Rational constructing of Zn0.5Cd0.5S-diethylenetriamine/g-C3N4 S-scheme heterojunction with enhanced photocatalytic H2O2 production

Zheng Liu Yuqing Bian Graham Dawson Jiawei Zhu Kai Dai

Citation:  Zheng Liu, Yuqing Bian, Graham Dawson, Jiawei Zhu, Kai Dai. Rational constructing of Zn0.5Cd0.5S-diethylenetriamine/g-C3N4 S-scheme heterojunction with enhanced photocatalytic H2O2 production[J]. Chinese Chemical Letters, 2025, 36(9): 111272. doi: 10.1016/j.cclet.2025.111272 shu

Rational constructing of Zn0.5Cd0.5S-diethylenetriamine/g-C3N4 S-scheme heterojunction with enhanced photocatalytic H2O2 production

English

  • As environmental pollution and the global energy crisis escalate, the search for sustainable energy sources and efficient environmental management strategies has become a critical focus of scientific research [1-6]. Photocatalysis, a green technology that harnesses solar energy to drive chemical reactions, has garnered considerable attention due to its potential in both energy generation and environmental remediation [7-10]. Specifically, photocatalytic hydrogen peroxide (H2O2) production (PHP) has emerged as a promising alternative to traditional chemical synthesis methods, owing to its low energy consumption and eco-friendly characteristics [11]. The efficiency of PHP is primarily influenced by the light absorption capacity, the separation efficiency of electron-hole pairs, and its surface activity sites [12,13]. However, conventional photocatalysts such as TiO2 are limited by their poor absorption in the visible light spectrum, which restricts their practical application, leading researchers to focus on semiconductors with narrower band gaps. Graphitic carbon nitride (CN) is a facilely synthesized, cost-effective non-metallic two-dimensional polymer material [14-16]. Its high stability, non-toxicity, suitable band gap, and robust redox capabilities render it one of the most promising photocatalysts for PHP [17]. However, its limited photocatalytic activity is still constrained by the rapid recombination of photo generated charge carriers. Specifically, the high concentration of non-conjugated groups within CN hinders charge transfer rates, resulting in a deficiency of active reaction sites. Realizing efficient H2O2 photosynthesis with CN in a sacrificial free system remains a challenge [18].

    To overcome these limitations, strategies such as constructing heterostructures, elemental doping, and dual-atom engineering have been employed [19,20]. While these methods partially modify the internal electric field of CN and enhance its photocatalytic activity for PHP, the formation of heterojunctions between CN and metal sulfides markedly improves light absorption, accelerates charge separation and transfer rates, and boosts overall photocatalytic performance. Among these approaches, the binary metal sulfide Zn1-xCdxS, a solid solution of CdS and ZnS, stands out due to the identical crystal coordination patterns and similar ionic radii of CdS and ZnS [21,22]. By adjusting the Zn/Cd molar ratio, semiconductor photocatalysts with tailored band gaps can be synthesized, facilitating the formation of well-aligned band structures at the interface with CN. This band alignment is crucial for optimizing charge transfer and minimizing recombination, thereby enhancing the efficiency of PHP [23-25]. Despite the significant potential of these two pure materials to overcome existing challenges through the exploration of new building blocks and the appropriate arrangement and combination of existing ones, constructing S-scheme heterojunctions appears to be a straightforward and effective strategy [26-30]. S-scheme heterojunctions are composed of reductive photocatalysts (RP) and oxidative photocatalysts (OP) with staggered energy levels. Compared to OP, RP possess higher conduction band (CB) and valence band (VB) positions: The close contact between their Fermi levels generates a strong internal electric field (IEF) that directs the band bending from RP to OP at the interface [31-37]. Under illumination, photogenerated charge carriers with low redox potential undergo recombination facilitated by the IEF and band bending, while those with high redox potential are retained in the reaction. Consequently, the formation of S-scheme heterostructures significantly enhances the separation and transport efficiency of charge carriers as well as their redox capabilities [38-40]. Due to these superior characteristics, S-scheme heterojunctions are widely applied in photocatalytic pollutant degradation, hydrogen evolution reactions (HER), and CO2 reduction reactions (CO2RR), among others [41]. Therefore, constructing an S-scheme heterojunction is an effective strategy for photocatalytic PHP [42-44]. In this study, we designed and synthesized an organic amine constrained ZCS-D/CN S-scheme heterojunction for efficient H2O2 photosynthesis. The optimized ZCS/CN achieved an efficient H2O2 yield of 5124 µmol g-1 h-1 in a sacrificial agent free system, which is significantly higher than the yield obtained from pure CN (24 µmol g-1 h-1) and pure ZCS-D (4012 µmol g-1 h-1). These results indicate that constructing S-scheme heterojunctions can achieve effective separation of charge carriers and optimization of redox capabilities, which has great potential for efficient H2O2 photosynthesis in sacrificial free systems.

    The synthesis process of the ZCS-D/CN catalyst is illustrated in Fig. 1a. Firstly, melamine was placed in a muffle furnace, where it was heated at a rate of 5 ℃ per minute to 550 ℃ and maintained at this temperature for 4 h, resulting in the calcined CN precursor [45]. Using a hydrothermal method, ZCS-D/CN catalyst was synthesized, during which ZCS-D nanoparticles were anchored onto the CN surface. As depicted in Fig. 1b, CN exhibits two characteristic diffraction peaks at 2θ values of 13.1° and 27.5°, which correspond to the (100) crystal plane of the uniformly triazine-based structure and the (002) crystal plane of the graphitic-like structure, respectively. The diffraction peaks of pure ZCS-D nanoparticles also match those of the hexagonal solid solution ZCS-D. Furthermore, the ZCS-D/CN composite displays diffraction peaks corresponding to both pure CN and pure ZCS-D characteristic peaks, confirming the successful formation of the composite heterostructure [46]. In FTIR spectrum (Fig. 1c), the peaks of CN in the range of 2900–3400 cm-1 correspond to the stretching vibrations of N—H bonds, while the characteristic peak at 1624.5 cm-1 corresponds to the stretching vibrations of C=N bonds. Compared to pure CN, the ZCS-D/CN composite shows shift in these peak positions, indicating that the chemical environment of N—H and C=N bonds changed after the formation of the composite. The absorption peaks in the range of 1050–1600 cm-1 are attributed to the stretching vibrations of aromatic C—N heterocycles, and these peaks also exhibit shifts in the composite material, which may be related to the introduction of ZCS-D. At 810 cm-1, the characteristic peak corresponds to the out-of-plane bending vibrations of the triazine unit, and the peak in the composite material shows a red shift, suggesting that the chemical environment or bond strength of the triazine structure has been altered. For the ZCS-D sample, the FTIR peaks in the range of 660–1150 cm-1 are attributed to the stretching vibrations of Zn-S bonds, while the absorption peak at 620 cm-1 corresponds to the stretching vibrations of Cd-S bonds. In the FTIR spectrum of the ZCS-D/CN composite, characteristic peaks of both CN and ZCS-D are observed, and their positions exhibit significant shifts compared to those of pure CN and ZCS-D. These changes indicate strong chemical interactions between the two components in the composite material, confirming the successful synthesis of the composite [47,48].

    Figure 1

    Figure 1.  (a) Schematic diagram of the synthesis pathway for ZCS-D/CN. (b) XRD pattern of CN, ZCS-D/CN, and ZCS-D. (c) FTIR spectra of CN and the ZCS-D/CN. (d) TEM and (e) HRTEM images of ZCS-D/CN. (f, g) Elemental mapping of ZCS-D/CN.

    To gain an in-depth understanding of the microscopic features of the photocatalyst, SEM and TEM were employed to characterize its morphology. It can be observed that CN exhibits an irregular layered structure (Fig. S1 in Supporting information), while ZCS-D sample displays a nanoparticle-like structure. The composite material shows nanoparticles loaded onto the layered structure, confirming the successful construction of the composite. In the HRTEM images, CN does not show distinct lattice fringes, indicating its disordered structure and lower crystallinity (Figs. 1d and e). On the other hand, ZCS exhibits clear lattice fringes with an interplanar spacing of 0.328 nm, corresponding to the (202) plane of ZCS. This phenomenon may be attributed to structural changes within the composite, resulting in a more ordered structure that facilitates the migration and transport of photo-generated charges within the composite material. Elemental mapping images reveal the uniform distribution of each constituent element in ZCS-D/CN composite, thereby confirming the successful synthesis of the composite material (Figs. 1f and g) [49,50].

    In XPS spectra of CN, ZCS-D, and ZCS-D/CN composites, no impurity elements were detected, indicating the high purity of the synthesized samples (Fig. 2a). In the high-resolution XPS spectrum of Zn 2p (Fig. 2b), two prominent peaks at 1045.02 and 1022.18 eV were observed, corresponding to Zn 2p1/2 and Zn 2p3/2, respectively. For the XPS spectrum of Cd 3d (Fig. 2c), distinct peaks appeared at 411.59 and 404.86 eV, which are assigned to Cd 3d5/2 and Cd 3d3/2. In the S 2p XPS spectrum (Fig. 2d), two significant peaks were detected at 162.45 and 161.24 eV, corresponding to S 2p3/2 and S 2p1/2, respectively. The C 1s XPS spectrum (Fig. 2e) revealed prominent peaks located at 284.62, 285.70, and 288.58 eV, which can be attributed to C=C, C–N, and C=O bonds. In the N 1s XPS spectrum (Fig. 2f), distinct peaks were observed at 401.07, 400.15, and 398.71 eV, corresponding to π-bonded nitrogen, C=N–C, and C–N bonds, respectively. Due to the formation of the S-scheme heterojunction between CN and ZCS-D, ZCS-D serves as the reduction site and donates electrons, leading to an increase in binding energy. Consequently, the Zn 2p, Cd 3d, and S 2p peaks in the composites exhibit a positive shift compared to pure ZCS-D. Conversely, CN, as the oxidation site, accepts electrons, causing a decrease in binding energy and resulting in a negative shift of the C 1s and N 1s peaks in the composites. Upon light irradiation, the binding energy of the reduction site decreases, while that of the oxidation site increases. As a result, the Zn 2p, Cd 3d, and S 2p peaks shift negatively, whereas the C 1s and N 1s peaks shift positively. These shifts in peak positions indicate the strong interaction between ZCS-D and CN within the composites, as well as the altered electronic environment under light irradiation. The positive shift suggests electron withdrawal or changes in bonding environments, while the negative shift implies electron donation or an increase in electron density [51].

    Figure 2

    Figure 2.  (a) The XPS spectra of pristine CN, ZCS-D/CN, and ZCS-D. The survey spectra: (b) Zn 2p, (c) Cd 3d, (d) S 3d, (e) C 1s, (f) N 1s.

    As shown in Fig. S2 (Supporting information), the catalyst exhibits a type IV isotherm. The specific surface areas of the two pure materials, CN and ZCS-D, are 1.72 and 4.29 m2/g, while the specific surface area of the composite material ZCS-D/CN is 7.58 m2/g, which is significantly higher than that of the two pure materials (Fig. S3 in Supporting information). When ZCS-D is in situ grown on the surface of CN, a strong interaction between them is formed, which greatly increases the specific surface area of the material. This provides more adsorption and reaction sites on the surface, enhances light absorption, and is beneficial for improving the efficiency of photocatalytic reactions [52,53].

    Fig. 3a presents the UV–vis DRS spectra of CN, ZCS-D, and the ZCS-D/CN composite. The band absorption edges for CN and ZCS-D are approximately 450 nm and 540 nm. Compared to pure CN, the ZCS-D/CN composite exhibits a significant red shift in the absorption edge and a marked increase in absorption intensity. The bandgap energies (Eg) of the samples were determined using the transformed Kubelka-Munk function, as shown in Fig. 3b For indirect and direct bandgap semiconductors, the parameter n in the Kubelka-Munk function is 2 and 1/2 (Eq. S1 in Supporting information). The calculated bandgap energies of CN and ZCS-D are 2.72 eV and 2.37 eV, respectively. The flat-band potentials (EFB) were measured using Mott-Schottky (M-S) plots, yielding values of −0.83 V for CN and −0.96 V for ZCS-D (Figs. S4 and S5 in Supporting information). A positive slope in the Mott-Schottky plots confirms that the catalysts are n-type semiconductors. For n-type semiconductors, EFB is typically located near the conduction band (CB), the conduction band edge (ECB) is obtained by subtracting 0.1 eV from the calculated EFB. The ECB of CN and ZCS are converted to −0.34 V and −0.45 V (vs. RHE, pH 6.9), respectively, through Eq. S2 (Supporting information). Considering the bandgap energies of CN and ZCS-D, the valence band potentials (EVB) were calculated using the following equation (Eq. S3 in Supporting information). The calculated EVB values for CN and ZCS-D are 2.38 V and 1.93 V, respectively. Based on this quantitative analysis, the band structure diagrams for CN and ZCS-D are illustrated in Fig. 3c. Compared to the standard reduction potential of O2/O2 (superoxide radicals), the ZCS-D/CN composite exhibits a more negative conduction band position −0.33 V vs. RHE). This implies that photogenerated electrons are more readily transferred to the conduction band under illumination, facilitating the generation of more active O2 radicals and thereby accelerating the catalytic reaction. Additionally, compared to the pure materials, the heterojunction provides a more negative conduction band position, promoting electron transfer and mobility, which enhances charge separation and electron transfer capabilities [54,55]. As shown in Fig. S6 (Supporting information) presents the photocurrent responses of pristine CN, ZCS-D, and the ZCS-D/CN composite electrodes. Both pristine CN and ZCS-D exhibit relatively low photocurrent densities, which can be attributed to the rapid recombination of photoexcited charge carriers. In contrast, the ZCS-D/CN composite demonstrates a significantly enhanced photocurrent density, indicating effective charge separation and improved electron transfer within the composite material. As illustrated in Fig. S7 (Supporting information), ZCS-D/CN composite exhibits the smallest impedance radius, demonstrating that the ZCS-D/CN composite exhibits the fastest charge transfer rate. This suggests that the formation of the composite reduces carrier transfer resistance, accelerates charge transfer, and thereby enhances photocatalytic activity. During the photocatalytic reaction process, the efficiency of charge separation and transfer is a critical factor determining material performance. PL spectra of CN, ZCS-D, and ZCS-D/CN composite under an excitation wavelength of 320 nm are shown in Fig. S8 (Supporting information). The pristine CN and ZCS-D exhibit prominent emission peaks at 550–560 nm, which are attributed to the direct recombination of photoexcited charge carriers. In contrast, the ZCS-D/CN composite displays emission peaks at similar wavelengths but with significantly reduced intensity. This reduction indicates that the synthesis of the composite effectively decreases the recombination rate of photoexcited charge carriers, thereby promoting the separation and transfer of electrons and holes [56-58].

    Figure 3

    Figure 3.  (a) UV–vis DRS and (b) corresponding plots of the converted Kubelka–Munk function versus photon energy for the prepared samples. (c) The band structure of CN and ZCS-D. (d) The photocatalytic PHP performance graph. (e) The 4 h cycle graph with four cycles. (f) Comparison of H2O2 production by ZCS-D/CN under the different conditions over 1 h: H2O, EDTA (0.5 mmol/L), NBT (0.06 mmol/L), and AgNO3 (0.5 mmol/L). (g) Controlled experiments: air atmosphere, Ar atmosphere and O2 atmosphere. (h) O2 was detected by EPR trapping experiments under light irradiation. (i) In situ FTIR spectra of ZCS-D/CN in photocatalytic reaction.

    The H2O2 production rate was measured using the potassium oxalate method, as shown in Fig. 3d. The H2O2 production rates under visible light irradiation and O2 flow for CN, ZCS-D, and ZCS-D/CN were 24, 4012, and 5124 µmol g-1 h-1, respectively. Compared to ZnS-D/CN and CdS-D/CN, ZCS-D/CN exhibits superior PHP performance (Fig. S9 in Supporting information). The H2O2 yield from CN is negligible, which can be attributed to its narrow light absorption range and small specific surface area. In contrast, the ZCS-D/CN composite showed a significant improvement in performance, achieving a 1.25-fold increase in H2O2 production compared to pure ZCS-D under the same conditions. The formation of an S-scheme heterojunction between CN and ZCS-D optimizes the electron migration path, where the electrons and holes are efficiently separated and localized on different surfaces. This reduces the recombination rate of electron-hole pairs, enabling more photogenerated charge carriers to participate in the reaction, which accelerates the photocatalytic process. Notably, the performance of pure ZCS-D is also superior to most photocatalysts reported in the literature. To assess the stability of the prepared material, photocatalytic H2O2 precipitation cycling experiments were conducted under the same testing conditions, as shown in Fig. 3e. After four cycles of 4 h each, the H2O2 yield decreased only slightly, which is primarily due to catalyst loss during the recycling process. This indicates that ZCS-D/CN is a stable photocatalyst with good durability. In the context of current research, the photocatalytic H2O2 production process driven by semiconductors involves a series of complex and continuous reactions, including oxygen reduction reaction (ORR) and water oxidation reaction (WOR), as shown in Eqs. S4-S11 (Supporting information).

    To identify the key active species involved in the H2O2 production process, radical inhibition experiments were conducted using ZCS-D/CN as the catalyst, as shown in Fig. 3f. Compared to the pure water system, the H2O2 concentration in NBT solution decreased slightly, in the NBT system, a small amount of photogenerated electrons were captured for the reduction of NBT, which reduced the number of electrons available for the oxygen reduction reaction (O2 → H2O2). Consequently, the H2O2 production decreased slightly. This indicates that photogenerated electrons are essential for H2O2 generation, but the introduction of NBT did not completely suppress H2O2 production, suggesting that holes (h+) and other pathways may still contribute to H2O2 formation. In AgNO3 solution, the H2O2 yield was significantly reduced. During the photocatalytic process, part of the H2O2 production might involve the water oxidation reaction (H2O → H2O2) driven by photogenerated holes. When AgNO3 captures the holes, it inhibits this reaction pathway, thus significantly reducing the H2O2 production. This demonstrates that photogenerated holes play a crucial role in H2O2 generation, especially in the water oxidation (WOR) pathway. Additionally, after the introduction of EDTA, there was a slight decrease in H2O2 production. This occurred because EDTA preferentially captured the photogenerated holes, reducing the hole-driven oxidation reactions (such as the water oxidation pathway), while allowing more photogenerated electrons to participate in the oxygen reduction pathway. Although EDTA captures some holes, inhibiting hole-driven WOR, photogenerated electrons can still participate in the oxygen reduction reaction (ORR) and generate O2, which eventually lead to the formation of H2O2. Therefore, the importance of photogenerated electrons in H2O2 production was not completely suppressed. The slight decrease in H2O2 production indicates that while the introduction of EDTA reduced the oxidation reaction involving holes, the reduction reaction driven by electrons remained active. Through the analysis above, it is clear that the separation and utilization efficiency of photogenerated carriers are crucial for photocatalytic H2O2 production, and that oxygen reduction and water oxidation are two important reaction pathways. To evaluate the role of O2 in photocatalytic H2O2 production, different atmospheres were introduced into the reaction system, as shown in Fig. 3g. In an air system, the H2O2 yield was significantly lower compared to the continuous oxygen supply system. Under continuous argon supply, the H2O2 yield was nearly zero. Therefore, during the ZCS-D/CN photocatalysis process, due to the inability of the ZCS-D/CN VB to meet the potential for generating OH, thermodynamic conditions were not satisfied, and the two-step 1e WOR pathway did not occur. The primary process in photocatalysis involves photogenerated electrons initially participating in the oxygen reduction reaction (ORR), where oxygen molecules are reduced to O2. Under suitable conditions, these O2 radicals then combine with protons to form hydrogen peroxide (H₂O₂). There might still be superoxide radical reaction pathways (Eqs. S4-S6 in Supporting information) and a two-electron direct reduction pathway (Eq. S7 in Supporting information). To investigate these pathways, room-temperature EPR was used to detect related reaction intermediates. Under light illumination, using DMPO as a trapping agent (Fig. 3h), a distinct O2 signal was detected. DMPO is a key intermediate in the successive superoxide radical reduction and the O two-electron direct reaction, confirming the existence of these pathways. The PHP of ZCS-D/CN is a complex process involving multiple evolution steps. In-situ FTIR revealed that, within the initial 30 min in the dark, the adsorption and desorption of O2 reached an equilibrium, showing weak vibrational signals (Fig. 3i). As the illumination time increased, the signals gradually intensified. Positive peaks corresponding to different intermediate species were observed, with bands at 1099 and 1243 cm-1 attributed to adsorbed O2 and peroxide (OOH*), respectively. Additionally, a significant peak at 1438 cm-1 is associated with the O—O vibration of adsorbed molecular oxygen species [59-62].

    To understand the charge transfer mechanism from a theoretical perspective, density functional theory (DFT) calculations were performed to investigate the impact of electronic transfer at the heterojunction interface of the composite material. During ORR process, the photogenerated electrons on the surface of the photocatalyst are initially captured by O2. The work functions of the CN (001) and ZCS-D (111) surfaces were found to be 4.58 and 4.44 eV, respectively (Figs. 4a and b). Since the Fermi level (Ep) value is the opposite of the work function when the vacuum level is set to zero, the Ep of CN is more negative than that of ZCS-D. Therefore, when CN and ZCS-D form a heterojunction, electrons will transfer from ZCS-D to CN until the Ep values reach equilibrium. According to Bader charge analysis, the electron transfer from ZCS-D to CN is 0.97 e (Fig. 4c). Additionally, the planar average charge density difference (CDD) of the ZCS-D/CN heterojunction was calculated, with the charge density on CN being positive and on ZCS-D being negative, further indicating the direction of electron transfer. This electron transfer at the interface generates an internal electric field that facilitates the driving of photogenerated electron transfer. The charge density difference was simulated using DFT, with charge accumulation regions and depletion regions represented in yellow and cyan, respectively (Figs. 4d-f) [57,63]. After the contact between ZCS-D and CN, electrons spontaneously transfer from ZCS-D to CN, indicating that ZCS-D, which is electron-deficient, has a higher Fermi level and can serve as RP, while CN, being electron-rich, acts as OP. As illustrated in Fig. 4g, the charge transfer and separation mechanism of ZCS-D/CN S-scheme heterojunction is elucidated. The photogenerated electrons with strong reductive properties in the CB of ZCS-D are effectively separated and retained, participating in the photocatalytic reaction. Thus, ZCS-D/CN S-scheme heterojunction provides a key strategy for enhancing photocatalytic hydrogen peroxide (PHP) activity [64-68].

    Figure 4

    Figure 4.  (a) Electrostatic potentials of CN (001) and (b) ZCS-D (111) surface. (c) Bader charge analysis of CN absorbed with ZCS-D. (d-f) The charge accumulation is shown as the yellow region, and the charge depletion is shown as the cyan region. (g) Charge transfer diagram of S-scheme heterojunction.

    In summary, we successfully synthesized a ZCS-D/CN composite photocatalyst with an S-scheme heterojunction structure via a hydrothermal method. The results demonstrate that the composite exhibits excellent performance and remarkable stability in photocatalytic H2O2 production. These studies confirm that the heterojunction effectively promotes the separation and migration of photogenerated charge carriers, significantly enhancing H2O2 yield. This work provides valuable insights for the design of high-efficiency photocatalytic systems and is expected to contribute to the advancement of green energy utilization and environmental remediation technologies.

    We declare that we have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Zheng Liu: Writing – original draft, Investigation, Data curation, Conceptualization. Yuqing Bian: Investigation, Data curation. Graham Dawson: Writing – review & editing, Resources. Jiawei Zhu: Resources, Investigation, Formal analysis. Kai Dai: Writing – review & editing, Supervision, Resources, Project administration, Funding acquisition, Conceptualization.

    This work was supported by the National Natural Science Foundation of China (No. 22278169), the Excellent Scientific Research and Innovation Team of Education Department of Anhui Province (No. 2022AH010028) and Anhui Provincial Quality Engineering Project (No. 2022sx134).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111272.


    1. [1]

      Q. Xu, L. Zhang, B. Cheng, et al., Chem 6 (2020) 1543–1559. doi: 10.1016/j.chempr.2020.06.010

    2. [2]

      X. Xu, C. Shao, J. Zhang, et al., Acta Phys. Chim. Sin. 40 (2023) 2309031.

    3. [3]

      M. Sayed, F. Xu, P. Kuang, et al., Nat. Commun. 12 (2021) 4936. doi: 10.1038/s41467-021-25007-6

    4. [4]

      L. Liu, Z. Wang, J. Zhang, et al., Adv. Mater. 35 (2023) 2300643. doi: 10.1002/adma.202300643

    5. [5]

      C. Nie, X. Wang, P. Lu, et al., J. Mater. Sci. Technol. 169 (2024) 182–198. doi: 10.1016/j.jmst.2023.06.011

    6. [6]

      J. Hua, Z. Dai, K. Cheng, et al., Nano Lett. 24 (2024) 14363–14372. doi: 10.1021/acs.nanolett.4c03978

    7. [7]

      F.N. Habarugira, D. Yao, W. Miao, et al., Chin. Chem. Lett. 35 (2024) 109886. doi: 10.1016/j.cclet.2024.109886

    8. [8]

      S. Ding, J. Duan, S. Chen, EcoEnergy 2 (2024) 45–82. doi: 10.1002/ece2.26

    9. [9]

      B. Liu, J. Cai, J. Zhang, et al., Chin. J. Catal. 51 (2023) 204–215. doi: 10.1016/S1872-2067(23)64466-3

    10. [10]

      X. Wang, C. Chen, Q. Wang, et al., Chin. J. Struct. Chem. 43 (2024) 100473.

    11. [11]

      H. Lv, Z. Li, P. Yin, et al., Chin. Chem. Lett. 36 (2025) 110457. doi: 10.1016/j.cclet.2024.110457

    12. [12]

      K. Meng, J. Zhang, B. Cheng, et al., Adv. Mater. 36 (2024) 2406460. doi: 10.1002/adma.202406460

    13. [13]

      B. Zhu, J. Liu, J. Sun, et al., J. Mater. Sci. Technol. 162 (2023) 90–98. doi: 10.1016/j.jmst.2023.03.054

    14. [14]

      K. Huang, G. Liang, S. Sun, et al., J. Mater. Sci. Technol. 193 (2024) 98–106. doi: 10.1016/j.jmst.2024.01.034

    15. [15]

      B. He, Z. Wang, P. Xiao, et al., Adv. Mater. 34 (2022) 2203225. doi: 10.1002/adma.202203225

    16. [16]

      Z. Wang, Y. Gao, T. Wang, et al., Chin. Chem. Lett. 36 (2025) 110602. doi: 10.1016/j.cclet.2024.110602

    17. [17]

      Z. Jiang, Q. Long, B. Cheng, et al., J. Mater. Sci. Technol. 162 (2023) 1–10. doi: 10.1016/j.jmst.2023.03.045

    18. [18]

      M. Gao, Z. Sun, Y. Gong, et al., J. Liaocheng Univ. Nat. Sci. Ed. 37 (2024) 39–48.

    19. [19]

      Y. Cui, J. Zhang, H. Chu, et al., Acta Phys. Chim. Sin. 40 (2024) 2405016. doi: 10.3866/PKU.WHXB202405016

    20. [20]

      Q. Wang, Z. Miao, Y. Zhang, et al., ACS Catal. 12 (2022) 4016–4025. doi: 10.1021/acscatal.1c05553

    21. [21]

      E. Cui, Y. Lu, J. Jiang, et al., Chin. J. Catal. 59 (2024) 126–136. doi: 10.1016/S1872-2067(23)64630-3

    22. [22]

      P. Su, J. Yu, P. Deng, et al., J. Liaocheng Univ. Nat. Sci. Ed. 37 (2024) 123–131.

    23. [23]

      R. Gao, R. Shen, C. Huang, et al., Angew. Chem. Int. Ed. 64 (2024) e202414229.

    24. [24]

      C. Bie, Z. Meng, B. He, et al., J. Mater. Sci. Technol. 173 (2024) 11–19. doi: 10.1016/j.jmst.2023.07.019

    25. [25]

      S. Yan, Y. Li, X. Yang, et al., Adv. Mater. 36 (2023) 2307967.

    26. [26]

      H. Ding, R. Shen, K. Huang, et al., Adv. Funct. Mater. 34 (2024) 2400065.

    27. [27]

      H. Zhang, C. Shao, Z. Wang, et al., J. Mater. Sci. Technol. 195 (2024) 146–154.

    28. [28]

      T. Yang, J. Wang, Z. Wang, et al., Chin. J. Catal. 58 (2024) 157–167.

    29. [29]

      C. Yang, Q. Rong, F. Shi, et al., Chin. Chem. Lett. 35 (2024) 109767.

    30. [30]

      Y. Tang, W. Wang, J. Ran, et al., Energy Environ. Sci. 17 (2024) 6482–6498.

    31. [31]

      Y. Bian, H. He, G. Dawson, et al., Sci. China Mater. 67 (2024) 514–523.

    32. [32]

      Z. Zhou, H. Yao, Y. Wu, et al., Acta Phys. Chim. Sin. 40 (2024) 2312010.

    33. [33]

      C. Yang, X. Li, M. Li, et al., Chin. J. Catal. 56 (2024) 88–103.

    34. [34]

      C. Cheng, B. He, J. Fan, et al., Adv. Mater. 33 (2021) 2100317.

    35. [35]

      L. Xiao, W. Ren, S. Shen, et al., Acta Phys. Chim. Sin. 40 (2024) 2308036.

    36. [36]

      C. Cheng, J. Yu, D. Xu, et al., Nat. Commun. 15 (2024) 1313.

    37. [37]

      C. Yang, B. Cheng, J. Xu, et al., EnergyChem 6 (2024) 100116.

    38. [38]

      S. Cao, B. Zhong, C. Bie, et al., Acta Phys. Chim. Sin. 40 (2024) 2307016.

    39. [39]

      Z. Jiang, B. Cheng, L. Zhang, et al., Chin. J. Catal. 52 (2023) 32–49.

    40. [40]

      J. Cai, B. Liu, S. Zhang, et al., J. Mater. Sci. Technol. 197 (2024) 183–193.

    41. [41]

      Z. Miao, Q. Wang, Y. Zhang, et al., Appl. Catal. B: Environ. 301 (2022) 120802.

    42. [42]

      M. Cai, Y. Wei, Y. Li, et al., EcoEnergy 1 (2023) 248–295.

    43. [43]

      X. Deng, J. Zhang, K. Qi, et al., Nat. Commun. 15 (2024) 4807.

    44. [44]

      J. Wang, Z. Wang, J. Zhang, et al., ACS Nano 18 (2024) 20740–20750.

    45. [45]

      C. Wang, C. You, K. Rong, et al., Acta Phys. Chim. Sin. 40 (2024) 2307045.

    46. [46]

      Z. Zhao, Z. Wang, J. Zhang, et al., Adv. Funct. Mater. 33 (2023) 2214470.

    47. [47]

      D. Chen, Z. Wang, J. Fu, et al., Sci. China Mater. 67 (2024) 541–549.

    48. [48]

      X. Gao, D. Zeng, J. Yang, et al., Chin. J. Catal. 42 (2021) 1137–1146.

    49. [49]

      H. Jiang, W. Ye, H. Zhen, et al., Chin. Chem. Lett. 36 (2025) 109984.

    50. [50]

      R. Shen, G. Liang, L. Hao, et al., Adv. Mater. 35 (2023) 2303649.

    51. [51]

      Y. Wu, Y. Yang, M. Gu, et al., Chin. J. Catal. 53 (2023) 123–133.

    52. [52]

      R.I. Rodríguez, V. Corti, L. Rizzo, et al., Nat. Catal. 7 (2024) 1223–1231.

    53. [53]

      Y. Xing, X. Wang, S. Hao, et al., Chin. Chem. Lett. 32 (2021) 13–20.

    54. [54]

      J. Yan, J. Zhang, J. Mater. Sci. Technol. 193 (2024) 18–21.

    55. [55]

      H. Wang, L. Yu, J. Jiang, et al., Acta Phys. Chim. Sin. 40 (2024) 2305047.

    56. [56]

      J. Yu, X. Yao, P. Su, et al., J. Liaocheng Univ. Nat. Sci. Ed. 37 (2024) 52–61.

    57. [57]

      F. Li, G. Zhu, J. Jiang, et al., J. Mater. Sci. Technol. 177 (2024) 142–180.

    58. [58]

      Z. Pan, W. Ding, H. Chen, et al., Chin. Chem. Lett. 35 (2024) 108567.

    59. [59]

      X. Ma, S. Li, Y. Gao, et al., Adv. Funct. Mater. 34 (2024) 2409913.

    60. [60]

      K. Dong, C. Shen, R. Yan, et al., Acta Phys. Chim. Sin. 40 (2024) 2310013.

    61. [61]

      B. He, P. Xiao, S. Wan, et al., Angew. Chem. Int. Ed. 62 (2023) e202313172.

    62. [62]

      H. Guo, L. Zhou, K. Huang, et al., Adv. Funct. Mater. 34 (2024) 2402650.

    63. [63]

      H. He, Z. Wang, J. Zhang, et al., Adv. Funct. Mater. 34 (2024) 2315426.

    64. [64]

      B. Zhu, C. Jiang, J. Xu, et al., Mater. Today. 82 (2025) 251–273.

    65. [65]

      W. Yu, C. Bie, Acta Phys. Chim. Sin. 40 (2024) 2307022.

    66. [66]

      C. Chen, J. Zhang, H. Chu, et al., Chin. J. Catal. 63 (2024) 81–108.

    67. [67]

      J.Y. Yue, J.X. Luo, Z.X. Pan, et al., Angew. Chem. Int. Ed. 63 (2024) e202405763.

    68. [68]

      S. Wan, W. Wang, B. Cheng, et al., Nat. Commun. 15 (2024) 9612.

  • Figure 1  (a) Schematic diagram of the synthesis pathway for ZCS-D/CN. (b) XRD pattern of CN, ZCS-D/CN, and ZCS-D. (c) FTIR spectra of CN and the ZCS-D/CN. (d) TEM and (e) HRTEM images of ZCS-D/CN. (f, g) Elemental mapping of ZCS-D/CN.

    Figure 2  (a) The XPS spectra of pristine CN, ZCS-D/CN, and ZCS-D. The survey spectra: (b) Zn 2p, (c) Cd 3d, (d) S 3d, (e) C 1s, (f) N 1s.

    Figure 3  (a) UV–vis DRS and (b) corresponding plots of the converted Kubelka–Munk function versus photon energy for the prepared samples. (c) The band structure of CN and ZCS-D. (d) The photocatalytic PHP performance graph. (e) The 4 h cycle graph with four cycles. (f) Comparison of H2O2 production by ZCS-D/CN under the different conditions over 1 h: H2O, EDTA (0.5 mmol/L), NBT (0.06 mmol/L), and AgNO3 (0.5 mmol/L). (g) Controlled experiments: air atmosphere, Ar atmosphere and O2 atmosphere. (h) O2 was detected by EPR trapping experiments under light irradiation. (i) In situ FTIR spectra of ZCS-D/CN in photocatalytic reaction.

    Figure 4  (a) Electrostatic potentials of CN (001) and (b) ZCS-D (111) surface. (c) Bader charge analysis of CN absorbed with ZCS-D. (d-f) The charge accumulation is shown as the yellow region, and the charge depletion is shown as the cyan region. (g) Charge transfer diagram of S-scheme heterojunction.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  58
  • HTML全文浏览量:  8
文章相关
  • 发布日期:  2025-09-15
  • 收稿日期:  2025-01-27
  • 接受日期:  2025-04-27
  • 修回日期:  2025-04-25
  • 网络出版日期:  2025-04-28
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章