Next Article in Journal
Ligand-Mediated Tuning of Pd-Au Nanoalloys for Selective H2O2 Production in Direct Synthesis from H2 and O2
Previous Article in Journal
Advanced TiO2-Based Photoelectrocatalysis: Material Modifications, Charge Dynamics, and Environmental–Energy Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Curly-like CN@CdS Z-Scheme Heterojunction to Boost Visible-Light-Driven H2O2 Evolution

1
Key Laboratory of Marine Chemistry Theory and Technology, Ministry of Education, College of Chemistry and Chemical Engineering, Ocean University of China, Qingdao 266100, China
2
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
3
Key Laboratory of Biofuels, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2025, 15(6), 543; https://doi.org/10.3390/catal15060543
Submission received: 30 April 2025 / Revised: 27 May 2025 / Accepted: 27 May 2025 / Published: 29 May 2025
(This article belongs to the Section Photocatalysis)

Abstract

Photocatalytic H2O2 production with H2O and O2 as resources is a promising technique. Herein, a g-C3N4@CdS (CN@CdS) Z-scheme heterojunction is prepared for photocatalytic H2O2 production with a rate of 167.5 μmol/h, which is 5.43 and 5.15 times higher than that of pure g-C3N4 and pure CdS. Photoelectronic characterization results reveal the existence of a strong built-in electric field between CdS and g-C3N4, which significantly facilitates the separation of photogenerated carriers and preserves the strong redox capacity. Density functional theory (DFT) calculations show that CN@CdS displays lower O2 adsorption energy and the H2O2 is more readily formed. This work provides a novel strategy for the design of photocatalysts with excellent H2O2 evolution efficiency.

1. Introduction

As one of the most important chemical substances in the world, H2O2 is widely applied in the paper industry, chemical synthesis, antimicrobial bleaching and wastewater treatment [1]. Currently, H2O2 is produced through an oxidation of anthraquinone process, which requires the direct use of hydrogen and oxygen gases and generates wastewater, exhaust gas and solid waste [2,3]. Photocatalytic H2O2 production, using H2O and O2 as raw materials and solar energy as the only driving force, has attracted attention for its green, economic and safety advantages [4].
Graphite-phase carbon nitride (g-C3N4), a graphite-phase nonmetallic polymer with tris-s-triazine as a repeating unit, is considered to be a common and effective photocatalyst. Since Wang et al. [5] first reported photocatalytic overall water splitting by g-C3N4 under visible light irradiation, g-C3N4 has been used for various fields. However, pristine g-C3N4 leads to inferior photocatalytic H2O2 production activity owing to disadvantages such as the high complexation efficiency of photogenerated carriers, small specific surface area and weak redox capacity. Meanwhile, single-component photocatalysts usually cannot simultaneously satisfy requirements for strong light absorption capacity, strong redox properties and efficient photogenerated carrier separation. Constructing a Z-scheme heterojunction has been an effective approach to improve the photocatalytic performance of g-C3N4-based photocatalysts [6]. Unlike conventional type-II heterojunctions, Z-scheme heterojunctions retain strongly reducing conduction-band electrons and strongly oxidizing valence-band holes in addition to improved charge separation efficiency, thus providing both highly efficient charge separation and strong redox capability [7].
Metal sulfides (CdS, MoS2, ZnS, etc.) have attracted much attention in the photocatalytic field due to their unique physicochemical properties [8]. In particular, CdS is a promising semiconductor material owing to its excellent band gap (≈2.4 eV) and superior response to visible light [9]. The d orbitals of Cd in CdS contribute to a significant role in the formation of the valence band (VB), which facilitates electron transfer in photocatalysis [10]. The suitable energy band position of CdS makes it an excellent choice for coupling with g-C3N4 to construct a Z-scheme heterojunction. Zhang et al. [11] prepared a novel Z-scheme O-CN/CdS heterojunction by the hydrothermal method. It is found that constructing a Z-scheme heterojunction promotes the separation of photogenerated carriers, which greatly enhances photocatalytic performance. Close interfacial contact between heterojunctions improves charge transport pathways, thereby facilitating interfacial charge separation and transfer. Lin et al. [12] designed and synthesized a novel defective C3N4/CdS direct Z-scheme heterojunction with significantly enhanced light absorption, redox properties, charge separation efficiency and carrier lifetime. However, the above Z-scheme heterojunction application process only utilizes the conduction band electrons for the target product synthesis, and the holes on the valence band of CdS are not utilized for production. Therefore, it is necessary to develop a photocatalyst that utilizes both a conduction band and valence band for the synthesis of target products.
It has been shown that co-modification can effectively improve the photocatalytic efficiency. A curly-like photocatalyst enables multiple reflections to maximize the utilization of light [13]. In this study, we constructed curly-like CN@CdS Z-scheme heterojunction photocatalysts for visible-light photocatalytic H2O2 production. The mechanism of photocatalytic H2O2 production from CN@CdS photocatalysts was revealed by DFT calculations, active group capture experiments and ESR characterization. The convoluted structure of the photocatalyst supplies more adsorption and active sites for the reaction, and the Z-scheme charge transfer pathway retains the strong redox ability of the photocatalyst. Therefore, the CN@CdS heterojunction exhibited significantly improved yields for photocatalytic H2O2 production. This study provides a rational design pathway for the construction of the Z-scheme heterojunction of g-C3N4-based photocatalysts aimed at enhancing charge separation efficiency and boosting photocatalytic H2O2 production.

2. Results

2.1. Structural Characterizations

The crystal structures of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3 were analyzed by XRD. As shown in Figure 1a, two characteristic peaks of g-C3N4 at 2θ = 12.9° and 27.7° can be observed, which correspond to the (100) and (002) crystal planes of g-C3N4 (JCPDS No. 87-1526), assigned to the in-plane structural stacking motifs of tri-s-triazine in g-C3N4 and the interplanar spacing between C3N4 layers, respectively [14]. The XRD spectrum of CdS shows diffraction peaks at 2θ = 25.1°, 26.2°, 28.3°, 43.9°, 48.0°and 51.9°, which correspond to the (100), (002), (101), (110), (103) and (112) crystal planes of the hexagonal wurtzite structure of CdS (JCPDS No. 41-1049) [15]. In addition, a series of CN@CdS photocatalysts were prepared with good crystalline phase matching with CdS and g-C3N4. For CN@CdS composites, the peak intensity of the CdS phase increases with its content. No impurity diffraction peaks are detected, indicating that no other chemical reactions occurred during the electrostatic self-assembly of g-C3N4 with CdS nanoparticles. To further understand the chemical state of CN@CdS composites, XPS was employed. The XPS survey spectrum (Figure 1b) reveals that C, N, Cd, S and O elements co-exist in CN@CdS-2. The observed presence of a small amount of elemental O can be attributed to the adsorption of H2O on the surface of CN@CdS-2. As depicted in the high-resolution spectrum of C 1s (Figure 1c), the characteristic peaks at 284.7 eV and 287.9 eV belong to graphitic carbon (C=C) and sp2-bonded carbon (N-C=N), respectively [16]. The three characteristic peaks of N 1s spectra (Figure 1d) are attributed to the sp2 hybridized nitrogen (C-N=C), N-(C)3 group and π-excited state [17]. In Figure 1e, the Cd 3d high-resolution spectra have two characteristic peaks corresponding with Cd 3d5/2 and Cd 3d3/2 of Cd2+ at 405.2 and 412.0 eV [18,19]. Similarly, the binding energies of S 2p (Figure 1f) bimodal peaks are located at 163.4 and 161.8 eV [20]. The XPS results further demonstrate the favorable loading of CdS nanoparticles on CN to form a CN@CdS heterojunction.
SEM, TEM and HRTEM were employed to observe the microscopic morphology of samples. Figure 2a shows SEM image of CdS, and it can be clearly seen that a large number of CdS nanospheres with diameters of about 100 nm were aggregated together to form CdS nanospheres with diameters of about 500 nm. Figure 2b shows the SEM image of g-C3N4, in which it is obvious that g-C3N4 has a curly structure. This structure can expose a higher surface area, which can reflect the light multiple times in the photocatalyst to maximize the utilization of the light source. In addition, it also can shorten the transmission distance of the photogenerated electrons and inhibit the complexation of the photogenerated carriers, which promotes the improvement of the photocatalytic performance. Figure 2c shows an SEM image of CN@CdS-2. After self-assembly by electrostatic force, close interfacial contacts are formed between CdS nanospheres and g-C3N4. Compared with pure CdS, the size of CdS nanospheres in CN@CdS-2 is significantly reduced, which reflects the positive effect of g-C3N4 in enhancing the dispersion of CdS and effectively suppressing the agglomeration of CdS nanospheres. Figure 2d,e show TEM images of g-C3N4 and CN@CdS-2, respectively, which are consistent with SEM results, exhibiting the curly structure of g-C3N4 and nano-microsphere structure of CdS. The HRTEM image of CN@CdS-2 is illustrated in Figure 2f, where two crystalline planes with lattice spacings of 0.34 and 0.31 nm can be observed, corresponding to the (002) and (101) crystalline planes of CdS [15,21]. The amorphous region is attributed to g-C3N4. The above results provide strong proof of the successful anchoring of CdS nanospheres on the g-C3N4 surface.

2.2. Optical and Photoelectrochemical Characterization

The light absorption properties of CdS, g-C3N4 and CN@CdS were investigated by UV-Vis DRS. The band edges of g-C3N4 and CdS are 452 and 583 nm, respectively, as seen in Figure 3a. As expected, the band edges of the CN@CdS heterojunction move to longer wavelengths with the increase in CdS loading. This indicates that the construction of the CN@CdS heterojunction can enhance the light absorption ability, which is more favorable to enhance the photocatalytic activity. The band gap (Eg) of each catalyst can be obtained by converting the UV-Vis DRS (Figure 3a) into Trac spectra (Figure 3b). The Eg can be derived from the Kubellka–Munk equation ( α h ν = A ( h ν E g ) n 2 , where α, A, h, ν and Eg are the absorption coefficient, scaling factor, Planck coefficient, optical frequency and band gap. Both g-C3N4 and CdS are direct bandgap semiconductors, so n = 1. From the above equation, it can be seen that by plots (αhν)1/2 versus hν, the slope of the curves is Eg. For g-C3N4 and CdS, the Eg are 2.75 and 2.30 eV (Figure 3b). For the CN@CdS heterojunction, the Eg of three catalysts are in between g-C3N4 and CdS. The Eg of CN@CdS-1, CN@CdS-2 and CN@CdS-3 are 2.74, 2.73 and 2.71 eV, respectively. The flat potential (Efb) of g-C3N4 and CdS are measured based on Mott Schottky (Figure 3c). The slopes of g-C3N4 and CdS are both positive, consistent with an n-type semiconductor. The Efb of g-C3N4 and CdS are −1.17 and −0.40 eV (vs. NHE), respectively. Usually the ECB of an n-type semiconductor approximates the Efb. The valence band positions (EVB) of g-C3N4 and CdS can be calculated from Eg = EVB − ECB to be 1.58 eV and 1.90 eV (vs. NHE), respectively.
DFT was used to investigate the interfacial charge transfer in the heterojunction. Combining the results of HRTEM and XRD, we selected the (101) crystal plane of CdS and the (002) crystal plane of g-C3N4 for DFT calculations. As can be observed from Figure 4a,b, the work functions (Φ) of the g-C3N4 (002) surface and CdS (101) surface are 5.657 and 6.936 eV, and the Fermi energy (Ef) level of CdS is higher than that of g-C3N4 (Figure 4c). Upon contact between the two, the CB of g-C3N4 bends upward at the interface while the CB of CdS bends downward, induced by an internal electric field (IEF). Upon photoexcitation, an e is spontaneously transferred from g-C3N4 to CdS under the effect of IEF, coulomb attraction and energy band bending until the two Fermi energy levels reach equilibrium. As a result, an e on the CB of CdS binds to h+ on the VB of g-C3N4 due to the presence of the IEF. The formation of a Z-scheme heterojunction between CdS and g-C3N4 improves the photogenerated carrier separation rate and maintains a high redox activity.
The effect of the Z-scheme heterojunction on photogenerated carrier separation efficiency has been investigated by PL and transient photocurrent response. Compared with g-C3N4, CN@CdS-2 exhibits remarkably low intensity, indicating the construction of the heterojunction effectively suppresses the complexation rate of the photogenerated carrier (Figure 5a). The PL spectrum of g-C3N4 shows a small peak at ~ 450 nm in addition to a main peak at 430 nm. The reason for the double peak may be due to the defect states derived from the structure, which form intermediate energy levels in the forbidden band, resulting in low-energy emission peaks [22]. The PL peak position also appears red-shifted, which is consistent with the results of the UV-Vis DRS. In addition, the prepared samples were also analyzed for photocurrent to compare their photogenerated carrier separation ability (Figure 5b). The highest photocurrent density is observed for CN@CdS-2, indicating the highest separation efficiency of its photogenerated carriers. All of the above results indicate that construction of the Z-scheme heterojunction can significantly enhance the separation efficiency of the photogenerated carriers and thus possibly boost the photocatalytic activity of the photocatalyst.

2.3. Photocatalytic Performance of H2O2 Production

The photocatalytic performance of prepared samples was evaluated by photocatalytic H2O2 production. The photocatalytic reaction was carried out under air conditions without adding any sacrificial agent and the specific experimental steps are illustrated in the Materials and Methods section. Figure 6a exhibits curves of H2O2 produced by g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3 with time. The concentrations of H2O2 increase with the longer irradiation time for all samples, indicating that all of them have the ability to produce H2O2 in the photocatalytic reaction system. After irradiation for 2 h, CN@CdS series samples all show superior photocatalytic H2O2 production performance to that of g-C3N4 and CdS. CN@CdS-2 exhibits the highest H2O2 yield, reaching up to 167.5 μM/h, which is 5.43 and 5.15 times higher than that of g-C3N4 and CdS, respectively. The excellent photocatalytic performance of CN@CdS may be attributed to the construction of the Z-scheme heterojunction, which enhances the light-absorption capacity and separation efficiency of photogenerated carriers. In addition, its curled structure multiplies adsorption and active sites for O2 and H2O, which promotes its visible-light photocatalytic H2O2 production performance. The yield of CN@CdS-3 is inferior to that of CN@CdS-2, which is probably due to the fact that the excessive loading of CdS will cover the active sites of the photocatalyst, leading to an inhibition of photocatalytic H2O2 production performance. To investigate whether the prepared samples could maintain high photocatalytic H2O2 production activity in real water samples, photocatalytic H2O2 production experiments were carried out in seawater and pure water. As illustrated in Figure 6b, g-C3N4 and CN@CdS-2 both have greater H2O2 yields in seawater than in pure water. CN@CdS-2 has an improved H2O2 yield from 167.5 to 287.5 μM/h. This result can be attributed to the ability of cations in seawater to play a facilitating role in photocatalytic reactions [23].
The stability of CN@CdS-2 was investigated. As demonstrated in Figure 6c, the photocatalytic H2O2 production of CN@CdS-2 can still maintain relative stability after three cycling experiments. The decrease in production during cycling may be attributed to unavoidable loss during recycling. The photocatalytic H2O2 production is normally associated with a decomposition of H2O2; hence, a kinetic model is developed to describe it. The production rate constant (Kf) and decomposition rate constant (Kd) of H2O2 are in accordance with the kinetic equations of zero and first order, respectively [24].
H 2 O 2 = K f K d 1 exp K d × t
K d = ln C t / C 0 / t
where Kf and Kd are the formation and decomposition rate constants, respectively. As shown in Figure 6d, CN@CdS-2 possesses the highest Kf value (2.804 μM/min) and the lowest Kd value (0.745 × 10−4 min−1). This indicates that the construction of a Z-scheme heterojunction is not only favorable to promote H2O2 generation but also to inhibit the decomposition of H2O2.
Table 1 demonstrates yields and AQY values of photocatalytic H2O2 production from CdS, g-C3N4 and CN@CdS-2 under 450 nm monochromatic light irradiation. As can be seen from Table 1, the AQY value of CN@CdS-2 is significantly higher than that of pristine CdS and g-C3N4, which is attributed to a Z-scheme heterojunction constructed to promote visible light absorption. In addition, we compared the performance of photocatalytic H2O2 synthesis on CN@CdS-2 with other photocatalysts reported in the literature, as shown in Table 2, demonstrating its excellent photocatalytic performance.

2.4. Possible Photocatalysis Mechanism

The pathways for photocatalytic H2O2 production include the ORR pathway and WOR pathway. Active substance capture experiments were performed to elucidate the mechanism of H2O2 generation. The experiments were carried out by bubbling Ar to eliminate O2. The results are illustrated in Figure 7a. When Ar was continuously passed into the reaction solution, the H2O2 generation decreased significantly to only 16.89% of that under saturated air. It is suggested that O2 is the main reactor for photocatalytic H2O2 production. In addition, P-benzoquinone (BQ), isopropanol (IPA), EDTA-2Na and AgNO3 were used as ·O2, ·OH, h+ and e trapping agents, respectively. The ·OH scavenger (IPA) has little inhibitory effect against H2O2 production on CN@CdS-2. The e and ·O2 capture agent significantly decreased H2O2 production, confirming that e and ·O2 are the main active species and the photocatalysts perform a two-electron reduction from O2 to produce H2O2. The H2O2 production increases by 10.18% after adding EDTA-2Na, which is attributed to the fact that EDTA-2Na consumes h+ and inhibits the complexation of photogenerated electron-hole pairs, thus allowing more e to participate in the surface reduction reaction to generate H2O2. To reveal the main contributing radical for photocatalytic H2O2 production, ESR experiments were performed using 5,5-dimethyl-1-pyrrolidine N-oxide (DMPO) as a radical trapping agent, as depicted in Figure 7b. The distinct peaks of DMPO-·O2 can be observed, confirming that H2O2 can be generated from the ORR pathway.
In addition, the Gibbs free energy diagram is calculated for the conversion of O2 to H2O2 to obtain further insight into the mechanism of highly reactive H2O2 production on CN@CdS (Figure 8). The adsorption of O2 (O2 → *O2) is the rate determining step of the 2e ORR process. During the conversion of O2 to H2O2, CN@CdS exhibits a smoother path compared to g-C3N4, suggesting a kinetic advantage [16]. In the O2 absorption process, a lower ΔG value is observed on CN@CdS (−1.77 eV) than g-C3N4 (−2.01 eV). In addition, OOH*, the key intermediate of 2e ORR, is more readily formed on CN@CdS than on g-C3N4. The relatively low ΔG value suggests that CN@CdS prefers to produce H2O2 via a two-step one-electron ORR pathway in comparison with g-C3N4.
Based on the above results, the mechanism of photocatalytic H2O2 production by the CN@CdS Z-scheme heterojunction is proposed, as shown in Figure 9. The photocatalytic H2O2 production process consists of the WOR pathway and ORR pathway, where the ORR pathway is the main source of H2O2. The high CB potential (−1.17 eV vs. NHE) on g-C3N4 is sufficient to drive a single-step two-electron ORR for the direct reduction of O2 to form H2O2 in the presence of protons (H+) (Equations (4–6)). In addition, the accumulation of holes (h+) on CdS accelerates the kinetics of the WOR pathway. h+ in VB on CdS oxidizes H2O in situ to produce H2O2 (Equation (7)). This process is accompanied by the 4e WOR process (Equation (8)). Based on the above discussion, it can be concluded that enhanced H2O2 production can be attributed to the high specific surface area of the convoluted structure and the construction of the Z-scheme heterojunction, which facilitates the spatial separation of redox centers and the exposure of active sites to promote WOR and ORR reactions.
CN @ CdS + h ν e + h +
O 2 + e · O 2
· O 2 + H + · OOH
· OOH + e + H + H 2 O 2
2 H 2 O + 2 h + H 2 O 2 + 2 H +
2 H 2 O + 4 h + O 2 + 4 H +

3. Materials and Methods

3.1. Materials

All of the reagents were purchased from Sinopharm Chemical Reagent Co. (Shanghai, China), and were not further treated. Urea (CO(NH2)2), melamine (C3H6N6), potassium iodide (KI), cadmium acetate (C4H6CdO4·2H2O), thiourea (CH4N2S), hydrochloric acid (HCl), ethylene glycol (C2H4O2) and ammonium molybdate (H8MoN2O4) were analytical reagents.

3.2. Synthesis of Photocatalysts

3.2.1. Synthesis of g-C3N4

g-C3N4 was obtained by a thermal polymerization of precursors. First, 10 g CO(NH2)2 and 2 g C3H6N6 were taken into a mortar and ground to a uniform mixture. The mixed powder was placed in a porcelain boat and placed in a tube furnace. The temperature was gradually increased to 550 °C in Ar atmosphere at a rate of 5 °C/min and held for 4 h. The product was allowed to cool naturally to room temperature. The product was ground and washed three times with deionized water to remove residual impurities. Finally, g-C3N4 was obtained by drying at 60 °C.

3.2.2. Synthesis of CdS

A total of 0.533 g C4H6CdO4·2H2O and 0.4567 g CH4N2S were dissolved in 60 mL of deionized water. After continuous stirring for 1 h, the mixed solution was transferred to a 100 mL autoclave and reacted at 180 °C for 12 h. The product was obtained by centrifugation and was washed three times alternately using deionized water and ethanol, respectively. The final CdS was obtained by drying at 60 °C overnight.

3.2.3. Synthesis of CN@CdS Z-Scheme Heterojunction

The g-C3N4@CdS (CN@CdS) Z-scheme heterojunction was obtained by an electrostatic self-assembly method as shown in Figure 10. A total of 0.1 g g-C3N4 was added into 50 mL deionized (DI) water and ultrasonically dispersed for 1 h. Subsequently, a certain amount of CdS was added proportionally into the g-C3N4 dispersion, and stirring continued for 24 h. The products were collected by centrifugation, washed with DI water and dried overnight at 60 °C, and different mass ratios of CN@CdS composites were obtained. According to the mass ratio of CdS to g-C3N4, the prepared composites were named g-C3N4@CdS-1(CN@CdS-1), g-C3N4@CdS-2(CN@CdS-2) and g-C3N4@CdS-3(CN@CdS-3), in which the percentages of the mass ratio of CdS to g-C3N4 were 10%, 15% and 20%, respectively.

3.3. Characterizations

The crystal structure of the samples was measured using a D8 Advance X-ray diffractometer from Bruker, Munich, Germany. X-ray photoelectron spectroscopy (XPS) was performed with an Esca Lab 250XI X-ray photoelectron spectrometer from Thermo Fisher Scientific, Waltham, MA, USA. The optical absorption properties of samples were analyzed by ultraviolet visible diffuse reflectance spectroscopy (UV-vis DRS) on a UV-3600 ultraviolet visible spectrometer manufactured by Shimadzu, Kyoto, Japan, with BaSO4 powder as a substrate blank, and the detection range was 200~800 nm. The S-4800 scanning electron microscope from Hitachi, Chiyoda City, Japan, was used to observe the surface morphology of the samples. TEM images were obtained using a JEM-2100 transmission electron microscope. The radical components were monitored by an electron paramagnetic resonance spectrometer (A300 ESR, Bruker, Munich, Germany). The photoluminescence (PL) spectra of the synthesized samples were analyzed by a fluorescence spectrometer, F-7000, from Hitachi, Chiyoda City, Japan, with an excitation wavelength of 366 nm.
The electrochemical properties of the samples were determined using an electrochemical workstation model CS350M (Wuhan COEETEST Instruments Co., Ltd., Wuhan, China). The tests were performed using three electrodes, a Pt electrode as the counter electrode, Ag/AgCl electrode as the reference electrode and FTO conductive glass coated with photocatalytic material as the working electrode (samples are coated with an area of 1 × 1 cm2), and 0.1 M NaSO4 solution as the electrolyte. When conducting transient photocurrent tests, the light source is a 300 W Xe lamp (CEL-HXF300, λ > 420 nm) with a switching cycle of 60 s. Mott–Schottky measurements were performed at a frequency of 1 kHz.

3.4. Photocatalytic Synthesis and Degradation of H2O2 and Analytical Methods

3.4.1. Photocatalytic Synthesis of H2O2

In the experiment, 50 mg of photocatalyst was dispersed in 50 mL of DI water, sonicated for 10 min and then placed in a 100 mL photochemical reactor. Before dark treatment for 30 min, the mixture was adjusted to pH = 3 with hydrochloric acid solution. The dark treatment process involves continuous stirring and the injection of enough air into the suspension to achieve an adsorption–dissolution equilibrium between the photocatalyst and dissolved oxygen in the water. A 300 W Xe lamp (CEL-HXF300, λ > 420 nm) was used as the light source, and the reaction temperature was controlled by a thermostatic water circulation system. After activating the light source, samples for measurement are taken every 20 min and the reaction solution is collected through a 0.22 μM PES filter head. To evaluate the stability of the photocatalyst, the photocatalyst was collected through centrifugation, washed with deionized water and subsequently dried at 60 °C for 6 h after each photocatalytic reaction experiment. Following the drying process, the samples were reintroduced into the photochemical reactor and subjected to a repeated photocatalytic H2O2 production experiment under the aforementioned experimental conditions.
The experimental procedure for apparent quantum yield (AQY) is as follows: 3 mg of photocatalyst was dispersed into 3 mL of DI water, and the pH of the suspension was adjusted to 3 with hydrochloric acid solution. Then, it was added to a test tube with a 1 × 1 cm2 area receiving light controlled using masking tape. A single-wavelength LED lamp (λ = 450 nm) was applied as the light source with a light intensity of 9.55 mW/cm2. Other experimental procedures were performed as the photocatalytic H2O2 production experiments. The AQY was calculated by the formula below:
AQY = N e N p × 100 % = 2 × n × N A × h × c S × P × t × λ × 100 %
where AQY—apparent quantum yield;
Ne—electrons involved in the photocatalytic reaction;
Np—incident photons;
n—the amount of H2O2 molecule (mol);
NA—Avogadro’s constant (6.022 × 1023/mol);
h—Planck’s constant (6.626 × 10−34 J·s);
c—the speed of light (3 × 108 m/s).
S—the area of irradiation (cm2);
P—the energy density of incident light (W/cm2);
t—the time of photocatalytic reaction (s);
λ—the wavelength of monochromatic light (nm).

3.4.2. Photocatalytic Decomposition of H2O2

The experimental conditions for the photocatalytic decomposition of H2O2 are identical to photocatalytic H2O2 production. 50 mg of photocatalyst was dispersed in 50 mL of H2O2 (0.5 mM) solution followed by ultrasonication for 10 min and then placement in a photochemical reactor. After being stirred in the dark for 30 min, the light source was activated and a 1 mL suspension was extracted every 20 min to measure the concentration of H2O2.

3.4.3. Analytical Methods

The concentration of H2O2 was determined according to the iodometric method [13]. The working curve can be established by configuring H2O2 solutions with different standard concentrations, as illustrated in Figure 11. Experimental details are as follows: 1 mL of reacted solution was added to 2 mL of 0.1 M KI solution and 50 μL of 0.01 M (NH4)2MoO4 solution, and the absorbance of the solution was measured by a UV-visible spectrophotometer at the wavelength of 350 nm. The resulting H2O2 concentration was calculated using a standard curve.

3.5. Computational Detail

In this study, the Perdew–Burke–Ernzerhof (PBE) exchange-correlation generalization and generalized gradient approximation generalization (GGA) in the Materials Studio software (2018) were used and a plane-wave truncation energy of 540 eV was set to perform the DFT calculation [30]. For each photocatalyst single-cell model, a 3 × 3 × 1 k-point grid was used for sampling and geometry optimization in the self-consistent field (SCF) calculations, while a 4 × 4 × 1 k-point grid was chosen for the density of states (DOS) calculations. The convergence thresholds for energy and force were set to 10−4 eV and 0.02 eV Å−1, respectively. To improve the accuracy and reliability of the calculations, we also used the DFT + U method to correct for van der Waals forces, and applied dipole corrections, the cancellation of symmetry settings and spin polarization. The calculation model for photocatalysts is shown in Figure 12.

4. Conclusions

In conclusion, CN@CdS Z-scheme heterojunctions with high photocatalytic activity and stability have been constructed by in situ anchoring CdS nanoparticles onto curly-like g-C3N4. The optimized CN@CdS shows significantly enhanced visible-light-driven photocatalytic H2O2 production, which is superior to that of pure CdS and g-C3N4. The convoluted structure of g-C3N4 provides abundant surface-anchored sites, which restrict the agglomeration of CdS nanoparticles. More importantly, the well-matched energy band structure and close contact between CdS and g-C3N4 induced the formation of IEF, which accelerated the photogenerated electron migration. The Z-scheme heterojunction has been constructed, which not only achieves a spatial separation of photogenerated carriers, but also maintains strong oxidation and reduction capabilities. The results of DFT, the ESR test and trapping experiments show that CN@CdS generates H2O2 through the dual channels of the ORR and WOR pathways, dominated by the ORR pathway. This study provides promising insights for the rational design of photocatalysts for efficient photocatalytic H2O2 production.

Author Contributions

Conceptualization, L.W., B.Y. and W.F.; methodology, L.W., B.Y. and W.F.; software, H.B. and D.S.; validation, L.W., D.S., B.Y. and W.F.; formal analysis, D.S.; investigation, B.Y. and W.F.; resources, L.W. and C.L.; data curation, D.S., B.Y. and W.F.; writing—original draft preparation, W.F.; writing—review and editing, L.W., D.S., B.Y. and W.F.; visualization, L.W.; supervision, L.W.; project administration, L.W.; funding acquisition, L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities (202364004), National Natural Science Foundation of China (No. 22002146) and the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (Grant No. 2018-k21, 2021-K15).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MDPIMultidisciplinary Digital Publishing Institute
DOAJDirectory of open access journals
TLAThree-letter acronym
LDLinear dichroism

References

  1. Qiu, J.; Dai, D.; Yao, J. Tailoring metal–organic frameworks for photocatalytic H2O2 production. Coordin. Chem. Rev. 2024, 501, 215597. [Google Scholar] [CrossRef]
  2. Fang, W.; Wang, L. S-Scheme heterojunction photocatalyst for photocatalytic H2O2 production: A review. Catalysts 2023, 13, 1325. [Google Scholar] [CrossRef]
  3. Chen, Z.; Yao, D.; Chu, C.; Mao, S. Photocatalytic H2O2 production systems: Design strategies and environmental applications. Chem. Eng. J. 2023, 451, 138489. [Google Scholar] [CrossRef]
  4. Zhao, Y.; Li, X.; Fan, X.; Wang, H.; Liu, Y.; Chen, Y.; Yang, T.; Ye, J.; Huang, H.; Li, H.; et al. Small-molecule catalyzed H2O2 production via a phase-transfer photocatalytic process. Appl. Catal. B-Environ. 2022, 314, 121499. [Google Scholar] [CrossRef]
  5. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2008, 8, 76–80. [Google Scholar] [CrossRef]
  6. Zhao, Y.; Liu, Y.; Cao, J.; Wang, H.; Shao, M.; Huang, H.; Liu, Y.; Kang, Z. Efficient production of H2O2 via two-channel pathway over ZIF-8/C3N4 composite photocatalyst without any sacrificial agent. Appl. Catal. B-Environ. 2020, 278, 119289. [Google Scholar] [CrossRef]
  7. Liao, Z.; Du, J.; Wang, L.; Yang, M.; Zhao, L.; Zhan, X.; Yang, H.; Yang, W.; Hou, H. Advances in Z-scheme photocatalysts for efficient photocatalytic H2O2 production: A comprehensive review. J. Alloys Compd. 2024, 976, 173322. [Google Scholar] [CrossRef]
  8. Mamiyev, Z.; Balayeva, N.O. Metal Sulfide Photocatalysts for Hydrogen Generation: A Review of Recent Advances. Catalysts. 2022, 12, 1316. [Google Scholar] [CrossRef]
  9. Tang, Y.; Ye, F.; Li, B.; Yang, T.; Yang, F.; Qu, J.; Yang, X.; Cai, Y.; Hu, J. Electronic structure modulation of oxygen-enriched defective CdS for efficient photocatalytic H2O2 production. Small 2024, 20, 2400376. [Google Scholar] [CrossRef]
  10. Wang, K.; Xing, Z.; Meng, D.; Zhang, S.; Li, Z.; Pan, K.; Zhou, W. Hollow MoSe2@Bi2S3/CdS core-shell nanostructure as dual Z-scheme heterojunctions with enhanced full spectrum photocatalytic-photothermal performance. Appl. Catal. B-Environ. 2021, 281, 119482. [Google Scholar] [CrossRef]
  11. Zhang, Z.; Ji, R.; Sun, Q.; He, J.; Chen, D.; Li, N.; Li, H.; Marcomini, A.; Xu, Q.; Lu, J. Enhanced photocatalytic degradation of 2-chlorophenol over Z-scheme heterojunction of CdS-decorated oxygen-doped g-C3N4 under visible-light. Appl. Catal. B-Environ. 2023, 324, 122276. [Google Scholar] [CrossRef]
  12. Wang, Z.; Wang, Z.; Zhu, X.; Ai, C.; Zeng, Y.; Shi, W.; Zhang, X.; Zhang, H.; Si, H.; Li, J.; et al. Photodepositing CdS on the active cyano groups decorated g-C3N4 in Z-Scheme manner promotes visible-light-driven hydrogen evolution. Small 2021, 17, 2102699. [Google Scholar] [CrossRef] [PubMed]
  13. Zhou, L.; Feng, J.; Qiu, B.; Zhou, Y.; Lei, J.; Xing, M.; Wang, L.; Zhou, Y.; Liu, Y.; Zhang, J. Ultrathin g-C3N4 nanosheet with hierarchical pores and desirable energy band for highly efficient H2O2 production. Appl. Catal. B-Environ. 2020, 267, 118396. [Google Scholar] [CrossRef]
  14. Zhu, Q.; Shi, Y.; Tao, Y.; Fu, L.; Zhang, C.; Shang, H.; Zhang, D.; Li, G. Binary heteroatom dopants enable carbon-based heterostructures for efficient photoelectrocatalytic H2O2 synthesis in a wide pH range. Chem. Eng. J. 2024, 482, 149066. [Google Scholar] [CrossRef]
  15. Xu, H.; Wu, L.; Jin, L.; Wu, K. Combination mechanism and enhanced visible-light photocatalytic activity and stability of CdS/g-C3N4 heterojunctions. J. Mater. Sci. Technol. 2017, 33, 30–38. [Google Scholar] [CrossRef]
  16. Yu, W.; Zhu, Z.; Hu, C.; Lin, S.; Wang, Y.; Wang, C.; Tian, N.; Zhang, Y.; Huang, H. Point-to-face Z-scheme junction Cd0.6Zn0.4S/g-C3N4 with a robust internal electric field for high-efficiency H2O2 production. J. Mater. Chem. A 2023, 11, 6384–6393. [Google Scholar] [CrossRef]
  17. Nguyen, T.K.A.; Pham, T.-T.; Gendensuren, B.; Oh, E.-S.; Shin, E.W. Defect engineering of water-dispersible g-C3N4 photocatalysts by chemical oxidative etching of bulk g-C3N4 prepared in different calcination atmospheres. J. Mater. Sci. Technol. 2022, 103, 232–243. [Google Scholar] [CrossRef]
  18. Zhao, T.; Xing, Z.; Xiu, Z.; Li, Z.; Yang, S.; Zhou, W. Oxygen-doped MoS2 nanospheres/CdS quantum dots/g-C3N4 nanosheets super-architectures for prolonged charge lifetime and enhanced visible-light-driven photocatalytic performance. ACS Appl. Mater. Inter. 2019, 11, 7104–7111. [Google Scholar] [CrossRef]
  19. Liu, Y.; Dai, F.; Zhao, R.; Huai, X.; Han, J.; Wang, L. Aqueous synthesis of core/shell/shell CdSe/CdS/ZnS quantum dots for photocatalytic hydrogen generation. J. Mater. Sci. 2019, 54, 8571–8580. [Google Scholar] [CrossRef]
  20. Guo, Y.; Huang, H.; He, Y.; Tian, N.; Zhang, T.; Chu, P.K.; An, Q.; Zhang, Y. In situ crystallization for fabrication of a core–satellite structured BiOBr–CdS heterostructure with excellent visible-light-responsive photoreactivity. Nanoscale. 2015, 7, 11702–11711. [Google Scholar] [CrossRef]
  21. Huang, M.; Liu, C.; Cui, P.; Wu, T.; Feng, X.; Huang, H.; Zhou, J.; Wang, Y. Facet-dependent photoinduced transformation of cadmium sulfide (CdS) nanoparticles. Environ. Sci. Technol. 2021, 55, 13132–13141. [Google Scholar] [CrossRef] [PubMed]
  22. Jiang, Y.; Sun, Z.; Tang, C.; Zhou, Y.; Zeng, L.; Huang, L. Enhancement of photocatalytic hydrogen evolution activity of porous oxygen doped g-C3N4 with nitrogen defects induced by changing electron transition. Appl. Catal. B-Environ. 2019, 240, 30–38. [Google Scholar] [CrossRef]
  23. Chu, C.; Yao, D.; Chen, Z.; Liu, X.; Huang, Q.; Li, Q.; Mao, S. Cyano-regulated organic polymers for highly efficient photocatalytic H2O2 production in various actual water bodies. Small 2023, 19, 2303796. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, X.; Zeng, Y.; Shi, W.; Tao, Z.; Liao, J.; Ai, C.; Si, H.; Wang, Z.; Fisher, A.C.; Lin, S. S-scheme heterojunction of core–shell biphase (1T-2H)-MoSe2/TiO2 nanorod arrays for enhanced photoelectrocatalytic production of hydrogen peroxide. Chem. Eng. J. 2022, 429, 131312. [Google Scholar] [CrossRef]
  25. Meng, K.; Zhang, J.; Cheng, B.; Ren, X.; Xia, Z.; Xu, F.; Zhang, L.; Yu, J. Plasmonic near-infrared-response S-scheme ZnO/CuInS2 photocatalyst for H2O2 production coupled with glycerin oxidation. Adv. Mater. 2024, 36, 2406460. [Google Scholar] [CrossRef]
  26. Lai, C.; Xu, M.; Xu, F.; Li, B.; Ma, D.; Li, Y.; Li, L.; Zhang, M.; Huang, D.; Tang, L.; et al. An S-scheme CdS/K2Ta2O6 heterojunction photocatalyst for production of H2O2 from water and air. Chem. Eng. J. 2023, 452, 139070. [Google Scholar] [CrossRef]
  27. Zhao, X.; You, Y.; Huang, S.; Wu, Y.; Ma, Y.; Zhang, G.; Zhang, Z. Z-scheme photocatalytic production of hydrogen peroxide over Bi4O5Br2/g-C3N4 heterostructure under visible light. Appl. Catal. B-Environ. 2020, 278, 119251. [Google Scholar] [CrossRef]
  28. Su, P.; Zhang, J.; Zhou, Y.; Wei, Z.; Zhao, S.; Yang, B.; Zhao, X.; Chen, J. Efficient photocatalytic production of hydrogen peroxide by Z-scheme resorcinol-formaldehyde resin/g-C3N4 heterostructure under visible light. Chem. Eng. J. 2023, 454, 140504. [Google Scholar] [CrossRef]
  29. Li, X.; Ye, F.; Zhang, H.; Ahmad, M.; Zeng, Z.; Wang, S.; Wang, S.; Gao, D.; Zhang, Q. Ternary rGO decorated W18O49@g-C3N4 composite as a full-spectrum-responded Z-scheme photocatalyst for efficient photocatalytic H2O2 production and water disinfection. J. Environ. Chem. Eng. 2023, 11, 110329. [Google Scholar] [CrossRef]
  30. Yu, X.; Xie, J.; Liu, Q.; Dong, H.; Li, Y. The origin of enhanced photocatalytic activity in g-C3N4/TiO2 heterostructure revealed by DFT calculations. J. Colloid Interf. Sci. 2021, 593, 133–141. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3. (b) The XPS survey spectra. High-resolution (c) C 1s, (d) N 1s, (e) Cd 3d and (f) S 2p XPS spectra of CN@CdS-2.
Figure 1. (a) XRD patterns of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3. (b) The XPS survey spectra. High-resolution (c) C 1s, (d) N 1s, (e) Cd 3d and (f) S 2p XPS spectra of CN@CdS-2.
Catalysts 15 00543 g001
Figure 2. SEM images of (a) CdS, (b) g-C3N4 and (c) CN@CdS-2. TEM images of (d) g-C3N4 and (e) CN@CdS-2. (f) HRTEM image of CN@CdS-2.
Figure 2. SEM images of (a) CdS, (b) g-C3N4 and (c) CN@CdS-2. TEM images of (d) g-C3N4 and (e) CN@CdS-2. (f) HRTEM image of CN@CdS-2.
Catalysts 15 00543 g002
Figure 3. (a) The UV-Vis DRS, (b) plots of (αhν)1/2 versus hν of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3, and Mott Schottky plots of (c) g-C3N4 and (d) CdS.
Figure 3. (a) The UV-Vis DRS, (b) plots of (αhν)1/2 versus hν of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3, and Mott Schottky plots of (c) g-C3N4 and (d) CdS.
Catalysts 15 00543 g003
Figure 4. Calculated work functions for the (a) g-C3N4 (002) plane, (b) CdS (101) plane and (c) charge-transfer processes in the Z-scheme CN@CdS-2 heterojunction.
Figure 4. Calculated work functions for the (a) g-C3N4 (002) plane, (b) CdS (101) plane and (c) charge-transfer processes in the Z-scheme CN@CdS-2 heterojunction.
Catalysts 15 00543 g004
Figure 5. (a) PL spectra and (b) the transient photocurrent response of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3.
Figure 5. (a) PL spectra and (b) the transient photocurrent response of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3.
Catalysts 15 00543 g005
Figure 6. (a) Photocatalytic H2O2 production of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3. (b) Photocatalytic H2O2 production in pure water and sea water on g-C3N4 and CN@CdS-2. (c) Cyclic experiments of, and CN@CdS-2 for, H2O2 production. (d) Formation rate constants (Kf) and decomposition rate constants (Kd) of H2O2 on g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3.
Figure 6. (a) Photocatalytic H2O2 production of g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3. (b) Photocatalytic H2O2 production in pure water and sea water on g-C3N4 and CN@CdS-2. (c) Cyclic experiments of, and CN@CdS-2 for, H2O2 production. (d) Formation rate constants (Kf) and decomposition rate constants (Kd) of H2O2 on g-C3N4, CdS, CN@CdS-1, CN@CdS-2 and CN@CdS-3.
Catalysts 15 00543 g006
Figure 7. (a) Free radical trapping results and (b) ESR spectrum for·O2 radicals generated over CN@CdS-2.
Figure 7. (a) Free radical trapping results and (b) ESR spectrum for·O2 radicals generated over CN@CdS-2.
Catalysts 15 00543 g007
Figure 8. Gibbs free energy diagram for the conversion of O2 to H2O2 by ORR process on g-C3N4 and CN@CdS.
Figure 8. Gibbs free energy diagram for the conversion of O2 to H2O2 by ORR process on g-C3N4 and CN@CdS.
Catalysts 15 00543 g008
Figure 9. Mechanistic diagram of photocatalytic H2O2 production by Z-scheme CN@CdS heterojunction.
Figure 9. Mechanistic diagram of photocatalytic H2O2 production by Z-scheme CN@CdS heterojunction.
Catalysts 15 00543 g009
Figure 10. Diagram illustrating the formation of CN@CdS Z-scheme heterojunction.
Figure 10. Diagram illustrating the formation of CN@CdS Z-scheme heterojunction.
Catalysts 15 00543 g010
Figure 11. The work curve of the concentration of H2O2 and absorption.
Figure 11. The work curve of the concentration of H2O2 and absorption.
Catalysts 15 00543 g011
Figure 12. Calculation model for photocatalysts.
Figure 12. Calculation model for photocatalysts.
Catalysts 15 00543 g012
Table 1. H2O2 production and AQY (%) of g-C3N4, CdS and CN@CdS-2 at 450 nm.
Table 1. H2O2 production and AQY (%) of g-C3N4, CdS and CN@CdS-2 at 450 nm.
Photocatalystg-C3N4CdSCN@CdS-2
H2O2 production (μM)38.9713.9786.7
AQY (%)1.010.362.24
Table 2. Comparison of photocatalytic H2O2 synthesis performance of CN@CdS-2 with that of reported photocatalysts.
Table 2. Comparison of photocatalytic H2O2 synthesis performance of CN@CdS-2 with that of reported photocatalysts.
PhotocatalystH2O2 Yield(μmol g−1 h−1)AQY (%)Refs
CN@CdS-2287.52.24 (450 nm)This work
ZnO/CuInS2911.2 (365 nm)[25]
CdS/K2Ta2O6160.89-[26]
Bi4O5Br2/g-C3N412411.8 (420 nm)[27]
Resorcinol-formaldehyde Resin/g-C3N4140-[28]
rGO decorated W18O49@g-C3N4 (r-CNW-2)49.41.2 (420 nm)[29]
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

Yu, B.; Fang, W.; Bai, H.; Li, C.; Shen, D.; Wang, L. Construction of Curly-like CN@CdS Z-Scheme Heterojunction to Boost Visible-Light-Driven H2O2 Evolution. Catalysts 2025, 15, 543. https://doi.org/10.3390/catal15060543

AMA Style

Yu B, Fang W, Bai H, Li C, Shen D, Wang L. Construction of Curly-like CN@CdS Z-Scheme Heterojunction to Boost Visible-Light-Driven H2O2 Evolution. Catalysts. 2025; 15(6):543. https://doi.org/10.3390/catal15060543

Chicago/Turabian Style

Yu, Bingkun, Weili Fang, Hongcun Bai, Chunhu Li, Dongcai Shen, and Liang Wang. 2025. "Construction of Curly-like CN@CdS Z-Scheme Heterojunction to Boost Visible-Light-Driven H2O2 Evolution" Catalysts 15, no. 6: 543. https://doi.org/10.3390/catal15060543

APA Style

Yu, B., Fang, W., Bai, H., Li, C., Shen, D., & Wang, L. (2025). Construction of Curly-like CN@CdS Z-Scheme Heterojunction to Boost Visible-Light-Driven H2O2 Evolution. Catalysts, 15(6), 543. https://doi.org/10.3390/catal15060543

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