y23A145).

Data Availability Statement

All data are available as download within the database.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [CrossRef]
  2. Li, K.; Peng, B.; Peng, T. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  3. Gong, E.; Ali, S.; Hiragond, C.B.; Kim, H.S.; Powar, N.S.; Kim, D.; Kim, H.; In, S.I. Solar fuels: Research and development strategies to accelerate photocatalytic CO2 conversion into hydrocarbon fuels. Energy Environ. Sci. 2022, 15, 880. [Google Scholar] [CrossRef]
  4. Wang, X.; Liu, Q.; Bai, Z.; Lei, J.; Jin, H. Thermodynamic investigations of the supercritical CO2 system with solar energy and biomass. Appl. Energy 2017, 227, 108–118. [Google Scholar] [CrossRef]
  5. Jia, J.; O’Brien, P.G.; He, L.; Qiao, Q.; Fei, T.; Reyes, L.M.; Burrow, T.E.; Dong, Y.; Liao, K.; Varela, M.; et al. Carbon Dioxide Reduction: Visible and Near-Infrared Photothermal Catalyzed Hydrogenation of Gaseous CO2 over Nanostructured Pd@Nb2O5. Adv. Sci. 2016, 3, 10. [Google Scholar] [CrossRef]
  6. Hou, S.; Gao, X.; Lv, X.; Zhao, Y.; Yin, X.; Liu, Y.; Fang, J.; Yu, X.; Ma, X.; Ma, T.; et al. Decade Milestone Advancement of Defect-Engineered g-C3N4 for Solar Catalytic Applications. Nano-Micro Lett. 2024, 16, 70. [Google Scholar] [CrossRef] [PubMed]
  7. Li, B.; Li, N.; Guo, Z.; Wang, C.; Zhu, Z.; Tang, X.; Andrew Penyikie, G.; Wang, X.; Huo, P. Understanding the effect of multiple configurations in polymeric carbon nitride for efficient solar-driven hydrogen peroxide photosynthesis. Chem. Eng. J. 2024, 491, 152022. [Google Scholar] [CrossRef]
  8. Ye, Q.; Zhou, Y.; Xu, Y.; Zhang, Q.; Shi, X.; Li, D.; Tian, D.; Jiang, D. Improved charge transfer in polymeric carbon nitride synergistically induced by the aromatic rings modification and Schottky junctions for efficient photocatalytic CO2 reduction. Chem. Eng. J. 2023, 463, 142395. [Google Scholar] [CrossRef]
  9. Song, X.; Wu, Y.; Zhang, X.; Li, X.; Zhu, Z.; Ma, C.; Yan, Y.; Huo, P.; Yang, G. Boosting charge carriers separation and migration efficiency via fabricating all organic van der Waals heterojunction for efficient photoreduction of CO2. Chem. Eng. J. 2020, 408, 127292. [Google Scholar] [CrossRef]
  10. Ma, W.; Wang, N.; Guo, Y.; Yang, L.; Lv, M.; Tang, X.; Li, S. Enhanced photoreduction CO2 activity on g-C3N4: By synergistic effect of nitrogen defective-enriched and porous structure, and mechanism insights. Chem. Eng. J. 2020, 388, 124288. [Google Scholar] [CrossRef]
  11. Jun, W.; Bi, J.; Chuan, J.; Cheng, J. Hierarchical Porous O-Doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 1603938. [Google Scholar]
  12. Jiang, T.; Wang, Z.; Wei, G.; Wu, S.; Huang, L.; Li, D.; Ruan, X.; Liu, Y.; Jiang, C.; Ren, F. Defective High-Crystallinity g-C3N4 Heterostructures by Double-End Modulation for Photocatalysis. ACS Energy Lett. 2024, 9, 1915–1922. [Google Scholar] [CrossRef]
  13. Liu, Q.; Zhao, X.; Song, X.; Liu, X.; Zhou, W.; Wang, H.; Huo, P. Pd Nanosheet-Decorated 2D/2D g-C3N4/WO3·H2O S-Scheme Photocatalyst for High Selective Photoreduction of CO2 to CO. Inorg. Chem. 2022, 61, 4171–4183. [Google Scholar] [CrossRef]
  14. Hu, H.; Hu, J.; Wang, X.; Gan, J.; Su, M.; Ye, W.; Zhang, W.; Ma, X.; Wang, H. Enhanced reduction and oxidation capability over the CeO2/g-C3N4 hybrid through surface carboxylation: Performance and mechanism. Catal. Sci. Technol. 2020, 10, 4712–4725. [Google Scholar] [CrossRef]
  15. Gao, W.; Li, Y.; Xiao, D.; Ma, D. Advances in photothermal conversion of carbon dioxide to solar fuels. J. Energy Chem. 2023, 83, 62–78. [Google Scholar] [CrossRef]
  16. Yang, X.; Peng, J.; Zhao, L.; Zhang, H.; Li, J.; Yu, P.; Fan, Y.; Wang, J.; Liu, H.; Dou, S. Insights on advanced g-C3N4 in energy storage: Applications, challenges, and future. Carbon Energy 2024, 9, 490. [Google Scholar] [CrossRef]
  17. Yin, S.; Sun, L.; Zhou, Y.; Li, X.; Li, J.; Song, X.; Huo, P.; Wang, H.; Yan, Y. Enhanced electron–hole separation in SnS2/Au/g-C3N4 embedded structure for efficient CO2 photoreduction. Chem. Eng. J. 2020, 406, 126776. [Google Scholar] [CrossRef]
  18. Ji, H.; Lin, L.; Chang, K. Plasma-assisted CO2 decomposition catalyzed by CeO2 of various morphologies. J. CO2 Util. 2022, 68, 102351. [Google Scholar] [CrossRef]
  19. Tran, D.; Pham, M.; Bui, X.; Wang, Y.; You, S. CeO2 as a photocatalytic material for CO2 conversion: A review. Sol. Energy 2022, 240, 443–466. [Google Scholar] [CrossRef]
  20. Yoko, A.; Wang, H.; Furuya, K.; Takahashi, D.; Seong, G.; Tomai, T.; Frenkel, A.I.; Saito, M.; Inoue, K.; Ikuhara, Y.; et al. Reduction of (100)-Faceted CeO2 for Effective Pt Loading. Chem. Mater. 2024, 36, 5611–5620. [Google Scholar] [CrossRef]
  21. Zhu, Z.; Chen, C.Y.; Wu, R.J. Hydrocarbon production by addition of Cu-ZnO on g-C3N4 for CO2 conversion. J. Chin. Chem. Soc. 2020, 67, 1654–1660. [Google Scholar] [CrossRef]
  22. Tang, X.; Shen, W.; Li, D.; Li, B.; Wang, Y.; Song, X.; Zhu, Z.; Huo, P. Research on cobalt-doping sites in g-C3N4 framework and photocatalytic reduction CO2 mechanism insights. J. Alloys Compd. 2023, 954, 170044. [Google Scholar] [CrossRef]
  23. Liang, J.; Zhang, L.; Li, C.; Mo, Z.; Ye, M.; Zhu, Z.; Sun, S.; Wong, J.W.C. Triclocarban transformation and removal in sludge conditioning using chalcopyrite–triggered percarbonate treatment. J. Hazard. Mater. 2024, 463, 132944. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, J.; Heil, T.; Zhu, B.; Tung, C.-W.; Yu, J.; Chen, H.M.; Antonietti, M.; Cao, S. A Single Cu-Center Containing Enzyme-Mimic Enabling Full Photosynthesis under CO2 Reduction. ACS Nano 2020, 14, 8584–8593. [Google Scholar] [CrossRef]
  25. Zhu, Z.; Ye, J.; Tang, X.; Chen, Z.; Yang, J.; Huo, P.; Ng, Y.H.; Crittenden, J. Vacancy-Rich CoSx@LDH@Co-NC Catalytic Membrane for Antibiotic Degradation with Mechanistic Insights. Environ. Sci. Technol. 2023, 57, 16131–16140. [Google Scholar] [CrossRef]
  26. Zhu, Z.; Xing, X.; Qi, Q.; Shen, W.; Wu, H.; Li, D.; Li, B.; Liang, J.; Tang, X.; Zhao, J.; et al. Fabrication of Graphene Modified CeO2/g-C3N4 Heterostructures for Photocatalytic Degradation of Organic Pollutants. Chin. J. Struct. Chem. 2023, 42, 100194. [Google Scholar] [CrossRef]
  27. Vikatakavi, A.; Mauri, S.; Rivera-Salazar, M.L.; Dobovičnik, E.; Pelatti, S.; D’Addato, S.; Torelli, P.; Luches, P.; Benedetti, S. Role of Metal Dopants in Hydrogen Dissociation on Cu:CeO2 and Fe:CeO2 Surfaces Studied by Ambient-Pressure X-ray Absorption Spectroscopy. ACS Appl. Energy Mater. 2024, 7, 2746–2754. [Google Scholar] [CrossRef]
  28. Li, S.; Lu, J.; Wang, G.; Li, X.; Liu, W.; Xin, C.; Zhan, H. Study on the performance of photothermal catalyzed carbon dioxide hydrogenation to CH4 over doped sodium tantalate-based perovskite catalysts. J. Environ. Chem. Eng. 2024, 12, 113108. [Google Scholar] [CrossRef]
  29. Pan, D.; Wang, Y.; Lin, L.; Zhang, Y.; Zhou, M.; Li, Z.; Xu, S. Cu/Ni-loaded oxygen vacancy-rich CeO2 rod derived from metal organic framework for efficient photothermal CO2 hydrogenation. J. Mater. Sci. Mater. Electron. 2024, 35, 791. [Google Scholar] [CrossRef]
  30. Wang, X.; Guo, Y.; Li, L.; Chen, L. Carbon nanotubes @Cu/Ni loaded BFO for photothermal catalytic conversion of CO2 by H2O. Mater. Chem. Phys. 2023, 309, 128417. [Google Scholar] [CrossRef]
  31. Wang, Y.; Ding, J.; Yin, Q.; Zhang, C.; Zeng, Y.; Xu, S.; Liang, Q.; Zhou, M.; Li, Z. Reinforcing the efficiency of photocatalytic CO2 conversion with H2O vapor through the integration of photothermal Cu/C@Bi/C supported on 3D g-C3N4 under full-spectrum solar irradiation. J. Environ. Chem. Eng. 2024, 12, 113008. [Google Scholar] [CrossRef]
  32. Kong, N.; Han, B.; Li, Z.; Fang, Y.; Feng, K.; Wu, Z.; Wang, S.; Xu, A.-B.; Yu, Y.; Li, C.; et al. Ruthenium Nanoparticles Supported on Mg(OH)2 Microflowers as Catalysts for Photothermal Carbon Dioxide Hydrogenation. ACS Appl. Energy Mater. 2020, 3, 3028–3033. [Google Scholar] [CrossRef]
  33. Zhang, H.; Ding, L.; Long, H.; Li, J.; Tan, W.; Ji, J.; Sun, J.; Tang, C.; Dong, L. Influence of CeO2 loading on structure and catalytic activity for NH3-SCR over TiO2-supported CeO2. J. Rare Earths 2020, 38, 883–890. [Google Scholar] [CrossRef]
  34. Chenari, H.M.; Riasvand, L.; Khalili, S. The synthesis condition and characterization of thermally treated CeO2 and ZnO CeO2 composite fibers. Ceram. Int. 2019, 45, 14223–14228. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Feng, W.; Yang, F.; Bao, X. Interface-controlled synthesis of CeO2(111) and CeO2(100) and their structural transition on Pt(111). Chin. J. Catal. 2019, 40, 204–213. [Google Scholar] [CrossRef]
  36. Smeekens, S.P.; Malireddi, R.K.; Plantinga, T.S.; Buffen, K.; Netea, M.G. Autophagy is redundant for the host defense against systemic Candida albicans infections. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 711–722. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, R.; Qiu, G.; Xiao, Y.; Tao, X.; Li, B. Optimal construction of WO3·H2O/Pd/CdS ternary Z-scheme photocatalyst with remarkably enhanced performance for oxidative coupling of benzylamines. J. Catal. 2019, 374, 378–390. [Google Scholar] [CrossRef]
  38. Zhang, N.; Li, X.; Ye, H.; Chen, S.; Ju, H.; Liu, D.; Lin, Y.; Ye, W.; Wang, C.; Xu, Q. Oxide Defect Engineering Enables to Couple Solar Energy into Oxygen Activation. J. Am. Chem. Soc. 2016, 138, 8928–8935. [Google Scholar] [CrossRef]
  39. Tong, R.; Sun, Z.; Zhong, X.; Wang, X.; Xu, J.; Yang, Y.; Xu, B.; Wang, S.; Pan, H. Enhancement of Visible-Light Photocatalytic Hydrogen Production by CeCO3OH in g-C3N4/CeO2 System. ChemCatChem 2019, 11, 1069–1075. [Google Scholar] [CrossRef]
  40. Saravanan, V.; Lakshmanan, P.; Palanisami, N.; Amalraj, J.; Pyarasani, R.D.; Ramalingan, C. 2D/3D- C3N4/CeO2 S-scheme heterojunctions with enhanced photocatalytic performance. Inorg. Chem. Commun. 2022, 291, 898. [Google Scholar] [CrossRef]
  41. Li, M.; Zhang, L.; Fan, X.; Wu, M.; Du, Y.; Wang, M.; Kong, Q.; Zhang, L.; Shi, J. Dual synergetic effects in MoS2/pyridine-modified g-C3N4 composite for highly active and stable photocatalytic hydrogen evolution under visible light. Appl. Catal. B Environ. 2016, 190, 36–43. [Google Scholar] [CrossRef]
  42. Michalow-Mauke, K.A.; Lu, Y.; Kowalski, K.; Graule, T.; Nachtegaal, M.; Kröcher, O.; Ferri, D. Flame-Made WO3/CeOx-TiO2 Catalysts for Selective Catalytic Reduction of NOx by NH3. ACS Catal. 2015, 5, 5657–5672. [Google Scholar] [CrossRef]
  43. Feng, C.; Luo, J.; Chen, C.; Zuo, S.; Ren, Y.; Wu, Z.-P.; Hu, M.; Ould-Chikh, S.; Ruiz-Martínez, J.; Han, Y.; et al. Cooperative tungsten centers in polymeric carbon nitride for efficient overall photosynthesis of hydrogen peroxide. Energy Environ. Sci. 2024, 17, 1520–1530. [Google Scholar] [CrossRef]
  44. Feng, C.; Bo, T.; Maity, P.; Zuo, S.; Zhou, W.; Huang, K.W.; Mohammed, O.F.; Zhang, H. Regulating Photocatalytic CO2 Reduction Kinetics through Modification of Surface Coordination Sphere. Adv. Funct. Mater. 2023, 34, 2309761. [Google Scholar] [CrossRef]
  45. Wu, G.; He, Z.; Wang, Q.; Wang, H.; Wang, Z.; Sun, P.; Mo, Z.; Liu, H.; Xu, H. Non-metal inducing charge rearrangement in carbon nitride to promote photocatalytic hydrogen production. J. Mater. Sci. Technol. 2024, 195, 1–8. [Google Scholar] [CrossRef]
  46. Mo, Z.; Miao, Z.; Yan, P.; Sun, P.; Wu, G.; Zhu, X.; Ding, C.; Zhu, Q.; Lei, Y.; Xu, H. Electronic and energy level structural engineering of graphitic carbon nitride nanotubes with B and S co-doping for photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2023, 645, 525–532. [Google Scholar] [CrossRef]
  47. Zhang, X.; Kim, D.; Yan, J.; Lee, L.Y.S. Photocatalytic CO2 Reduction Enabled by Interfacial S-Scheme Heterojunction between Ultrasmall Copper Phosphosulfide and g-C3N4. ACS Appl. Mater. Interfaces 2021, 13, 9762–9770. [Google Scholar] [CrossRef]
  48. Cai, S.; Chen, J.; Li, Q.; Jia, H. Enhanced Photocatalytic CO2 Reduction with Photothermal Effect by Cooperative Effect of Oxygen Vacancy and Au Cocatalyst. ACS Appl. Mater. Interfaces 2021, 13, 14221–14229. [Google Scholar] [CrossRef]
  49. Chen, Y.; Lin, X.; Li, W.; Sun, H.; Jia, S.; Zhou, Y.; Hao, Y.; Liu, Z.; Yin, S.; Guo, C.; et al. Harnessing Photo-to-Thermal Conversion in Sulfur-Vulcanized Mxene for High-Efficiency Solar-to-Carbon-Fuel Synthesis. Adv. Funct. Mater. 2024, 34, 2400121. [Google Scholar] [CrossRef]
Figure 1. Structures of monolayer CeO2 (a), g-C3N4 (b), Cu@g-C3N4/CeO2 (c).
Figure 1. Structures of monolayer CeO2 (a), g-C3N4 (b), Cu@g-C3N4/CeO2 (c).
Catalysts 14 00546 g001
Figure 2. TEM images of the g-C3N4 (a), CeO2 (b), Cu@g-C3N4/CeO2 (c), SEM images of g-C3N4 (d), CeO2 (e), Cu@g-C3N4/CeO2 (f), EDX mapping of Cu@g-C3N4/CeO2 (g1g5).
Figure 2. TEM images of the g-C3N4 (a), CeO2 (b), Cu@g-C3N4/CeO2 (c), SEM images of g-C3N4 (d), CeO2 (e), Cu@g-C3N4/CeO2 (f), EDX mapping of Cu@g-C3N4/CeO2 (g1g5).
Catalysts 14 00546 g002
Figure 3. XRD patterns (a), XPS survey spectra (b) of g-C3N4, CeO2 and Cu@g-C3N4/CeO2, high resolution of C1s of g-C3N4 and Cu@g-C3N4/CeO2(c), O1s XPS spectra of CeO2 and Cu@g-C3N4/CeO2 (d), N1s XPS spectra of g-C3N4 and Cu@g-C3N4/CeO2 (e), Ce 3d XPS spectra of CeO2 and Cu@g-C3N4/CeO2 (f).
Figure 3. XRD patterns (a), XPS survey spectra (b) of g-C3N4, CeO2 and Cu@g-C3N4/CeO2, high resolution of C1s of g-C3N4 and Cu@g-C3N4/CeO2(c), O1s XPS spectra of CeO2 and Cu@g-C3N4/CeO2 (d), N1s XPS spectra of g-C3N4 and Cu@g-C3N4/CeO2 (e), Ce 3d XPS spectra of CeO2 and Cu@g-C3N4/CeO2 (f).
Catalysts 14 00546 g003
Figure 4. PL spectra (a), FL spectra (b), UV–Vis DRS (c), transient photocurrent response (d) and EIS Nyquist plots (e) of g-C3N4 and CeO2 and Cu@g-C3N4/CeO2, Mott–Schottky (f) of Cu@g-C3N4/CeO2.
Figure 4. PL spectra (a), FL spectra (b), UV–Vis DRS (c), transient photocurrent response (d) and EIS Nyquist plots (e) of g-C3N4 and CeO2 and Cu@g-C3N4/CeO2, Mott–Schottky (f) of Cu@g-C3N4/CeO2.
Catalysts 14 00546 g004
Figure 5. CO yields with condensed water ((a), Xenon lamp, 1000 mW/cm2, 6 °C); the CO yields over g-C3N4 and CeO2 and Cu@g-C3N4/CeO2 hybrids without condensed water ((b), Xenon lamp, 1000 mW/cm2); the recycling test for CO2 reduction using Cu@g-C3N4/CeO2 without condensed water ((c), Xenon lamp, 1000 mW/cm2).
Figure 5. CO yields with condensed water ((a), Xenon lamp, 1000 mW/cm2, 6 °C); the CO yields over g-C3N4 and CeO2 and Cu@g-C3N4/CeO2 hybrids without condensed water ((b), Xenon lamp, 1000 mW/cm2); the recycling test for CO2 reduction using Cu@g-C3N4/CeO2 without condensed water ((c), Xenon lamp, 1000 mW/cm2).
Catalysts 14 00546 g005
Figure 6. (a) XRD comparison of photocatalysis before and after photocatalysis; (b) Evolution of CO after 4 h of reaction under various reaction conditions. In the case of N2, no light, normal and without photocatalyst, respectively.
Figure 6. (a) XRD comparison of photocatalysis before and after photocatalysis; (b) Evolution of CO after 4 h of reaction under various reaction conditions. In the case of N2, no light, normal and without photocatalyst, respectively.
Catalysts 14 00546 g006
Figure 7. In situ FTIR spectra of CO2 adsorption (a) and reaction (b) of Cu@g-C3N4/CeO2 collected at different time intervals.
Figure 7. In situ FTIR spectra of CO2 adsorption (a) and reaction (b) of Cu@g-C3N4/CeO2 collected at different time intervals.
Catalysts 14 00546 g007
Figure 8. Schematic diagram of the mechanism of CO2 reduction by Cu@g-C3N4/CeO2 under visible light irradiation.
Figure 8. Schematic diagram of the mechanism of CO2 reduction by Cu@g-C3N4/CeO2 under visible light irradiation.
Catalysts 14 00546 g008
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

Zhou, Y.; Cai, J.; Sun, Y.; Jia, S.; Liu, Z.; Tang, X.; Hu, B.; Zhang, Y.; Yan, Y.; Zhu, Z. Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight. Catalysts 2024, 14, 546. https://doi.org/10.3390/catal14080546

AMA Style

Zhou Y, Cai J, Sun Y, Jia S, Liu Z, Tang X, Hu B, Zhang Y, Yan Y, Zhu Z. Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight. Catalysts. 2024; 14(8):546. https://doi.org/10.3390/catal14080546

Chicago/Turabian Style

Zhou, Yiying, Junxi Cai, Yuming Sun, Shuhan Jia, Zhonghuan Liu, Xu Tang, Bo Hu, Yue Zhang, Yan Yan, and Zhi Zhu. 2024. "Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight" Catalysts 14, no. 8: 546. https://doi.org/10.3390/catal14080546

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop