C–C coupling regulation to enhance the stability of ambient pressure photothermal CO2 hydrogenation to multi-hydrocarbon compounds

Xianhua Bai a, Linjia Han ab, Jialin Wang c, Yanhong Luo abd, Yiming Li a, Jiangjian Shi a, Yaguang Li *c, Dongmei Li abd and Qingbo Meng *ade
aBeijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing, 100190, P. R. China. E-mail: qbmeng@iphy.ac.cn
bSchool of Physical Sciences, University of Chinese Academy of Sciences, Beijing, 100049, P. R. China
cResearch Center for Solar Driven Carbon Neutrality, Engineering Research Center of Zero-Carbon Energy Buildings and Measurement Techniques, Ministry of Education, The College of Physics Science and Technology, Institute of Life Science and Green Development, Hebei University, Baoding, 071002, P. R. China. E-mail: liyaguang@hbu.edu.cn
dSongshan Lake Materials Laboratory, Dongguan, 523808, P. R. China
eCenter of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing, 100049, P. R. China

Received 5th May 2025 , Accepted 4th June 2025

First published on 5th June 2025


Abstract

Ambient pressure photothermal CO2 hydrogenation for producing multi-hydrocarbon (C2+: CxHy, where carbon number >1) compounds is a highly valuable way to recycle CO2 and an important path to achieve carbon neutrality. It suffers from carbon deposition during the C–C coupling process that results in low catalytic stability. To overcome this challenge, a Fe3C/ZnO heterostructure was designed to realize ambient pressure photothermal CO2 hydrogenation that can not only achieve a C2+ generation rate of ∼1.9 mmol g−1 h−1, 67.9% C2+ selectivity and a CO2 conversion rate of 29.8% under natural sunlight irradiation, but also extend the stable reaction duration from 40 hours to 200 hours. In situ DRIFTS and theoretical calculations demonstrate that the Fe3C/ZnO heterostructures could significantly reduce the adsorption of CHx intermediates and activate the HCO* intermediates to regulate the C–C formation pathway of photothermal CO2 hydrogenation from the traditional CHx intermediates to HCO* and CO* intermediates, thus mitigating surface carbon deposition. This study contributes to the advancement of new catalysts designed for outdoor photothermal CO2 hydrogenation aimed at robustly producing C2+ compounds under ambient pressure.


Introduction

Solar-driven photothermal CO2 hydrogenation for the production of C2+ compounds not only addresses the energy consumption issues associated with traditional CO2 hydrogenation reactions but also offers significant economic benefits and practical application potential.1 Specifically under ambient pressure conditions, achieving C2+ production from photothermal CO2 hydrogenation can be integrated with the hydrogen production system of photovoltaic water electrolysis, reducing storage and transportation costs of H2 and presenting a promising pathway for carbon cycling.2,3 However, CO2 hydrogenation to C2+ compounds via the CHx intermediate coupling pathway often results in carbon deposition under ambient pressure, which can cover active sites and lead to catalyst deactivation.4,5 Currently, the state-of-the-art C2+ production method from photothermal CO2 hydrogenation only has ∼60 hours of operation.6,7 Therefore, improving catalyst stability and extending catalyst lifespan remain as urgent challenges that need to be addressed for the photothermal CO2 hydrogenation to C2+ compounds under ambient pressure.

Herein, we employ a polyvinylpyrrolidone (PVP) soft template method to construct a Fe3C/ZnO heterostructure on a nitrogen-doped carbon support, which is designed to regulate the catalytic reaction pathway. During CO2 hydrogenation, C2+ production proceeds via the coupling of HCO* and CO* intermediates. In situ DRIFTS and theoretical calculations demonstrate that the Fe3C/ZnO heterostructures can reduce the catalyst's adsorption of CHx intermediates, effectively preventing surface carbon deposition. As a result, this heterostructure catalyst achieves photothermal CO2 hydrogenation to C2+ compounds under ambient pressure and 1.2 kW m−2 light irradiation, with a C2+ generation rate of ∼1.9 mmol g−1 h−1, C2+ selectivity of 67.9% and a CO2 conversion rate of 29.8%. Furthermore, it maintained 97% C2+ selectivity over a continuous 200 hour photothermal CO2 hydrogenation reaction, demonstrating excellent catalytic stability. These results make it an ideal material for outdoor, solar-driven ambient pressure photothermal CO2 hydrogenation to C2+.

Results and discussion

The heterostructures composed of Fe3C and ZnO (Fe3C/ZnO) were synthesized utilizing a polyvinylpyrrolidone (PVP) template method, and X-ray diffraction (XRD) analysis was employed to verify the phase composition of the obtained sample. The XRD pattern revealed distinct characteristic peaks of Fe3C/ZnO corresponding to ZnO and Fe3C (Fig. 1a) with a small peak at approximately 44.5° attributed to the (110) plane of metallic Fe (PDF#87-0722), and the weight ratio of Fe and Zn is about 1.3 (Table S1). To further analyze the electronic and magnetic properties of the iron species, 57Fe Mössbauer spectroscopy (MS) was employed. The MS spectra of the Fe3C/ZnO catalyst included one sextet and one doublet (Fig. 1b). The sextet is indicative of the presence of iron in the form of Fe3C, and the doublet corresponds to metallic iron (Table S2).8 X-ray photoelectron spectroscopy (XPS) was employed to investigate the electronic structures of the Fe3C/ZnO catalyst. As shown in Fig. 1c, the Fe 2p XPS spectrum of the Fe3C/ZnO catalyst displayed one satellite peak alongside three prominent peaks positioned at approximately 707.6, 710.8, and 713.3 eV, corresponding to Fe0, Fe2+, and Fe3+.9,10 Furthermore, the Zn 2p spectrum of Fe3C/ZnO confirm that the valence state of Zn is +2 (Fig. 1d), consistent with the XRD patterns. The C 1s spectrum in Fig. S1a displayed three distinct peaks around 284.8, 285.8, and 289.3 eV, confirming that carbon is present in the catalyst as C–C, C–O, and C[double bond, length as m-dash]O groups.11 Additionally, the deconvoluted N 1s XPS spectrum (Fig. S1b) exhibited distinct peaks around 398.5 eV and 400.8 eV, attributed to pyrrolic nitrogen and pyridinic nitrogen, indicating the existence of nitrogen-doped carbon.12,13
image file: d5cy00535c-f1.tif
Fig. 1 (a) XRD profiles; (b) Mössbauer spectroscopy; (c) Fe 2p XPS and (d) Zn 2p XPS spectra of the Fe3C/ZnO catalyst.

We used scanning electron microscopy (SEM) and transmission electron microscopy (TEM) to investigate the morphological structure of the synthesized sample. The Fe3C/ZnO catalyst exhibited a thin nanosheet structure, which resulted from the use of PVP as a template (Fig. 2a). The TEM image further confirmed the nanosheet structure and showed densely packed nanoparticles, approximately 7.1 nm in size, for Fe3C/ZnO (Fig. 2b). The high-resolution TEM image indicated that Fe nanoparticles in the Fe3C/ZnO sample are encapsulated within a graphite–carbon shell featuring a porous structure (Fig. S2), suggesting a core–shell architecture. This encapsulation contributes to the high specific surface area of Fe3C/ZnO (Fig. S3). A heterostructure can be seen in Fig. 2c, where the Fe3C (210) and ZnO (110) faces show lattice parameters of 0.254 and 0.209 nm, respectively. In Fig. 2d, energy-dispersive X-ray spectroscopy (EDX) analysis was utilized to reveal the spatial distributions of various elements present in the Fe3C/ZnO catalyst. The mapping images showed that the iron species, in the form of nanoparticles, are randomly distributed throughout the nanosheets, while the zinc species were closely associated with the Fe nanoparticles and well-dispersed throughout the structure (Fig. S4). These combined structural and electronic characterization results conclude that Fe3C/ZnO catalysts, composed of ZnO and Fe3C, are supported by nitrogen-doped carbon nanosheets, which collectively contribute to enhance CO2 hydrogenation catalytic activity and stability.


image file: d5cy00535c-f2.tif
Fig. 2 (a) SEM image; (b) TEM image (inset is particle diameter distribution map); (c) HRTEM image; (d) elemental mapping images of the Fe3C/ZnO catalyst.

To facilitate efficient photothermal catalysis, a homemade photothermal device utilizing a TiC/Cu-based absorber was employed to completely capture the solar spectrum and transform it into heat energy, which drives the CO2 hydrogenation reaction as depicted in Fig. 3a and S5.14–16 This device enabled the catalyst within to achieve temperatures of 321 °C solely with light illumination of 1.2 kW m−2 (Fig. 3b). The performance of the Fe3C/ZnO catalyst was evaluated under varying light intensities ranging from 0.8 to 1.4 kW m−2 in the photothermal device. Under ambient pressure, the conversion rate of CO2 increased with the light intensity (Fig. S6), whereas the selectivity for CO and C2+ products initially increased, but subsequently declined as light intensity increased. Notably, Fe3C/ZnO exhibited peak C2+ selectivity at a light intensity of 1.2 kW m−2, indicating that this intensity is optimal for photothermal catalysis. For comparison, we used the same PVP templated method to achieve the nitrogen-doped carbon nanosheet supported Fe species (abbreviated as Fe/NC, Fig. S7–S10). We also evaluated the performance of the Fe/NC catalyst at ambient pressure and 1.2 kW m−2 light illumination, and the CO2 conversion rate of the Fe/NC catalyst is only 25.1% with a C2+ selectivity of 54.1% (Fig. 3c and S11). Remarkably, the in situ introduction of Zn to the Fe catalyst both enhances the CO2 conversion of 29.9% and facilitates the C–C coupling, leading to a C2+ selectivity of 67.9%. This performance exceeds those of the state-of-the-art photo-driven CO2 hydrogenation catalysts (Table 1: line 2–11).6,17–25 Upon increasing the space velocity from 2400 to 7200 mL g−1 h−1, a slight decline of C2+ selectivity was observed, with a decrease from 67.9% to 60.6% (Fig. S12). Furthermore, the conversion of CO2 decreased from 29.8% to 24.9%. These findings underscore the robustness of Fe3C/ZnO across a diverse range of operation circumstances. The stability of catalysts is a vital factor that significantly influences effectiveness and longevity in practical applications, particularly in industrial processes. However, there exists a relative paucity of research dedicated to this aspect.19–21 Therefore, we undertook further investigations into the stability of Fe/NC and Fe3C/ZnO catalysts under 1.2 kW m−2 light illumination and at ambient pressure. The selectivity of C2+ for the Fe/NC catalyst sharply reduced from 54.1% to 38.3% within 40 hours, with the CO2 conversion rate reduced from 25.1% to 20.6% (Fig. 3d and S13a). In contrast, the Fe3C/ZnO catalyst still maintained 97% of the initial C2+ selectivity after 200 hours, with a CO2 conversion rate of 29.1% (Fig. S13b). The space–time yield (STY) of C2+ for the Fe/NC catalyst after a 40-hour test was 0.42 mmol g−1 h−1 (Fig. 3e). Meanwhile, the STY of C2+ for the Fe3C/ZnO catalyst slightly decreased from 1.94 to 1.84 mmol g−1 h−1 after a testing duration of 200 hours. These results demonstrate that the Fe3C/ZnO catalyst exhibits excellent catalytic stability for the ambient photothermal CO2 hydrogenation reaction.


image file: d5cy00535c-f3.tif
Fig. 3 (a) Photothermal catalysis CO2 hydrogenation scheme. (b) Temperatures of the catalyst within the TiC/Cu-based photothermal device under varying light intensities. (c) Photothermal CO2 hydrogenation performance of Fe/NC and Fe3C/ZnO catalysts. (d) C2+ selectivity of Fe/NC and Fe3C/ZnO catalysts during stability test. (e) C2+ STY of Fe/NC and Fe3C/ZnO catalysts during the stability test.
Table 1 Comparison of CO2 hydrogenation performance between Fe3C/ZnO and other reported catalysts under various reaction conditions
Catalysts Light intensity (W cm−2) Pressure (MPa) Duration (h) C2+ production rate (mmol g−1 h−1) Hydrocarbon distribution Ref.
CH4 C2+
Fe3C/ZnO 0.12 0 200 1.94 32.1 67.9 This work
5wt%Ag/TiO2 0.15 0 32 0.00038 71.8 28.2 17
NiFe3Ox 5 0.27 1 ∼1.17 81.1 18.9 18
CoFe/Al2O3 5.2 0.18 10 ∼0.42 62.9 37.1 19
Fe3C 2.05 0 12 0.89 90.9 9.1 20
Fe/FeOx/Al2O3–MgO 1.88 0.18 22 0.32 47.1 52.9 21
Co7Cu1Mn1Ox 0.234 0 54 1.4 91.2 8.8 6
Co–CoOx/MAO 1.2 0.3 50 1.16 37.5 62.5 22
CuOx@p-ZnO 0.1 0.5 32 0.00027 67.1 32.9 23
Graphdiyne/In2O3 2 0.16 20 ∼0.001 84.0 16.0 24
K–Ru/Fe3O4 2.05 0.1 5 0.63 34.9 62.5 25


To better understand the intermediates and remarkable stability of the Fe3C/ZnO catalyst during the process of CO2 hydrogenation, in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) measurements were performed. Fig. 4a illustrates several key C1 intermediates detected on Fe/NC and Fe3C/ZnO catalysts. These include surface carbonate species (image file: d5cy00535c-t1.tif), which exhibit characteristic peaks around 1270 cm−1. Bicarbonate species (image file: d5cy00535c-t2.tif) are identified by a peak around 1648 cm−1, while carboxylate species (COOH*) displayed a peak around 1560 cm−1. Additionally, CO* species are detected at approximately 2088 cm−1.26,27 However, the peak of CHx (about 1690 cm−1) related to carbon deposition is only observed in the Fe/NC catalyst.28,29 This indicates that the presence of ZnO in the Fe3C/ZnO catalyst reduces surface CHx species accumulation, thereby alleviating carbon deposition.


image file: d5cy00535c-f4.tif
Fig. 4 (a) CO2 hydrogenation DRIFTS spectra of Fe/NC and Fe3C/ZnO catalysts. (b) The structure model of Zn6O6/Fe3C(112). (c) The reaction energy profile for CO2 hydrogenation on Fe(110) and Zn6O6/Fe3C(112) surfaces.

To achieve a comprehensive understanding of mechanisms, density functional theory (DFT) was performed to calculate the reaction energies involved in hydrogenation process. The Fe(110) surface was used to simulate the Fe/NC catalyst (Fig. S14), while a Zn6O6 cluster on the Fe3C(112) surface (Fig. 4b) was used to represent the Fe3C/ZnO catalyst. The rate-determining step in the hydrogenation process involves coupling HCO* and CO* intermediates according to DRIFTS results (Fig. 4c). The reaction energy for this step is 3.30 eV on the Fe(110) surface, whereas on the Zn6O6/Fe3C(112) surface it is only 1.19 eV, suggesting that the introduction of zinc alters the catalyst's electronic structure and enhances its efficiency for CO2 hydrogenation. In addition, we also calculated the coupling energy of CH2 on Fe(110) surface, which was found to be 10.92 eV. This high energy suggests that Fe/NC is prone to CH2 retention, further contributing to carbon deposition. To confirm calculation results, the amount of carbon deposition on various catalysts was assessed through thermogravimetric analysis (TGA) in an Ar atmosphere. The weight loss of Fe3C/ZnO before and post testing was 51.3 wt% and 50.8 wt% (Fig. S15a and b), respectively. This slight difference indicates minimal carbon deposition, which contributes to the excellent stability of Fe3C/ZnO in photothermal CO2 hydrogenation. In contrast, the weight loss of Fe/NC before and after testing was 14.6 wt% and 18.2 wt%, respectively (Fig. S15c and d). This significant weight loss increase suggests considerable surface carbon deposition, leading to a notable decrease in CO2 hydrogenation activity and C2+ selectivity. The spent Fe3C/ZnO maintains its nanosheet morphology (Fig. S16), and the phase composition of the spent catalyst includes Fe3C and ZnO (Fig. S17). Moreover, EDS mapping of the spent catalyst reveals that Zn can migrate to the surface to prevent the oxidation of active sites by water (Fig. S18).30,31 These characterization results indicate that Fe3C/ZnO can retain both its structural and compositional stability during the CO2 hydrogenation process.

Conclusions

In this work, a Fe3C/ZnO catalyst was synthesized using a soft template method, which can realize ambient pressure photothermal CO2 hydrogenation that can achieve a C2+ generation rate of ∼1.9 mmol g−1 h−1, 67.9% C2+ selectivity and a CO2 conversion rate of 29.8% under natural sunlight irradiation. The heterostructure of Fe3C and ZnO is critical in modulating the C–C coupling pathway, facilitating a transition from the conventional CHx intermediate pathway to the HCO* and CO* intermediates during the CO2 hydrogenation. In situ DRIFTS and theoretical calculations demonstrate that Fe3C/ZnO heterostructures reduce the catalyst's adsorption of CHx intermediates and reduce carbon deposition, thereby enhancing the catalyst's stability over 200 hours of continuous operation. This study contributes to the advancement of new catalysts designed for photothermal CO2 hydrogenation aimed at producing C2+ compounds under ambient pressure and exposure to outdoor light conditions.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

X. Bai: investigation, formal analysis, and writing – original draft. L. Han: formal analysis and validation. J. Wang: investigation and data curation. Y. Luo: writing – original draft. Y. M. Li: investigation. J. Shi: Funding acquisition. Y. G. Li: methodology and writing – review & editing. D. Li: resource. Q. Meng: supervision, project administration, funding acquisition, and writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was funded by the National Natural Science Foundation of China (with Grant no. 52227803 (Q. M.) and 52222212 (J. S.). Additionally, J. S. also gratefully for the support received from the Youth Innovation Promotion Association of the Chinese Academy of Sciences.

References

  1. H. Lin, S. Luo, H. Zhang and J. Ye, Joule, 2022, 6, 294–314 CrossRef CAS .
  2. F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C. F. Elkjær, J. S. Hummelshøj, S. Dahl, I. Chorkendorff and J. K. Nørskov, Nat. Chem., 2014, 6, 320–324 CrossRef CAS PubMed .
  3. P. De Luna, C. Hahn, D. Higgins, S. A. Jaffer, T. F. Jaramillo and E. H. Sargent, Science, 2019, 364, eaav3506 CrossRef CAS PubMed .
  4. J. Xie, H. M. Torres Galvis, A. C. J. Koeken, A. Kirilin, A. I. Dugulan, M. Ruitenbeek and K. P. de Jong, ACS Catal., 2016, 6, 4017–4024 CrossRef CAS PubMed .
  5. E. T. C. Vogt, D. Fu and B. M. Weckhuysen, Angew. Chem., Int. Ed., 2023, 62, e202300319 CrossRef CAS PubMed .
  6. Z.-H. He, Z.-H. Li, Z.-Y. Wang, K. Wang, Y.-C. Sun, S.-W. Wang, W.-T. Wang, Y. Yang and Z.-T. Liu, Green Chem., 2021, 23, 5775–5785 RSC .
  7. S. Ning, J. Wang, X. Wu, L. Li, S. Zhang, S. Chen, X. Ren, L. Gao, Y. Hao, C. Lv, Y. Li and J. Ye, Adv. Funct. Mater., 2024, 34, 2400798 CrossRef CAS .
  8. Y. Shi, Z. Li, Q. Hao, R. Li, Y. Li, L. Guo, S. Ouyang, H. Yuan and T. Zhang, Adv. Funct. Mater., 2024, 34, 2308670 CrossRef CAS .
  9. T. Yamashita and P. Hayes, Appl. Surf. Sci., 2008, 254, 2441–2449 CrossRef CAS .
  10. A. Ramirez, L. Gevers, A. Bavykina, S. Ould-Chikh and J. Gascon, ACS Catal., 2018, 8, 9174–9182 CrossRef CAS .
  11. P. Peng, L. Shi, F. Huo, C. Mi, X. Wu, S. Zhang and Z. Xiang, Sci. Adv., 2019, 5, eaaw2322 CrossRef CAS PubMed .
  12. R. Kumar, M. Mooste, Z. Ahmed, S. Akula, I. Zekker, M. Marandi, M. Käärik, J. Leis, A. Kikas, A. Treshchalov, M. Otsus, J. Aruväli, V. Kisand, A. Tamm and K. Tammeveski, Ind. Chem. Mater., 2023, 1, 526–541 RSC .
  13. Y. Zeng, C. Li, B. Li, J. Liang, M. J. Zachman, D. A. Cullen, R. P. Hermann, E. E. Alp, B. Lavina, S. Karakalos, M. Lucero, B. Zhang, M. Wang, Z. Feng, G. Wang, J. Xie, D. J. Myers, J.-P. Dodelet and G. Wu, Nat. Catal., 2023, 6, 1215–1227 CrossRef CAS .
  14. Y. Li, X. Bai, D. Yuan, F. Zhang, B. Li, X. San, B. Liang, S. Wang, J. Luo and G. Fu, Nat. Commun., 2022, 13, 776 CrossRef CAS PubMed .
  15. Y. Li, X. Bai, D. Yuan, C. Yu, X. San, Y. Guo, L. Zhang and J. Ye, Nat. Commun., 2023, 14, 3171 CrossRef CAS PubMed .
  16. Y. Li, F. Meng, Q. Wu, D. Yuan, H. Wang, B. Liu, J. Wang, X. San, L. Gu and Q. Meng, Sci. Adv., 2024, 10, eadn5098 CrossRef CAS PubMed .
  17. M. Tahir and N. A. S. Amin, Int. J. Hydrogen Energy, 2017, 42, 15507–15522 CrossRef CAS .
  18. A. V. Puga and A. Corma, Top. Catal., 2018, 61, 1810–1819 CrossRef CAS .
  19. G. Chen, R. Gao, Y. Zhao, Z. Li, G. I. N. Waterhouse, R. Shi, J. Zhao, M. Zhang, L. Shang, G. Sheng, X. Zhang, X. Wen, L.-Z. Wu, C.-H. Tung and T. Zhang, Adv. Mater., 2018, 30, 1704663 CrossRef PubMed .
  20. C. Song, X. Liu, M. Xu, D. Masi, Y. Wang, Y. Deng, M. Zhang, X. Qin, K. Feng, J. Yan, J. Leng, Z. Wang, Y. Xu, B. Yan, S. Jin, D. Xu, Z. Yin, D. Xiao and D. Ma, ACS Catal., 2020, 10, 10364–10374 CrossRef CAS .
  21. Z. Li, J. Liu, R. Shi, G. I. N. Waterhouse, X.-D. Wen and T. Zhang, Adv. Energy Mater., 2021, 11, 2002783 CrossRef CAS .
  22. S. Ning, H. Ou, Y. Li, C. Lv, S. Wang, D. Wang and J. Ye, Angew. Chem., Int. Ed., 2023, 62, e202302253 CrossRef CAS PubMed .
  23. W. Wang, C. Deng, S. Xie, Y. Li, W. Zhang, H. Sheng, C. Chen and J. Zhao, J. Am. Chem. Soc., 2021, 143, 2984–2993 CrossRef CAS PubMed .
  24. W. Li, Y. Zhang, Y. Wang, W. Ran, Q. Guan, W. Yi, L. Zhang, D. Zhang, N. Li and T. Yan, Appl. Catal., B, 2024, 340, 123267 CrossRef CAS .
  25. C. Song, Z. Wang, J. Zhao, X. Qin, M. Peng, Z. Gao, M. Xu, Y. Xu, J. Yan, Y. Bi, M. Wang, L. Chen, Z. Yin, X. Liu, J. Liu and D. Ma, Chem Catal., 2024, 4, 100960 CrossRef CAS .
  26. M. K. Khan, P. Butolia, H. Jo, M. Irshad, D. Han, K.-W. Nam and J. Kim, ACS Catal., 2020, 10, 10325–10338 CrossRef CAS .
  27. N. Liu, J. Wei, J. Xu, Y. Yu, J. Yu, Y. Han, K. Wang, J. I. Orege, Q. Ge and J. Sun, Appl. Catal., B, 2023, 328, 122476 CrossRef CAS .
  28. B. C. Smith, Spectroscopy, 2016, 31, 28–34 Search PubMed .
  29. B. Wang, J. Li, H. Huang, B. Liang, Y. Zhang, L. Chen, K. Tan, Z. Chai, S. Wang, J. T. Wright, R. W. Meulenberg and S. Ma, ACS Cent. Sci., 2024, 10, 426–438 CrossRef CAS PubMed .
  30. C. Zhang, C. Cao, Y. Zhang, X. Liu, J. Xu, M. Zhu, W. Tu and Y.-F. Han, ACS Catal., 2021, 11, 2121–2133 CrossRef CAS .
  31. K. Wu, Z. Zhang, R. Shan, L. Li, J. Wang, B. Hou, Y. Xu and M. Ding, Appl. Catal., B, 2024, 353, 124067 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5cy00535c

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