Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Novel hierarchical MnxOy-functionalized graphene for enhanced visible-light photodegradation performance

Pham Huong Quynha, Ta Thi Huonga, Tran Y Doan Tranga, Pham Thi Mai Huonga, Nguyen Thi Thu Phuong*a, Pham Nguyet Anhb, Nguyen Thanh Trungcd, Vu Thi Thuyef, Dang Van Thanhfg and Nguyen Long Tuyen*h
aHanoi University of Industry, 298 Cau Dien Street, Bac Tu Liem District, Hanoi 100000, Vietnam. E-mail: phuongntt@haui.edu.vn
bThuyloi University, 175 Tay Son Street, Dong Da District, Hanoi 100000, Vietnam
cInstitute of Physics, Vietnam Academy of Science and Technology, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet, Hanoi 10072, Vietnam
dResearch and Development Center for Advanced Technology, Ha Noi, Vietnam
eInstitute of Science and Technology, TNU-University of Sciences, Tan Thinh Ward, Thai Nguyen 24000, Vietnam
fFaculty of Basic Sciences, TNU-University of Medicine and Pharmacy, 284 Luong Ngoc Quyen St., Thai Nguyen, Vietnam
gVNU Key Laboratory of Green Environment, Technology and Waste Utilization (GreenLab), University of Science, Vietnam National University, Hanoi, Vietnam
hHung Vuong University, Nong Trang, Viet Tri City, Phu Tho 35000, Vietnam. E-mail: nguyenlongtuyen@hvu.edu.vn

Received 6th May 2025 , Accepted 17th September 2025

First published on 30th September 2025


Abstract

Novel hierarchical MnxOy-functionalized graphene photocatalysts were successfully synthesized by simultaneous precipitation combined with electrochemical plasma methods. Comprehensive characterization, such as XRD, Raman, FT-IR, XPS, SEM, and TEM results, showed the formation of hierarchically structured MnxOy-graphene nanocomposites (GMP) via a change in the pH of the initial solution. The visible-light photodegradation performance of the as-prepared GMP was significantly enhanced compared to that of their counterparts owing to the synergistic effects of graphene and MnxOy. Specifically, the GMP10 sample displayed the highest efficiency for methylene blue decomposition, reaching 84.5% after 150 minutes of illumination. The stability of the material was confirmed through four consecutive degradation cycles, during which the efficiency of methylene blue degradation decreased by 6.9% (from 84.5% to 77.6%).


1 Introduction

Manganese oxides (MnxOy), including MnO, Mn3O4, Mn2O3, MnO2, and similar components, have been widely studied for their practical applications in adsorption,1 supercapacitors,2,3 photocatalysis4,5 due to the advantages of low cost, easy availability, low toxicity, large surface area, and high specific capacitance. However, their poor electronic conductivity and structural instability under various synthesis conditions hinder their applications.6 To overcome these obstacles, carbon materials, especially graphene, have been functionalized with MnxOy to take advantage of their synergistic effects in boosting electrocatalytic activity, catalyzing the decomposition of organic substances, and facilitating the charge storage process in supercapacitors.7,8 As a result, many methods have been developed to prepare hybrid materials, such as the sol–gel method,9 solvothermal method,10 hydrothermal method11 and microwave irradiation.12 These different approaches influence particle size distribution, particle morphology, nanoparticle dispersion, and the bonding between nanoparticles and the graphene layer, thus affecting the overall characteristics of the material. For example, Yingsi et al. used three methods to fabricate graphene-supported metal oxide nanomaterials: direct impregnation, homogeneous oxidative precipitation with hydrogen peroxide, and ammonia-catalyzed hydrolysis.13 The authors demonstrated that the ammonia-catalyzed hydrolysis method produced large and uneven particle sizes, whereas the oxidative precipitation approach resulted in uniform particle sizes and the best dispersion. The materials obtained by these three methods exhibited different catalytic activities depending on their particle size. The research also indicated that parameters such as temperature, reaction time, and pH significantly influenced the structure and properties of the final material. Recently, Atif et al. reported the green synthesis of MoO3 and MnO nanoparticles on graphene oxide, resulting in efficient photocatalysts. They found that GO/MnO demonstrated the highest removal efficiency of 98% for methylene blue (MB), significantly greater than that of pure MnO (60%).14 Jarvin et al. successfully prepared an Mn3O4-rGO nanocomposite by the solvothermal method, and the as-synthesized material achieved a good photocatalytic dye degradation efficiency of 60% within 120 min.15 Moreover, Mn2O3 anchored on graphitic carbon nitride was reported by Lalitha and colleagues.16 GCN-decorated Mn2O3 exhibited excellent photocatalytic performance for MB degradation.

Electrochemical discharge, a hybrid of electric discharge and electrolysis, is widely used for nanomaterial synthesis. Co-precipitation, a simple and cost-effective technique, enables the preparation of high-purity compounds with tailored structures due to its simplicity in device design, its low cost, and minimal equipment requirements for nanomaterial fabrication. Since plasma-liquid interactions involve both plasma and liquid, the properties of the final product can be tuned by adjusting the plasma parameters or the composition of the electrolyte medium, or by co-precipitation in the reaction solution. To the best of our knowledge, the use of electrochemical discharge and co-precipitation to effectively tune the Mn valence on graphene to provide an efficient photocatalyst has not yet been investigated.

Therefore, in this study, we hypothesize that Mn with multiple valence states can be effectively tuned during the co-precipitation reaction of Mn-containing salt precursors with the plasma-reacting electrolyte for synthesizing MnxOy/graphene by pH adjustments. Furthermore, these MnxOy/graphene compounds exhibited higher catalytic efficiency than single-phase MnxOy when tested using methylene blue (MB) dye as a simulated wastewater model.

2 Experiments

Manganese oxide was prepared using the conventional co-precipitation method. Solution (1) was a mixture of 200 mL of NaOH solution (10%) and 50 mL of (NH4)2SO4 solution (5%), with the pH adjusted to 12. Solution (2) was MnCl2 solution (2 M) dropped into mixture (1) at a rate of ∼1 drop/s. Ultrasonic vibration was then applied continuously during the dropwise addition and for an additional 30 minutes. The resulting material was denoted as MnxOy.

3 Results and discussion

Fig. 1(a) shows the XRD patterns of the MnxOy and graphene/MnxOy composite samples. The MnxOy sample was indexed to the tetragonal Mn3O4 structure (JCPDS No. 151-4120), and its diffraction peaks appeared in the composite samples, confirming the presence of Mn3O4. However, the composite samples also exhibited additional peaks indicating the presence of monoclinic MnOOH (JCPDS No. 900-9774) and MnO2 (JCPDS No. 901-6546). The sharp diffraction peaks indicated that the product was well-crystallized. From GMP8 to GMP12, the intensity ratio of the 26.1° peak (MnOOH phase) to the 36.1° peak ((211) facet of Mn3O4) gradually increases, implying decreased Mn3O4 crystallinity and increased MnOOH formation. Furthermore, the graphene peak at 26.01° overlaps with the 26.1° peak ((001) facet of MnOOH or (110) facet of MnO2), broadening the peak and confirming the presence of graphene. The XRD results show that the composite samples prepared by the plasma-assisted electrochemical exfoliation method contained additional components, such as MnOOH and MnO2, in addition to the Mn3O4 component. In contrast, the MnxOy samples prepared by the chemical precipitation method consisted solely of the Mn3O4 phase. The appearance of MnOOH and MnO2 in the composite samples was believed to result from the plasma discharge process and the change in the solution pH.
image file: d5ra03186a-f1.tif
Fig. 1 Schematic of MnxOy/graphene and photocatalytic degradation of MB.

The FT-IR results in Fig. 2(b) were used to investigate the bonding structures of the samples. For the graphene sample, the 1575 cm−1 absorption peak corresponds to the aromatic C[double bond, length as m-dash]C stretching vibration, which shifts to higher wavenumbers and broadens in the case of the composites. Additionally, the peaks at 497–533 cm−1 and 593–638 cm−1 in the spectra of the MnxOy and composite samples were attributed to the Mn–O bond vibrations and O–Mn–O stretching, respectively,17,18 confirming the coexistence of graphene and manganese oxide. Fig. 2(c and d) shows the Raman spectra of graphene, MnxOy, and their hybrid samples. Both graphene and the composites exhibit D bands (1348 cm−1) and G bands (1578 cm−1), corresponding to the breathing mode of the carbon rings and the stretching mode of the sp2 carbon atoms. However, the ID/IG intensity ratios for the GMP8 (0.78), GMP10 (0.85), and GMP12 (0.83) composite samples were higher than that of the graphene sample (0.52), indicating that MnxOy could be incorporated into the graphene structure, thereby increasing the degree of defects. Typical MnxOy bands were also observed; for example, GMP8 showed a peak at 642 cm−1, attributed to Mn–O stretching in the octahedral MnO6 structure of MnO2, while GMP12 exhibited a 654 cm−1 band indicative of Mn–O vibrations in the spinel Mn3O4 structure. The slight shift in this peak for the other composites possibly resulted from the mixing of the different components.


image file: d5ra03186a-f2.tif
Fig. 2 (a) XRD patterns; (b) FT-IR spectra and (c) Raman spectra of graphene, MnxOy, and GMP8–GMP10 composite materials; and (d) extended Raman spectra from part (c).

The XPS spectra were used to determine the oxidation state and composition of the MnxOy material and GMP8–GMP12 composites. As presented in Fig. 3(a), the MnxOy sample showed two O 1s peaks at 530.2 eV and 531.7 eV, corresponding to the lattice oxygen of Mn3O4 and surface-adsorbed oxygen, respectively, confirming its pure Mn3O4 phase, as observed using XRD. In contrast, the GMP8–GMP12 composites exhibited four O 1 s peaks at 529.6, 530.2, 531.3–531.5, and 532.9–533.5 eV, which correspond to O–Mn4+ bonds, Mn3O4 lattice oxygen, surface hydroxyl groups (Mn–OH/C–OH), and adsorbed water, respectively.19,20 These findings indicate that the composites contained both Mn4+ and Mn3+ (Mn3O4) components. Fig. 3(b) shows the Mn 2p spectra of the MnxOy sample, featuring Mn 2p3/2 and Mn 2p1/2 peaks at 641.5 eV and 653.3 eV, respectively, whereas in the composites, these peaks were shifted to 642.2–642.3 eV and 653.9–654 eV due to the formation of Mn4+ (MnO2) bonds. The energy separation of 11.8 eV for MnxOy and 11.7 eV for the composites is consistent with previous Mn3O4 data.21,22 Fig. 3(c) presents the high-resolution C 1s spectra of the GMP8–GMP12 composites deconvoluted into three peaks at 284.5 eV (C[double bond, length as m-dash]C/C–C), 285.7–286.2 eV (epoxy C–O), and 287.9–288.5 eV (carbonyl C[double bond, length as m-dash]O).23 The low intensity of the C–O and C[double bond, length as m-dash]O peaks indicates that graphene is in a low oxidation state.


image file: d5ra03186a-f3.tif
Fig. 3 XPS spectra of MnxOy and GMP8–GMP12 nanocomposite samples: (a) O 1s; (b) Mn 2p and (c) C 1s.

The SEM and TEM images in Fig. 4(a and c) clearly show agglomeration of the MnxOy nanoparticles. In contrast, GMP10 contains a large number of MnxOy nanoparticles (20–80 nm) anchored on the graphene sheets (Fig. 4(b)). The TEM images of GMP10 also verify the numerous folds at the edges of the graphene sheets (Fig. 4(d)), revealing a generally homogeneous shape with some pores. The combination of the graphene sheets with the MnxOy nanoparticles reduced material agglomeration while increasing the porosity of the composite, thereby enhancing its photocatalytic activity. The interplanar spacings of 0.264 nm and 0.288 nm were well observed, corresponding to the (202) plane of HMnO2 and the (200) plane of Mn3O4, respectively (Fig. 4(e)). Therefore, the observed interplanar spacings for the GMP10 sample further confirm the polycrystalline nature of the material.


image file: d5ra03186a-f4.tif
Fig. 4 SEM (a and b) and TEM (c and d) images of manganese oxide and the GMP10 composite material; and (e) HR-TEM images of the GMP10 composite material.

Fig. 5(a and b) shows the diffuse reflectance absorption spectra (DRS) for graphene, MnxOy, and the composite samples at various pH values. Both the MnxOy and composite samples exhibited strong UV absorption, but the MnxOy sample showed the least visible-light absorption. In contrast, the composite samples displayed enhanced visible-light absorption, increasing from GMP12 to GMP8 due to the increasing MnO2 content and the presence of graphene, which broadened the absorption band (Fig. 5(a)). Using the Kubelka–Munk model and the Tauc equation, the bandgaps for MnxOy, GMP12, GMP10, and GMP8 were determined to be 1.72, 1.56, 1.42, and 1.3 eV, respectively. These results suggest that graphene and the MnO2 components reduced the bandgap, thereby enhancing the electron–hole separation under visible light.


image file: d5ra03186a-f5.tif
Fig. 5 (a) Diffuse reflectance absorption spectrum; (b) Tauc plots; (c) N2 adsorption–desorption isotherms and (d) Nyquist plot of graphene, MnxOy, and GMP8–GMP12 composite samples.

The BET-specific surface of all the samples was measured, and the results are presented in Fig. 5(c). The isotherms of all samples exhibit a hysteresis loop and are classified as type IV isotherms, indicating a porous material. Specifically, the BET-specific surface areas of GMP8, GMP10, GMP12 and MnxOy were found to be 16.26, 25.16, 20.98 and 14.87 m2 g−1, respectively. The enhanced surface area of the composite materials compared to that of bare MnxOy is due to the uniform distribution of MnxOy nanoparticles on the graphene supports, providing more active sites for photocatalytic activity.

The electrochemical impedance spectroscopy (EIS) results (Fig. 5(d)) show that the composite samples had a smaller Nyquist radius than the MnxOy sample, indicating enhanced charge transfer, likely due to the highly conductive graphene component. Photocurrent measurements for the MnxOy, GMP8-GMP12 samples were also performed (SI, Fig. S1). In 0.1 M Na2SO4 solution, the GMP8–GMP12 heterojunction samples showed substantially higher photocurrents than the fabricated MnxOy sample. The GMP10 sample produced the largest photocurrent, approximately three times that of MnxOy. The photocurrent remained stable over many on/off light cycles, indicating efficient charge separation and suppressed recombination. Together with the EIS data, these observations demonstrate improved charge separation and lower charge-transfer impedance than in the composite samples. Overall, the formation of heterojunctions in the composites reduces electron–hole recombination and enhances carrier separation, which explains the improved photocatalytic performance.

The photocatalytic activity of the manganese oxide/graphene composites was evaluated by monitoring the decolorization of MB dye in water using a 400 W Xenon lamp. As shown in Fig. 6(a and b), the removal efficiency of MB by the GMP8–GMP12 composite samples (with 20 mg of catalyst) was superior to that of the MnxOy sample. In particular, the GMP10 sample exhibited the highest decomposition efficiency for MB, reaching 65.4% after 150 minutes of illumination (not including dark absorption). The GMP8 and GMP12 samples achieved degradation efficiencies of 40.8% and 53.1%, respectively, while the MnxOy sample reached only 30.1%. These results indicate that the incorporation of graphene enhanced the photocatalytic performance of the composite samples, in which graphene not only served as a substrate for the MnxOy nanoparticles, but also facilitated charge transfer. Using the Langmuir–Hinshelwood model (ln(Ct/C0) = −kt), the reaction rate constants (k) for the GMP12, GMP10, and GMP8 samples were determined to be 0.00495 min−1, 0.00716 min−1, and 0.00367 min−1, respectively. Table S1 compares the photocatalytic performance of the GMP10 composite sample with that reported in recent studies.


image file: d5ra03186a-f6.tif
Fig. 6 Comparison of photocatalytic ability of MnxOy and GMP8–GMP12 composite catalysts (catalyst dosage: m = 20 mg) for degrading MB dye: (a) C/C0 plot and (b) ln(C0/C) plot; (c) effects of the amount of GMP10 for MB decomposition; (d) photocatalytic cycling experiments of GMP10, using 80 mg of catalyst; (e) XRD patterns of GMP10 before and after stability test and (f) SEM of GMP10 after stability test.

Experiments varying the amount of photocatalyst were performed to determine the optimal catalyst loading; the results are shown in Fig. 6(c). The results suggest that the photocatalytic degradation efficiency increased with catalyst loading from 20 mg to 120 mg, reaching a maximum degradation efficiency of 83.3%. However, when the catalyst amount was increased to 120 mg, the degradation efficiency decreased slightly to 81.2%. These results can be explained by the fact that the photocatalytic degradation efficiency increased as the amount of catalyst increased since there were more reaction sites. However, excessive catalyst concentrations can hinder light penetration and restrict the interaction between light and the catalyst's surface, reducing the photocatalytic efficiency.

Additionally, photocatalytic cycling experiments were performed to evaluate the reusability of the material. As shown in Fig. 6(d), the degradation efficiencies of MB with 80 mg of the GMP10 catalyst were 83.3%, 83.7%, 80.7%, and 78.4% in four consecutive cycles. These results indicate that the photocatalytic performance of GMP10 remained relatively stable over four cycles. Furthermore, we also performed post-characterizations of the photocatalyst after investigating its stability (Fig. 6e and f). The XRD patterns confirm that the crystallinity of the material was well preserved, while the post-stability SEM images show a negligible change.

The photocatalytic mechanism is illustrated in Fig. 7. The energy levels, EC and EV, are referenced from Zhao et al.24 When MnxOy is irradiated with light of energy greater than its bandgap, electron–hole pairs are generated on the surface of the MnxOy particles. The photogenerated charges on the MnxOy particles can participate in a series of redox reactions to form radicals on the catalyst's surface. These radicals can subsequently interact with the MB dye molecules adsorbed on the catalyst's surface, resulting in the formation of degradation products. During this process, the photogenerated charged particles on the MnxOy particles can move to another location through the graphene, or to another particle (another oxide) due to the heterostructure, thus preventing recombination, which increases the photocatalytic degradation efficiency. Additionally, heterojunctions formed between the different MnxOy phases (such as Mn3O4, MnO2, and HMnO2) may also contribute to the improved photocatalytic activity, as the photogenerated charge carriers can be rapidly transferred between these phases. Some of the representative reactions involved in the photocatalytic degradation process of MnxOy-based materials are presented as follows:

 
MnxOy + → MnxOy (e + h+) (1)
 
e + O2 → O2˙ (2)
 
h+ + H2O → H+ + OH˙ (3)
 
OH˙ + OH˙ → H2O2 (4)
 
H2O2 + O2˙ [thin space (1/6-em)]→ OH˙·+ OH + O2 (5)
 
(h+, OH˙, O2˙) + MB → decomposition products (6)


image file: d5ra03186a-f7.tif
Fig. 7 Mechanism of MB photocatalysis by GMP10.

The quenching experiments were also performed with the free radical scavengers: EDTA (photohole quencher, h+), isopropyl alcohol (IPA) (hydroxyl radical scavenger, ˙OH) and ascorbic acid (AA) (superoxide radical scavenger, ˙O2). For each photocatalytic experiment, 8 mL of the scavenger solution (0.1 M) was added to suppress the main active species. The photocatalytic efficiency with AA, IPA, and EDTA as quenchers was 58.8%, 30.5% and 72.1%, respectively. The degradation efficiency for IPA as a quencher decreased most significantly, indicating that ˙OH played a dominant role in the degradation process. The contribution of ˙O2 also played a significant role, as shown by the reduced efficiency of 58.8%. Additionally, h + participated in the degradation process, but its contribution was not much (Fig. 8).


image file: d5ra03186a-f8.tif
Fig. 8 Quenching experiments with EDTA, IPA, and AA quenchers.

4 Conclusions

In this study, manganese oxide/graphene composites were successfully fabricated using plasma-assisted electrochemical exfoliation combined with co-precipitation. The graphene component enhanced conductivity, improving the decomposition efficiency for MB. Adjusting the pH of the precursor solution (NaOH 10% + (NH4)2SO4 5%) influenced the manganese oxide composition (MnO2, Mn3O4, MnOOH), enabling control over the material structures for specific applications. Photocatalytic evaluation under visible light revealed varying efficiencies due to material absorption, the graphene component, and heterojunctions within the composites. These findings offer a simple approach for developing graphene-based composites with diverse manganese oxide forms for the degradation of organic pollutants and other applications.

Author contributions

Pham Huong Quynh: processed data and prepared the manuscript. Nguyen Thi Thu Phuong, Ta Thi Huong, Tran Y Doan Trang, Pham Thi Mai Huong, Vu Thi Thuy conducted experiments and processed data. Dang Van Thanh, Pham Nguyet Anh, Nguyen Thanh Trung helped with material characterization and provided critical comments. Dang Van Thanh, Vu Thi Thuy, Nguyen Long Tuyen conceived, designed, and supervised the experiments. All authors read and approved the final manuscript.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

The supporting data has been submitted along with the manuscript and will be available as SI when the article online. See DOI: https://doi.org/10.1039/d5ra03186a.

Acknowledgements

Authors would like to express their sincere gratitude to Hanoi University of Industry for sponsoring the project with grant number HĐ 66-2024-RD/HĐ-ĐHCN.

References

  1. D. Zhou, Q. Liu, Q. Cheng, Y. Zhao, Y. Cui, T. Wang and B. Han, Chin. Sci. Bull., 2012, 57, 3059–3064 CrossRef CAS.
  2. S. Jadhav, R. S. Kalubarme, C. Terashima, B. B. Kale, V. Godbole, A. Fujishima and S. W. Gosavi, Electrochim. Acta, 2019, 299, 34–44 CrossRef CAS.
  3. A. Pramitha, S. S. Hegde, B. R. Bhat, C. Yadav, S. Chakraborty, A. Ravikumar, S. D. George, Y. N. Sudhakar and Y. Raviprakash, Phys. Scr., 2024, 99, 105922 CrossRef.
  4. S.-L. Chiam, S.-Y. Pung and F.-Y. Yeoh, Environ. Sci. Pollut. Res., 2020, 27, 5759–5778 CrossRef CAS.
  5. A. A. Amer, S. M. Reda, M. A. Mousa and M. M. Mohamed, RSC Adv., 2017, 7, 826–839 RSC.
  6. Y. Kong, R. Jiao, S. Zeng, C. Cui, H. Li, S. Xu and L. Wang, Nanomaterials, 2020, 10, 367 CrossRef CAS PubMed.
  7. R. Rajendiran, A. Patchaiyappan, S. Harisingh, P. Balla, A. Paari, B. Ponnala, V. Perupogu, U. Lassi and P. K. Seelam, J. Water Process Eng., 2022, 47, 102704 CrossRef.
  8. Y. Gu, J. Wu, X. Wang, W. Liu and S. Yan, ACS Omega, 2020, 5, 18975–18986 CrossRef CAS PubMed.
  9. M. Azarang, A. Shuhaimi, R. Yousefi and M. Sookhakian, J. Appl. Phys., 2014, 116, 084307 CrossRef.
  10. Z. Alves, C. Nunes and P. Ferreira, Nanomaterials, 2021, 11(8), 2149 CrossRef CAS.
  11. B. Saravanakumar, R. Mohan and S.-J. Kim, Mater. Res. Bull., 2013, 48, 878–883 CrossRef CAS.
  12. R. Singhmar, S. Sahoo, S. Choi, J. H. Choi, A. Sood and S. S. Han, Eur. Polym. J., 2024, 220, 113462 CrossRef CAS.
  13. Y. Wu, H. Yu, H. Wang and F. Peng, Chin. J. Catal., 2014, 35, 952–959 CrossRef CAS.
  14. A. Hussain, A. Zahid, S. Ali, N. Khalid, A. Ashraf, A. Latif, M. A. Farrukh and M. Jabeen, J. Taiwan Inst. Chem. Eng., 2024, 163, 105642 CrossRef CAS.
  15. M. Jarvin, S. A. Kumar, G. Vinodhkumar, E. Manikandan and S. S. R. Inbanathan, Mater. Lett., 2021, 305, 130750 CrossRef CAS.
  16. L. Kamarasu, S. Sree Nannapaneni, S. Arunachalam, P. Arumugam and N. Kumar Katari, Inorg. Chem. Commun., 2022, 145, 109949 CrossRef CAS.
  17. H. Visser, C. E. Dubé, W. H. Armstrong, K. Sauer and V. K. Yachandra, J. Am. Chem. Soc., 2002, 124, 11008–11017 CrossRef CAS.
  18. Y. Zheng, W. Pann, D. Zhengn and C. Sun, J. Electrochem. Soc., 2016, 163, D230–D238 CrossRef CAS.
  19. J. Zhao, J. Nan, Z. Zhao, N. Li, J. Liu and F. Cui, Appl. Catal., B, 2017, 202, 509–517 CrossRef CAS.
  20. P. Wu, S. Dai, G. Chen, S. Zhao, Z. Xu, M. Fu, P. Chen, Q. Chen, X. Jin and Y. Qiu, Appl. Catal., B, 2020, 268, 118418 CrossRef CAS.
  21. R. Yang, T. Tao, Y. Dai, Z. Chen, X. Zhang and Q. Song, Catal. Commun., 2015, 60, 96–99 CrossRef CAS.
  22. F. Zeng, Y. Pan, Y. Yang, Q. Li, G. Li, Z. Hou and G. Gu, Electrochim. Acta, 2016, 196, 587–596 CrossRef CAS.
  23. X. Chen, X. Wang and D. Fang, Fullerenes, Nanotubes Carbon Nanostruct., 2020, 28, 1048–1058 CrossRef CAS.
  24. J. Zhao, Z. Zhao, N. Li, J. Nan, R. Yu and J. Du, Chem. Eng. J., 2018, 353, 805–813 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.