Next Article in Journal
Reduction of p-Nitrophenol with Modified Coal Fly Ash Supported by Palladium Catalysts
Previous Article in Journal
Variations in Power Conversion Efficiency on n-Type Dye-Sensitized Solar Cells with Synthesized TiO2 Nanoparticle: A Thickness Effect of Active Layer
Previous Article in Special Issue
Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic 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.
Submission received: 31 July 2024 / Revised: 30 August 2024 / Accepted: 4 September 2024 / Published: 6 September 2024
(This article belongs to the Special Issue Mineral-Based Composite Catalytic Materials)

Abstract

:
It is important to improve the separation ability of photogenerated electrons and the adsorption capacity of carbon dioxide (CO2) for efficient photoreduction of CO2. Here, we synthesized ZnCdS quantum dots (ZCS-QDs) and cerium dioxide nanosheets (CeO2) using the solvothermal method and calcination method. We combined CeO2 and ZCS-QDs to effectively enhance the charge separation efficiency, and the lifetime of photogenerated electrons was increased 4.5 times. The CO evolution rate of the optimized composite (ZCS-QDs/CeO2) was up to 495.8 μmol g−1 h−1, and it had 100% product selectivity. In addition, the stability remained high after five cycles. The CO2 adsorption capacity of the catalyst surface was observed by in situ FTIR. The test results showed that improving CO2 capture ability and promoting photogenic electron separation had positive effects on enhancing photoreduction of CO2. This study provides a reference for constructing a zero-dimensional–two-dimensional (0D–2D) heterojunction and explores potential CO2 reduction reaction mechanisms.

Graphical Abstract

1. Introduction

In recent years, the challenge of mitigating CO2 emissions and converting this greenhouse gas into useful and renewable energy sources has been a focal point for contemporary research [1]. Photocatalysis is a promising strategy, particularly due to its potential for sustainability and relatively lower costs compared to other CO2 reduction methods [2,3], and is widely used because of its low cost and sustainable advantages [4,5,6]. The steps of the photocatalytic CO2 reduction reaction are complicated and affected by many factors, including light absorption, charge carrier generation, separation, and transfer, as well as surface reactions involving the adsorption and reduction of CO2 [7]. In the field of photocatalysis, in order to improve the reaction rate and regulate the reaction pathway, researchers have made a lot of efforts [8], including reducing the particle size [9,10], loading metal onto the surface [11], dispersing to various carriers [12], etc. These methods improve the catalytic activity of the photocatalyst but only to a limited extent. Therefore, selecting a suitable photocatalyst to enhance the light absorption capacity and improve the photoelectron transport efficiency may be an effective means to improve the photocatalytic efficiency. Numerous studies have found that the selection of catalysts and system design plays a pivotal role in enhancing the overall efficiency of this process [13,14,15].
Two-dimensional (2D) cerium dioxide (CeO2) is a highly versatile material that has garnered significant interest in the realm of photocatalysis due to its excellent chemical and structural stability. In addition, the reversible switch between the Ce(III) and Ce(IV) states confers unique redox properties [16,17,18]. However, its wide bandgap typically restricts its light absorption in the ultraviolet region, limiting its efficiency under visible light; this limitation hampers the direct conversion of CO2 into fuels. Combining CeO2 with other semiconductors to form CeO2 heterojunctions (such as narrower-bandgap catalysts capable of absorbing visible light) seems to present an opportunity for strategic enhancement of light harvesting and charge separation efficiency. This approach leverages synergistic effects between the components, enhances light absorption, and promotes charge carrier dynamics.
The selection and engineering of catalysts that can not only effectively adsorb CO2 but also facilitate efficient charge transfer are central to improving the photocatalytic conversion of CO2 into renewable energy sources. Zero-dimensional (0D) quantum dot (QD) materials have been extensively investigated as the building blocks of photocatalysts due to their special quantum size effect, electron transfer properties, and their efficient absorption to visible light [19,20,21]. Ternary ZnCdS creates good conditions for adjusting the optical properties from the ultraviolet to the visible light range [22], and can adjust the appropriate bandgap range [23]. In fact, the use of QDs in combination with semiconductors in the field of photocatalysis has been reported in the past few years. For example, Yang et al. used Ag QDs anchored with CeO2, and the Ag QD/CeO2 composite improved the utilization rate of visible light, and the photocatalytic elimination of acetaldehyde and water decomposition performance improved 7.08 and 6.83 times, respectively [24]. Jiang et al. reported a modification of CeO2 using g-C3N4 QDs, which improved the photoelectric response and showed significant performance improvement in the test of photoreduction of CO2 [25]. Sre et al. deposited ZnCdS quantum dots on g-C3N4 for photoinactivation of E. coli cells. The prepared composite photocatalyst enhanced the structural stability, as well as increased a large number of reaction sites, and improved the photocatalyst efficiency [26]. So, it can be inferred that if ZnCdS-QDs are introduced on the CeO2 surface and they construct a 0D–2D heterojunction, the photocatalytic CO2 activity and the recycling ability of the photocatalysts will be enhanced.
Inspired by the consideration above, ZnCdS quantum dots were combined with a CeO2 heterojunction (ZCS-QDs/CeO2) and designed to optimize the energy band alignment and CO2 reduction activity. The ZCS-QDs/CeO2 exhibits a substantial increase in the lifetime of photogenerated electrons and achieves a remarkable CO evolution rate of 495.8 μmol g−1 h−1 without any sacrificial agents during the CO2 reduction reaction, even showing good stability after five cycles. The in situ DRIFTS result provides direct evidence of the CO2 adsorption dynamics on the CeO2 surface and confirms the vital role of CeO2 in enhancing substrate adsorption. Also, the observation of intermediates and monitoring changes in water and CO2 consumption offer deeper insights into the photocatalytic process, contributing to a better understanding of the mechanism behind the improved efficiency.

2. Results and Discussion

2.1. Photocatalyst Characterization

Scheme 1 shows the synthesis strategy of ZCS-QDs/CeO2. Figure 1a shows the XRD patterns. For the CeO2, the four major diffraction peaks are located at 28.5°, 33.1°, 47.5°, and 56.3°, corresponding to the (111), (200), (220), and (311) lattice planes of CeO2 (PDF#43-1002), respectively. For the ZCS-QDs, the six major diffraction peaks are located at 26.7°, 28.3°, 45.1°, 47.3°, 53.4°, and 56.2°, corresponding to the (002), (101), (102), (110), (103), and (200) lattice planes of ZnCdS (PDF#40-0834), respectively. The relatively small amount of ZCS-QDs and the partial coincidence of peak positions result in the ZCS-QDs’ diffraction peak being covered by the CeO2 diffraction peak; thus, no diffraction peak was observed for the ZCS-QDs in the ZCS-QDs/CeO2 photocatalyst. The TEM image of CeO2 is shown in Figure 1b, exhibiting a smooth two-dimensional sheet structure. The ZnCdS is uniformly dispersed and has a diameter of about 8 nm belonging to the morphology of the quantum dots (Figure 1c,d). From Figure 1e, it can be seen that quantum dots are fully grown on the 2D nanosheets, indicating that our catalyst of 0D–2D ZCS-QDs/CeO2 is well assembled. Moreover, the 12% ZCS-QDs/CeO2 was characterized by HRTEM. As shown in Figure 1f, the measured lattice spacing of 0.33 nm and 0.31 nm corresponds to (002) of ZCS-QDs and (111) of CeO2, respectively. From HRTEM mapping (Figure 1g), the elements of S, Cd, Zn, O, and Ce were detected, providing further proof that the ZCS-QDs/CeO2 has been successfully synthesized.

2.2. Surface Composition and Photoelectric Analysis

The chemical state of the catalyst was investigated using X-ray photoelectron spectroscopy (XPS). From the XPS survey spectrum of Figure 2a, it can be seen that the Ce, O, S, Cd, and Zn elements are found in the 12% ZCS-QDs/CeO2 catalyst. Figure 2b shows the Ce 3d levels for CeO2 and 12% ZCS-QDs/CeO2. For CeO2, the peaks at 913.9 and 895.4 eV are from the Ce 3d9 4f0 final state, and the peaks at 904.7, 898.2, 886.1, and 879.5 eV are from the Ce 3d9 4f2 and Ce 3d9 4f1 final states [27]. However, for 12% ZCS-QDs/CeO2, the peak position of 895.4 eV shifted to 896.1 eV. This was attributed to the interaction of ZCS QDs and CeO2, and resulted in the electron transfer in ZCS-QDs/CeO2. For the O 1s of CeO2 (Figure 2c), two characteristic peaks can be observed at 529.1 and 526.8 eV, representing lattice oxygen and adsorbed oxygen, respectively [28], in the spectrum of 12% ZCS-QDs/CeO2. When CeO2 is combined with the nanostructure, the surface oxygen site will greatly affect the reduction reaction of the semiconductor at the interface [29]. In addition, oxygen sites are often involved in constructing the support interface for reactivity, as well as the reaction site of various adsorbents on the carrier [30,31]. Therefore, the new characteristic peak at 530.7 eV is attributed to S-O [32]. For the S 2p spectrum of 12% ZCS-QDs/CeO2 (Figure 2d), 159.5 and 158.6 eV are attributed to S 2p1/2 and S 2p3/2 [33], respectively, while the characteristic peak at 162.4 eV was attributed to the newly formed S-O. The formation of S-O fully confirms that a close chemical interaction is formed between ZCS-QDs and CeO2, which provides good conditions for electron transfer during the reaction. Two characteristic peaks at 409.2 and 402.4 eV can be observed in the Cd 3d spectrum (Figure 2e), corresponding to Cd 3d3/2 and Cd 3d5/2, respectively [33]. Two characteristic peaks can be observed at 1042.4 and 1019.4 eV in the Zn 2p spectrum (Figure 2f), corresponding to Zn 2p1/2 and Zn 2p3/2 [34], respectively.
The light absorption capacity of the photocatalyst was measured by UV–Vis DRS [35] (Figure 3a). The CeO2 showed strong absorption properties in the ultraviolet region and an absorption range of ZCS-QDs located at 480 nm. However, 12% ZCS-QDs/CeO2 has a unique enhanced adsorption peak around 450–620 nm. Also, the bandgap value was calculated and showed in Figure 3b using the following formula:
(αhv)1/n = A(hvEg)
where α, hv, Eg, and A represent the absorption coefficient, photon energy, bandgap energy, and proportionality constant, respectively. ZCS-QDs/CeO2 has the narrowest bandgap of 2.61 eV, and the bandgaps of ZCS-QDs and CeO2 were 2.67 eV and 2.98 eV. In addition, the valence band (VB) positions of CeO2 and ZCS-QDs were determined by VB-XPS (Figure 3c,d); the corresponding VBs are 1.45 eV and 2.38 eV. Based on the above results, the band structures of CeO2 and ZCS-QDs were obtained, as shown in Figure 3e.
Figure 4a shows the results of TR-PL spectroscopy. The lifetimes of CeO2 and ZCS-QDs are 2.53 ns and 6.67 ns, and 12% ZCS-QDs/CeO2 has the longest average decay life with 30.17 ns, meaning that 12% ZCS-QDs/CeO2 has excellent charge separation ability. As shown in Figure 4b, transient photocurrent responses (TPRs) were recorded, and all the catalysts exhibited stable photocurrent densities. It is worth noting that 12% ZCS-QDs/CeO2 has the highest photocurrent density compared to CeO2 and ZCS-QDs. Also, 12% ZCS-QDs/CeO2 has the smallest EIS radius (Figure 4c). This means that 12% ZCS-QDs/CeO2 has the smallest charge transfer resistance. The above confirmed that 12% ZCS-QDs/CeO2 has the best charge transfer ability and shows the best photocatalytic ability.

2.3. Photocatalytic Activity

As shown in Figure 5a, the CO2 photoreduction activity of CeO2, ZCS-QDs, and x-ZCS-QDs/CeO2 (x means the different amount of ZCS-QDs) was determined. The CO yield of ZCS-QDs is 151.8 μmol g−1 h−1. For the x ZCS-QDs/CeO2, the CO2 reduction performance significantly increased with all values of x, and 12% ZCS-QDs/CeO2 achieved the highest performance (495.8 μmol g−1 h−1), increasing 3.3 times. However, the CO2 reduction activity of 15% ZCS-QDs/CeO2 decreased because the excessive number of ZCS-QDs cover the CO2 adsorption sites on the CeO2 surface and inhibit CO2 adsorption. In order to exclude the effect of decomposition of organic matter from the catalyst, a controlled experiment under N2 was conducted. No CO was detected (Figure 5b), which confirmed that all of the CO product comes from CO2 conversion. To investigate the stability of 12% ZCS-QDs/CeO2, cyclic experiments were conducted. It can be seen from Figure 5c that, after five cycles, the 12% ZCS-QDs/CeO2 still had high catalytic activity, meaning that the catalyst has good stability. This study was compared with values previously reported for CeO2-based materials, such as CeO2/TiO2 (CO yield of 61.9 μmol g−1 h−1) [36], CuO-CeO2/TiO2 NT (CO yield of 9.2 μmol g−1 h−1) [37], CeO2/g-C3N4 (CO yield of 31 μmol g−1 h−1) [38], and CoAl-LDH/CeO2/RGO (CO yield of 5.5 μmol g−1 h−1) [39]. Among many CeO2-based photocatalysts, ZCS-QDs have the highest CO evolution rate. It is fully confirmed that ZCS-QDs/CeO2 has great advantages in the field of efficient photocatalytic reduction of CO2.

2.4. Mechanism of CO2 Photoreduction

It is known that the CO2 adsorption capacity of catalysts is crucial for the photocatalytic reduction reaction. Thus, CO2 adsorption tests were performed by in situ DRIFTS. As shown in Figure 6a,b, the adsorption capacity of CeO2 for CO2 is significantly higher than that of ZCS-QDs (the peak at 2340 cm−1 is a characteristic peak of CO2), indicating that CeO2 has strong CO2 adsorption capacity. However, due to the wide bandgap of CeO2 and the fact that it has fewer active sites, it is difficult to effectively convert the surface-adsorbed CO2 into CO, CH4, and other high value-added products. Although, the CO2 adsorption strength of 12% ZCS-QDs/CeO2 decreased slightly compared with that of CeO2, it was still higher than that of ZCS-QDs (Figure 6c). The weakened CO2 adsorption could be caused by the ZCS-QDs covering some of the adsorption sites of CeO2.
In order to investigate the photoreduction CO2 mechanism, in situ DRIFTS was carried out. Using the pre-reaction equilibrium state as a blank background, positive or negative infrared signals during the reaction were used to represent the increase or disappearance of substances on the interface. Figure 7a,b represent the CO2 photoreduction processes of CeO2 and ZCS-QDs. The peaks at 2339 cm−1 and 1658 cm−1 correspond to CO2 and H2O consumption signals, respectively [40]. By comparing the signal strength, the A2339/A1658 of CeO2 is 0.27, and the A2339/A1658 of ZCS-QDs is 0.31. In 12% ZCS-QDs/CeO2 (Figure 7c), A2344/A1668 is 2.08; this result suggests that CeO2 and ZCS-QDs mainly consume H2O during the photoreaction, while 12% ZCS-QDs/CeO2 consumes CO2 and converts it into CO. The above results show that after CeO2 captures CO2, photogenerated electrons migrate to the catalyst surface and react with CO2, greatly inhibiting competitive reactions of H2O splitting. Moreover, in Figure 7c, during 12% ZCS-QDs/CeO2 photoreaction, a series of peaks can be observed. The positive signal peak appearing at 2355 cm−1 is the protonated CO2 intermediate (O=C=O-H+) [41]. The peaks at 1468 cm−1 correspond to COOH* species [42]. In addition, the peak at 1402 cm−1 corresponds to carbonate (m-CO32−), the peak at 1267 cm−1 belongs to bidentate carbonate (b-CO32−,) and the peak at 1145 cm−1 represents bicarbonate (HCO3) [43,44,45].
Based on the discussion of the above results, the 0D–2D ZCS-QDs/CeO2 heterojunction is used for selective photoreduction of CO2 to CO. Under visible light (300 W xenon lamp) irradiation, both CeO2 and ZCS-QDs can absorb ultraviolet light and part of the visible light spectrum to produce electrons and holes [46,47,48]. As CeO2 and ZCS-QDs have a CB position of −0.7 eV and −1.22 eV, they both have the potential for CO2 reduction of CO. When the two materials were combined together, a Z-heterojunction was formed. The electrons on the CB of CeO2 recombine with the holes from the VB of ZCS-QDs; this not only improves the efficiency of photogenerated carrier separation but also the high-reducing-capacity photogenerated electrons on ZCS-QDs react with CO2 and further improve the efficiency of photoreduction of CO2 to CO (Figure 8). The CO2 photoreduction process over ZCS-QDs/CeO2 can be inferred as follows:
ZCS-QDs/CeO2 + hv → ZCS-QDs/CeO2 (e + h+)
CO2 → CO2*
CO2* + e + H+ → COOH*
COOH* + e + H+ → CO* (g) + H2O
CO* → CO
H2O + h+ → OH* + H+
OH* + h+ → 1/2 O2 (g) + H+

3. Experimental Section

3.1. Methods

3.1.1. Materials

Sulfur powder (99.999%), 1-octadecence (ODE, 90%), oleic acid (OA, 90%), cadmium oxide (CdO, 99.99%), zinc acetate (Zn(OAc)2, 99.99%), 11-mercaptoundecanoic acid (MUA, 95%), n-hexane (97%) tetramethylammonium hydroxide pentahydrate (97%), cerium nitrate hexahydrate (Ce(NO3)3·6H2O, 99.5%), ammonium bicarbonate (NH4HCO3, 99.995%), acetone (99.5%), methyl alcohol (99.5%), and ethyl alcohol (99.7%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

3.1.2. Synthesis of ZCS-QDs

The ZCS-QD synthesis method was improved according to the reported procedure [49]. For the S-precursor, 0.51 g of sulfur powder was dissolved in 16 mL of ODE and heated to 150 °C under N2. For the ZnCd-precursor, 0.256 g of CdO and 0.368 g of Zn(OAc)2 were dissolved in a mixture of 13 mL of ODE and 6 mL of OA and then heated to 300 °C under N2.
A total of 4 mL of S-precursor was injected into the ZnCd-precursor and held at 300 °C for 10 min. When brought to room temperature, the quantum dots were precipitated with acetone and then dispersed in n-hexane for preservation.

3.1.3. Ligand Exchange of ZCS-QDs

For ligand exchange, we used methods that have been previously reported [50]. A total of 20 mg of MUA was dissolved into 15 mL of methanol and the pH of the solution was adjusted to >10 using tetramethylammonium hydroxide pentahydrate. Then, 20 mg worth of ZCS-QDs was added to the above mixed solution, and stirred overnight at 60 °C in a N2 atmosphere. The QDs were precipitated with ethyl acetate, then cleaned with acetone and dispersed in water for preservation.

3.1.4. Synthesis of CeO2

For solution A, 2.78 g of Ce(NO3)3·6H2O was added to 100 mL of water and stirred for 30 min. For solution B, 1.50 g of NH4HCO3 was added to 100 mL of water and stirred for 30 min. The two solutions, A and B, were mixed and stirred for 1h, left stationary at 30 °C for 24 h, and then centrifugally dried. Last calcination was performed at 500 °C for 4 h in an air atmosphere.

3.1.5. Synthesis of x ZCS-QDs/CeO2

A certain amount of CeO2 was added to ZCS-QD solutions of different volumes, stirred away from light for 24 h, and centrifugally dried to obtain x ZCS-QDs/CeO2.

3.2. Characterizations

Transmission electron microscopy (TEM) analyses were performed on an H-7800 microscope (Hitachi, Tokyo, Japan) with an acceleration voltage of 120 kV. High-resolution transmission electron microscopy (HRTEM) was carried out on a JEM-2010 transmission electron microscope (Tokyo, Japan). The XRD patterns were examined on a Shimazu-6100 powder X-ray diffractometer (Tokyo, Japan) using Cu Kα radiation at a scan rate of 10° min−1. Transient fluorescence spectra were produced on an Edinburgh FLS1000 (Edinburgh, UK). The surface chemical states of the samples were studied by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA). The optical properties of the photocatalyst were studied using UV–Vis diffuse reflectance spectroscopy (DRS, UV-2450 Shimadzu, Kyoto, Japan). Time-resolved photoluminescence (TRPL, Edinburgh FLS1000, UK) attenuation curves were obtained using a photoluminescence detector. In situ diffuse reflection infrared Fourier transform spectroscopy (DRIFTS) experiments were conducted on a Nicolet iS10 (Thermo) machine. In a typical procedure, the catalyst sample was sealed in a reaction chamber with a quartz window. CO2 and H2O were carried into the reaction chamber by N2 flow until equilibrium. Taking the equilibrium system before reaction as the blank background, IR signals were collected in situ during the incident irradiation of a 365 nm LED lamb (3W, Education Au-light Co., Ltd., Beijing, China) through the quartz glass window. In the in situ CO2 adsorption experiment, CO2 with progressively increasing concentration was introduced into the chamber and infrared signals were collected until the adsorption equilibrium was reached.

3.3. CO2 Photoreduction Experiments

CO2 photoreduction was carried out in a sealed 150 mL quartz reactor with a 300 W Xenon lamp (1000 mW cm−2, Education Au-light Co., Ltd., Beijing, China) as the white light source. In a typical procedure, 10 mg of catalyst was dispersed in 3 mL of deionized water. The resulting mixture was dropped on quartz glass and then dried at 60 °C. The dried catalyst was placed in the quartz reactor with 1 mL of H2O. CO2 was then introduced into the reactor for 30 min to completely remove the air. During the reaction, a Xenon lamp illuminated the catalyst sample through the quartz window (22.1 cm2). Gas products were detected by a gas chromatograph (GC-7920, Beijing, China) equipped with a hydrogen flame ionization detector (FID).

3.4. Photoelectrochemical Measurements

The photoelectric chemistry experiment was tested on the electrochemistry workstation (CHI 660E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China). The reaction was carried out in a homemade standard three-electrode cell. The reference electrode and the reverse electrode were saturated Ag/AgCl and Pt electrodes, respectively, and 0.5 M Na2SO4 solution was used as the electrolyte. The 0.02 g photocatalyst, 0.03 mL oleic acid, 0.01 g PVP were dispersed into 3 mL ethanol and ultrasonic for 1 h. The prepared suspension is then uniformly dropped onto the FTO glass. Using 300 W xenon lamp as light source, the effective area is about 1 cm2.

4. Conclusions

In summary, the 0D–2D ZCS-QDs/CeO2 heterojunction was successfully prepared using solvothermal and calcination methods, and it was used in a high-efficiency photocatalytic CO2 reduction reaction. When using no sacrificial agent, the CO evolution rate reached 495.8 μmol g−1 h−1, which was 3.3 times higher than that of the ZCS-QD monomer, and has 100% selectivity. After five cycles, the catalytic activity was still high. This efficient photoconversion of CO2 can be attributed to the combination of two monomers, ZCS-QDs and CeO2, which promote the effective separation of photogenerated electrons, and the lifetime of photogenerated electrons is increased by 4.5 times. In addition, in situ FTIR was used to observe the changes in H2O and CO2 consumption during the reaction and monitor the formation of intermediate products. The result shows that the adsorption of CO2 on the surface of CeO2 solves the problem of low porosity on the surface of ZCS-QDs, creating good conditions for efficient photoreduction of CO2. The formation of intermediates such as COOH*, m-CO32−, b-CO32−, HCO3, and OH* was monitored. Therefore, a well-designed ZCS-QDs/CeO2 heterojunction can be used as an effective composite photocatalyst. In the future design of photocatalysts, synergistically improving both the CO2 capture ability and the photoelectron separation efficiency is an effective means to improve the efficiency of the photocatalyst. This provides a new strategy for efficient CO2 conversion.

Author Contributions

Investigation, data curation, and writing of the original draft, J.Y. and X.T.; resources, Y.S. and J.C.; writing—review, M.C., B.H. and Y.Z.; conceptualization, supervision, and writing—review and editing, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 22208127), the Senior Talent Research Foundation of Jiangsu University (No.23JDG030, 22GDG017), the RGC Postdoctoral Fellowship Scheme of Hong Kong (RGC-PDFS-2324-2S04), and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX24_3952, SJCX24_2419). This work was financially supported by the research project approval of Jiangsu University (23A046, Y23A145).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the valorization of exhaust carbon: From CO2 to chemicals, materials, and fuels. technological use of CO2. Chem. Rev. 2014, 114, 1709–1742. [Google Scholar]
  2. 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. 2021, 406, 126776. [Google Scholar] [CrossRef]
  3. 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]
  4. 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]
  5. 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]
  6. 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] [PubMed]
  7. Li, H.; Cheng, C.; Yang, Z.; Wei, J. Encapsulated CdSe/CdS nanorods in double-shelled porous nanocomposites for efficient photocatalytic CO2 reduction. Nat. Commun. 2022, 13, 6466. [Google Scholar] [CrossRef]
  8. Chang, X.; Wang, T.; Gong, J. CO2 Photo-reduction: Insights into CO2 Activation and Reaction on Surfaces of Photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar]
  9. Pathak, P.; Meziani, M.J.; Li, Y.; Cureton, L.T.; Sun, Y.-P. Improving photoreduction of CO2 with homogeneously dispersed nanoscale TiO2 catalysts. Chem. Commun. 2004, 10, 1234–1235. [Google Scholar] [CrossRef]
  10. Kočí, K.; Obalová, L.; Matějová, L.; Plachá, D.; Lacný, Z.; Jirkovský, J.; Šolcová, O. Effect of TiO2 particle size on the photocatalytic reduction of CO2. Appl. Catal. B Environ. 2009, 89, 494–502. [Google Scholar] [CrossRef]
  11. Wang, Y.; Li, B.; Zhang, C.; Cui, L.; Kang, S.; Li, X.; Zhou, L. Ordered mesoporous CeO2-TiO2 composites: Highly efficient photocatalysts for the reduction of CO2 with H2O under simulated solar irradiation. Appl. Catal. B Environ. 2013, 130–131, 277–284. [Google Scholar] [CrossRef]
  12. Li, Y.; Wang, W.-N.; Zhan, Z.; Woo, M.-H.; Wu, C.-Y.; Biswas, P. Photocatalytic reduction of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts. Appl. Catal. B Environ. 2010, 100, 386–392. [Google Scholar] [CrossRef]
  13. Zhu, Z.; Ye, J.; Tang, X.; Chen, Z.; Yang, J.; Huo, P.; Ng, Y.; 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]
  14. Zhao, X.; Xu, M.; Song, X.; Zhou, W.; Liu, X.; Yan, Y.; Huo, P. Charge separation and transfer activated by covalent bond in UiO-66-NH2/RGO heterostructure for CO2 photoreduction. Chem. Eng. J. 2022, 437, 135210. [Google Scholar] [CrossRef]
  15. 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]
  16. Hammond, O.S.; Edler, K.J.; Bowron, D.T.; Torrente-Murciano, L. Deep eutectic-solvothermal synthesis of nanostructured ceria. Nat. Commun. 2017, 8, 14150. [Google Scholar] [CrossRef]
  17. Wang, Z.; Yu, R. Hollow Micro/Nanostructured Ceria-Based Materials: Synthetic Strategies and Versatile Applications. Adv. Mater. 2018, 31, e1800592. [Google Scholar] [CrossRef]
  18. Huang, X.; Zhang, K.; Peng, B.; Wang, G.; Muhler, M.; Wang, F. Ceria-Based Materials for Thermocatalytic and Photocatalytic Organic Synthesis. ACS Catal. 2021, 11, 9618–9678. [Google Scholar] [CrossRef]
  19. Selopal, G.S.; Zhao, H.; Wang, Z.M.; Rosei, F. Core/Shell Quantum Dots Solar Cells. Adv. Funct. Mater. 2020, 30, 1908762. [Google Scholar] [CrossRef]
  20. Ning, Z.; Tian, H.; Yuan, C.; Fu, Y.; Qin, H.; Sun, L.; Ågren, H. Solar cells sensitized with type-II ZnSe–CdS core/shell colloidal quantum dots. Chem. Commun. 2011, 47, 1536–1538. [Google Scholar] [CrossRef] [PubMed]
  21. Kouhnavard, M.; Ikeda, S.; Ludin, N.; Khairudin, N.A.; Ghaffari, B.; Mat-Teridi, M.; Ibrahim, M.; Sepeai, S.; Sopian, K. A review of semiconductor materials as sensitizers for quantum dot-sensitized solar cells. Renew. Sustain. Energy Rev. 2014, 37, 397–407. [Google Scholar] [CrossRef]
  22. Yakoubi, A.; Ben Chaabane, T.; Aboulaich, A.; Mahiou, R.; Balan, L.; Medjahdi, G.; Schneider, R. Aqueous synthesis of Cu-doped CdZnS quantum dots with controlled and efficient photoluminescence. J. Lumin 2016, 175, 193–202. [Google Scholar] [CrossRef]
  23. Saha, S.; Sain, S.; Meikap, A.; Pradhan, S. Microstructure characterization and electrical transport of nanocrystalline CdZnS quantum dots. Phys. E Low-Dimens. Syst. Nanostruct. 2015, 66, 59–66. [Google Scholar] [CrossRef]
  24. Yang, H.; Xu, B.; Zhang, Q.; Yuan, S.; Zhang, Z.; Liu, Y.; Nan, Z.; Zhang, M.; Ohno, T. Boosting visible-light-driven photocatalytic performance of waxberry-like CeO2 by samarium doping and silver QDs anchoring. Appl. Catal. B Environ. 2021, 286, 119845. [Google Scholar] [CrossRef]
  25. Jiang, H.; Li, X.; Chen, S.; Wang, H.; Huo, P. g-C3N4 quantum dots-modified mesoporous CeO2 composite photocatalyst for enhanced CO2 photoreduction. J. Mater. Sci. Mater. Electron. 2020, 31, 20495–20512. [Google Scholar] [CrossRef]
  26. Sre, V.V.; Okla, M.K.; Janani, B.; Abdel-Maksoud, M.A.; Al-Amri, S.S.; Alaraidh, I.A.; Alatar, A.A.; Khan, S.S. A novel sunlight driven Z scheme ZnCdS QDs deposited over g-C3N4 photocatalyst for photoinactivation of E. coli cells. J. Water Process. Eng. 2024, 59, 104957. [Google Scholar] [CrossRef]
  27. Morgan, D.J. Photoelectron spectroscopy of ceria: Reduction, quantification and the myth of the vacancy peak in XPS analysis. Surf. Interface Anal. 2023, 55, 845–850. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Li, J.; Zhou, W.; Liu, X.; Song, X.; Chen, S.; Wang, H.; Huo, P. Rational design of Ag/CuO@ZnIn2S4 S-scheme plasmonic photocatalyst for highly selective CO2 conversion. Appl. Catal. B Environ. 2024, 342, 123449. [Google Scholar]
  29. Boaro, M.; Colussi, S.; Trovarelli, A. Ceria-Based Materials in Hydrogenation and Reforming Reactions for CO2 Valorization. Front. Chem. 2019, 7, 28. [Google Scholar] [CrossRef]
  30. Mullins, D.R. The surface chemistry of cerium oxide. Surf. Sci. Rep. 2015, 70, 42–85. [Google Scholar] [CrossRef]
  31. Rodriguez, J.A.; Grinter, D.C.; Liu, Z.; Palomino, R.M.; Senanayake, S.D. Ceria-based model catalysts: Fundamental studies on the importance of the metal–ceria interface in CO oxidation, the water–gas shift, CO2 hydrogenation, and methane and alcohol reforming. Chem. Soc. Rev. 2017, 46, 1824–1841. [Google Scholar] [CrossRef] [PubMed]
  32. Kim, S.S.; Britcher, L.; Kumar, S.; Griesser, H.J. XPS Study of Sulfur and Phosphorus Compounds with Different Oxidation States. Sains Malays. 2018, 47, 1913–1922. [Google Scholar] [CrossRef]
  33. Wu, L.; Su, F.; Liu, T.; Liu, G.-Q.; Li, Y.; Ma, T.; Wang, Y.; Zhang, C.; Yang, Y.; Yu, S.-H. Phosphorus-Doped Single-Crystalline Quaternary Sulfide Nanobelts Enable Efficient Visible-Light Photocatalytic Hydrogen Evolution. J. Am. Chem. Soc. 2022, 144, 20620–20629. [Google Scholar] [CrossRef]
  34. Sun, Y.; Hao, Y.; Lin, X.; Liu, Z.; Sun, H.; Jia, S.; Chen, Y.; Yan, Y.; Li, X. Efficient electron transport by 1D CuZnInS modified 2D Ti3C2 MXene for enhanced photocatalytic hydrogen production. J. Colloid Interface Sci. 2024, 653, 396–404. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, S.-J.; Choi, M.; Hong, G.; Hahn, S.K. Controlled afterglow luminescent particles for photochemical tissue bonding. Light Sci. Appl. 2022, 11, 314. [Google Scholar] [CrossRef] [PubMed]
  36. Zhao, J.; Wang, Y.; Li, Y.; Yue, X.; Wang, C. Phase-dependent enhancement for CO2 photocatalystic reduction over CeO2/TiO2 catalysts. Catal. Sci. Technol. 2016, 6, 7967–7975. [Google Scholar] [CrossRef]
  37. Kibar, M.E. Preparation of copper oxide-cerium oxide/nanotube-titanium dioxide photocatalyst for CO2 conversion in solar light. React. Kinet. Catal. Lett. 2021, 134, 937–950. [Google Scholar] [CrossRef]
  38. Hu, H.; Hu, J.; Wang, X.; Gan, J.; Su, M.; Ye, W.; Zhang, W.; Mac, X.; Huihu, W. Enhanced reduction and oxidation capability over CeO2/g-C3N4 hybrid through surface carboxylation: Performance and mechanism. Catal. Sci. Technol. 2020, 10, 4712–4725. [Google Scholar] [CrossRef]
  39. Li, Z.; Liu, Z.; Li, Y.; Wang, Q. Flower-like CoAl layered double hydroxides modified with CeO2 and RGO as efficient photocatalyst towards CO2 reduction. J. Alloys Compd. 2021, 881, 160650. [Google Scholar] [CrossRef]
  40. Su, B.; Kong, Y.; Wang, S.; Zuo, S.; Lin, W.; Fang, Y.; Hou, Y.; Zhang, G.; Zhang, H.; Wang, X. Hydroxyl-Bonded Ru on Metallic TiN Surface Catalyzing CO2 Reduction with H2O by Infrared Light. J. Am. Chem. Soc. 2023, 145, 27415–27423. [Google Scholar] [CrossRef]
  41. Yin, S.; Zhou, Y.; Liu, Z.; Wang, H.; Zhao, X.; Zhu, Z.; Yan, Y.; Huo, P. Elucidating protonation pathways in CO2 photoreduction using the kinetic isotope effect. Nat. Commun. 2024, 15, 437. [Google Scholar] [CrossRef] [PubMed]
  42. Xu, J.; Ju, Z.; Zhang, W.; Pan, Y.; Zhu, J.; Mao, J.; Zheng, X.; Fu, H.; Yuan, M.; Chen, H.; et al. Efficient Infrared-Light-Driven CO2 Reduction Over Ultrathin Metallic Ni-doped CoS2 Nanosheets. Angew. Chem. Int. Ed. 2021, 60, 8705–8709. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, L.; Huang, H.; Chen, Z.; Yu, H.; Wang, K.; Huang, J.; Yu, H.; Zhang, Y. Synergistic Polarization Engineering on Bulk and Surface for Boosting CO2 Photoreduction. Angew. Chem. Int. Ed. 2021, 60, 18303–18308. [Google Scholar] [CrossRef] [PubMed]
  44. 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. 2021, 408, 127292. [Google Scholar] [CrossRef]
  45. Zhu, K.; Zhu, Q.; Jiang, M.; Zhang, Y.; Shao, Z.; Geng, Z.; Wang, X.; Zeng, H.; Wu, X.; Zhang, W.; et al. Modulating Tit2g Orbital Occupancy in a Cu/TiO2 Composite for Selective Photocatalytic CO2 Reduction to CO. Angew. Chem. Int. Ed. 2022, 61, e202207600. [Google Scholar] [CrossRef]
  46. Chen, Y.; Yu, S.; Fan, X.-B.; Wu, L.-Z.; Zhou, Y. Mechanistic insights into the influence of surface ligands on quantum dots for photocatalysis. J. Mater. Chem. A 2023, 11, 8497–8514. [Google Scholar] [CrossRef]
  47. Xiao, Y.; Jiang, Y.; Zhou, E.; Zhang, W.; Liu, Y.; Zhang, J.; Wu, X.; Qi, Q.; Liu, Z. In-suit fabricating an efficient electronic transport channels via S-scheme polyaniline/Cd0.5Zn0.5S heterojunction for rapid removal of tetracycline hydrochloride and hydrogen production. J. Mater. Sci. Technol. 2023, 153, 205–218. [Google Scholar] [CrossRef]
  48. Xu, S.; Jiang, G.; Zhang, H.; Gao, C.; Chen, Z.; Liu, Z.; Wang, J.; Du, J.; Cai, B.; Li, Z. Boosting Photocatalytic CO2 Methanation through Interface Fusion over CdS Quantum Dot Aerogels. Small 2024, e2400769. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Rogers, C.R.; Weiss, E.A. Energy Transfer from CdS QDs to a Photogenerated Pd Complex Enhances the Rate and Selectivity of a Pd-Photocatalyzed Heck Reaction. J. Am. Chem. Soc. 2019, 142, 495–501. [Google Scholar] [CrossRef]
  50. Aldana, J.; Wang, Y.A.; Peng, X. Photochemical Instability of CdSe Nanocrystals Coated by Hydrophilic Thiols. J. Am. Chem. Soc. 2001, 123, 8844–8850. [Google Scholar] [CrossRef]
Scheme 1. Synthesis diagram of ZCS-QDs/CeO2.
Scheme 1. Synthesis diagram of ZCS-QDs/CeO2.
Catalysts 14 00599 sch001
Figure 1. (a) XRD patterns of CeO2, ZCS-QDs, and x ZCS-QDs/CeO2 (x = 5, 8, 10, 12, and 15%), (bd) TEM image of CeO2 and ZCS-QDs, (e) TEM image of 12% ZCS-QDs/CeO2, (f) HRTEM image of 12% ZCS-QDs/CeO2, (g) EDS elemental mapping images of 12% ZCS-QDs/CeO2.
Figure 1. (a) XRD patterns of CeO2, ZCS-QDs, and x ZCS-QDs/CeO2 (x = 5, 8, 10, 12, and 15%), (bd) TEM image of CeO2 and ZCS-QDs, (e) TEM image of 12% ZCS-QDs/CeO2, (f) HRTEM image of 12% ZCS-QDs/CeO2, (g) EDS elemental mapping images of 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g001
Figure 2. XPS spectra of (a) survey spectra, (b) Ce 3d, (c) O 1s, (d) S 2p of 12% ZCS-QDs/CeO2, (e) Cd 3d of 12% ZCS-QDs/CeO2, and (f) Zn 2p of 12% ZCS-QDs/CeO2.
Figure 2. XPS spectra of (a) survey spectra, (b) Ce 3d, (c) O 1s, (d) S 2p of 12% ZCS-QDs/CeO2, (e) Cd 3d of 12% ZCS-QDs/CeO2, and (f) Zn 2p of 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g002
Figure 3. (a) UV–Vis DRS, (b) plots of (αhv)1/2 versus (hv), (c,d) VB-XPS spectra of CeO2 and ZCS-QDs, and (e) bandgap structure diagram of CeO2 and ZCS-QDs.
Figure 3. (a) UV–Vis DRS, (b) plots of (αhv)1/2 versus (hv), (c,d) VB-XPS spectra of CeO2 and ZCS-QDs, and (e) bandgap structure diagram of CeO2 and ZCS-QDs.
Catalysts 14 00599 g003
Figure 4. (a) TR-PL decay spectra of CeO2, ZCS-QDs, and 12% ZCS-QDs/CeO2; (b) TPR and (c) EIS of CeO2, ZCS-QDs, and 12% ZCS-QDs/CeO2.
Figure 4. (a) TR-PL decay spectra of CeO2, ZCS-QDs, and 12% ZCS-QDs/CeO2; (b) TPR and (c) EIS of CeO2, ZCS-QDs, and 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g004
Figure 5. (a) Photocatalytic activity of CeO2, ZCS-QDs, and x ZCS-QDs/CeO2; (b) photocatalytic performance of 12% ZCS-QDs/CeO2 under CO2 and N2; (c) cycling test of 12% ZCS-QDs/CeO2.
Figure 5. (a) Photocatalytic activity of CeO2, ZCS-QDs, and x ZCS-QDs/CeO2; (b) photocatalytic performance of 12% ZCS-QDs/CeO2 under CO2 and N2; (c) cycling test of 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g005
Figure 6. DRIFTS spectra of CO2 adsorption of (a) ZCS-QDs, (b) CeO2, and (c) 12% ZCS-QDs/CeO2.
Figure 6. DRIFTS spectra of CO2 adsorption of (a) ZCS-QDs, (b) CeO2, and (c) 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g006
Figure 7. DRIFTS spectra of (a) CeO2, (b) ZCS-QDs, and (c) 12% ZCS-QDs/CeO2.
Figure 7. DRIFTS spectra of (a) CeO2, (b) ZCS-QDs, and (c) 12% ZCS-QDs/CeO2.
Catalysts 14 00599 g007
Figure 8. Photocatalytic CO2 reduction mechanism of 0D–2D ZCS-QDs/CeO2.
Figure 8. Photocatalytic CO2 reduction mechanism of 0D–2D ZCS-QDs/CeO2.
Catalysts 14 00599 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

Yan, J.; Sun, Y.; Cai, J.; Cai, M.; Hu, B.; Yan, Y.; Zhang, Y.; Tang, X. Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic CO2 Reduction and Mechanism Insight. Catalysts 2024, 14, 599. https://doi.org/10.3390/catal14090599

AMA Style

Yan J, Sun Y, Cai J, Cai M, Hu B, Yan Y, Zhang Y, Tang X. Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic CO2 Reduction and Mechanism Insight. Catalysts. 2024; 14(9):599. https://doi.org/10.3390/catal14090599

Chicago/Turabian Style

Yan, Junzhi, Yuming Sun, Junxi Cai, Ming Cai, Bo Hu, Yan Yan, Yue Zhang, and Xu Tang. 2024. "Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic CO2 Reduction and Mechanism Insight" Catalysts 14, no. 9: 599. https://doi.org/10.3390/catal14090599

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