Excessive carbon dioxide (CO2) emissions from fossil fuel combustion are a primary factor in the disruption of the natural carbon cycle. Electrochemical conversion of CO2 into high-value products (CO, CH4, C2H4, HCOOH, C2H5OH, etc.) offers a promising pathway to reverse entropy and recycle carbon [1-4]. However, the electrochemical CO2 reduction reaction (ECRR) is hindered by the kinetically favorable hydrogen evolution reaction (HER) and the challenge of achieving a highly selective reduction of CO2 to specific products with similar thermodynamic reduction potentials [5, 6]. Among the ECRR products, methane (CH4) is the main component of petroleum with the highest heat of combustion of 56 kJ/g [7], considered a promising energy to solve the global energy crisis. Various electro-reduction catalysts have been extensively investigated to achieve a highly selective ECRR to CH4 [8-12]. For example, a carbon-encapsulated copper-doped cerium oxide composite (Cu/CeO2@C) achieves a CH4 Faraday efficiency of 80.3% due to the synergistic effect of Cu/CeO2 and in situ-formed carbon with robust channels to assist charge transfer [13]. However, the complicated multiple proton and electron transfer steps, high overpotential, and poor selectivity of intermediate generation still block efficient CO2 utilization by electrochemical reduction [14-16].
Among studied catalysts, copper is the only metal that simultaneously possesses positive hydrogen and negative oxygen adsorption energy, allowing the formation of the following deep intermediates [17, 18]. The anchoring of Cu NPs on ZIF-8 (Cu/ZIF-8) achieves a maximum C2+ FE of 61.0% at a current density of 400 mA/cm2 [19]. The single-atom Ir/CuFeS2 QDs catalyst exhibited a suitable redox potential and active catalytic components, tailoring the chemical microenvironment for precisely engineered single-atom active sites [20]. Ag2Cu2O3 nanowires with Cu···Ag Lewis acid-base dual sites formed the highly stable Ag···C = O···Cu adsorption of the critical intermediate for CO2 to CH4 [21]. As demonstrated experimentally, the high-valence Cu2+ species would facilitate the hydrogenation and formation of hydrocarbon intermediates, showing more preference for the selectivity of CH4 [22]. For example, recent reports have shown that Cu2+ species in the solid solution (e.g., Cu-Ce-Ox) maintain high valence to benefit the adsorption of the *CO intermediate, further the hydrogenation for high CH4 faradaic efficiency [22]. In addition, a novel layered coordination polymer (CuPEDOT) keeps the Cu2+-EDOT coordination stable and promotes the hydrogenation of *CO intermediates, facilitating CH4 formation and prohibiting the dimerization of *CO to C2+ products [23]. The proof-of-concept catalysts of SrTi1-xCuxO3 also unraveled that the Cu2+ can promote activation/adsorption of reaction intermediates to produce CH4 [24]. However, Cu2+ species are thermodynamically unstable with cathodic potential applied and inevitably reduced to Cu0 [22, 23, 25]. The strategy to stabilize Cu2+ in its high valence is speculated to facilitate the generation of CH4 in ECRR.
Perovskite oxides (ABO3-δ) possess a flexible electronic structure and adjustable elemental components, which enable them to meet the meticulous requirements under different catalytic environments, including HER, oxygen evolution reaction (OER), nitrogen electroreduction reaction (NRR) [26-29]. By incorporating or substituting different metals, the electronic structure and charge density distributions of catalysts can be altered to formulate a suitable allotropic environment [26, 30-32]. Sun et al. designed Cu-MOF and CuBi double perovskite catalysts, CuBi@Cu-MOF-15 achieved a FEHCOOH of 93% at −1.1 V and a current density of 91.0 mA/cm2, for CuBi@Cu-MOF-15 is more favorable to stabilize *HCOO [33], while other CuBi@Cu-MOF achieved 56% of FEHCOOH [34]. In the ECRR catalyst studies with CH4 as the product, the combination of Cu with oxyphilic metal La can lower the energy barrier from *CO to *CHxO, facilitating the C—O cleavage for CH4 formation [35]. The perovskite catalyst La2CuO4 has been applied to the electrochemical methanation of CO2 via the structural evolution that generates the Cu/La2CuO4 heterostructure [36]. Furthermore, by introducing different A-site metals into perovskite catalysts, Ca2CuO3 and Sr2CuO3 showed different performance in terms of either selectivity or electrochemical activity, with uncoordinated copper sites on Ca2CuO3 contributing to the hydrogenation of *CO and *CHO intermediates to *CH2O [37]. There has also been a deepening of the realization that the A-site metal affects the electroreduction process according to its physicochemical property rather than a single supporting function, as previously thought [29, 37].
Here, we propose that the A-site components exert certain effects on the electronic structure of perovskites, which in turn affect the stability of the B-site metal, leading to distinctive intermediate preference and electrochemical properties. We introduced alkaline earth metal into the copper-based perovskite to test this conjecture, synthesizing LaCa0.4CuO3-δ for ECRR. After the introduction of Ca, an electrochemical test and in situ spectroscopic analysis revealed that the Cu2+ species in LaCa0.4CuO3-δ remain stable during ECRR, while that of La2CuO4 and commercial CuO are reduced to Cu0. The LaCa0.4CuO3-δ with stable Cu2+ species exhibit a FECH of 59.6% ± 3.8% at −1.3 V vs. reversible hydrogen electrode (RHE) in H-cell, and 38.7% ± 5.7% FECH in membrane electrode assembly (MEA) configuration with a partial current density of 155.0 mA/cm2. In situ Raman and density functional theory (DFT) calculations unraveled that the hydrogenation of *CH2O to *CH3O was promoted on LaCa0.4CuO3-δ, promoting the formation of CH4.
The LaCa0.4CuO3-δ perovskite catalyst was synthesized by a combined sol-gel and calcination method (Fig. 1). The quantitative elements component of LaCa0.4CuO3-δ perovskite catalyst was measured by inductively coupled plasma optical emission spectrometry (ICP-OES), which indicated that the atomic ratio of Ca/La/Cu was 0.4:1:1 (Table S1 in Supporting information), and thus the obtained electrocatalysts were named to LaCa0.4CuO3-δ.
|
Download:
|
| Fig. 1. Schematic of the synthesis route of LaCa0.4CuO3-δ by sol-gel method. | |
X-ray diffraction (XRD, Fig. 2a) analysis showed that the synthesized samples possessed orthorhombic phases with Cmca space group for both LaCa0.4CuO3-δ (PDF #82–1444) and La2CuO4 (PDF #89–8845). Compared to La2CuO4, the main peak corresponding to (131) plane of LaCa0.4CuO3-δ was slightly shifted to a higher angle (Fig. 2b), demonstrating the successful incorporation of Ca through doping (Fig. 2c).
|
Download:
|
| Fig. 2. (a) XRD patterns of commercial CuO, La2CuO4, and LaCa0.4CuO3-δ. (b) Zoom XRD patterns of La2CuO4 and LaCa0.4CuO3-δ in the 2-thera range of 30°−32.5°. (c) Schematic illustration of the structural transformation of La2CuO4 and LaCa0.4CuO3-δ. (d) TEM image and (e) HR-TEM image of LaCa0.4CuO3-δ. (f) EDS Elemental mapping images of LaCa0.4CuO3-δ. | |
Scanning electron microscopy (SEM) images reveal that commercial CuO, La2CuO4, and LaCa0.4CuO3-δ show a similar nanoparticle morphology (Figs. S1a-f in Supporting information). Transmission electron microscopy (TEM) image shows that LaCa0.4CuO3-δ gives a particle size of ca. 100 nm (Fig. 2d). The lattice spacings of 0.286 and 0.267 nm in the high-resolution TEM (HR-TEM) images corresponding to the (131) and (200) planes of LaCa0.4CuO3-δ (Fig. 2e and Fig. S2 in Supporting information). Energy-dispersive spectra (EDS) element mapping confirmed the uniform distribution of La, Cu, Ca, and O in LaCa0.4CuO3-δ (Fig. 2f), further confirming the successful introduction of Ca into the perovskite lattice.
The ECRR catalytic performance was firstly carried out in H-cell (Fig. S3 in Supporting information) with 0.1 mol/L KHCO3 aqueous as electrolyte under ambient conditions using a three-electrode system. All potentials shown in the H-cell have been converted to RHE with iR correction unless otherwise stated. The linear sweep voltammetry (LSV) curves of LaCa0.4CuO3-δ, La2CuO4, and commercial CuO are similar, especially at large current densities (Fig. S4 in Supporting information). The electrochemical surface areas (ECSA) of the catalysts were investigated using double-layer capacitance measurements, and these three catalysts also showed no significant difference in ECSA (Figs. S5a-d in Supporting information).
However, for the ECRR performance, with the modulation and elemental introduction of A-site, the CH4 selectivity showed an obvious increase. LaCa0.4CuO3-δ exhibited a FECH of 59.6% ± 3.8% at −1.3 V (vs. RHE) (Fig. 3a and Fig. S6 in Supporting information), while La2CuO4 yielded a ca. 18.3% FECH and commercial CuO had less than 10% (Fig. 3b and Fig. S7 in Supporting information). In addition, the commercial CuO favored C2+ products, and La2CuO4 had H2 as the main product, whereas LaCa0.4CuO3-δ showed a high preference for CH4 instead of these two types of co-products (Fig. 3c and Fig. S8 in Supporting information), whose CH4 selectivity is also better than other previously reported catalysts (Table S3 in Supporting information). Other doped samples with different Ca contents were also conducted for ECRR performance, which showed the Ca doping influenced the selectivity of CH4 better compared to La2CuO4 (Fig. S9 in Supporting information), and the LaCa0.4CuO3-δ performed best among these samples. Tafel slopes were also performed to understand the catalytic kinetics (Fig. S10 in Supporting information). LaCa0.4CuO3-δ gave a lower Tafel slope value than La2CuO4 and CuO, reflecting the fast catalytic kinetics of LaCa0.4CuO3-δ to produce CH4. Furthermore, the stability of LaCa0.4CuO3-δ under a constant potential of −1.3 V (vs. RHE) was evaluated in the H-cell for more than 9200 s (about 2.5 h) with an average FECH of about 54.6% (Fig. S11 in Supporting information).
|
Download:
|
| Fig. 3. FEs for ECRR products of (a) LaCa0.4CuO3-δ and (b) La2CuO4 in H-cell. (c) Ratios of FECH4 to FEC2+ in H-cell for all catalysts. FEs for ECRR products of (d) La2CuO4 and (e) LaCa0.4CuO3-δ in MEA. (f) The CH4 partial current density of catalysts at different total current densities in MEA. | |
The performance of LaCa0.4CuO3-δ, La2CuO4, and commercial CuO were also evaluated in a two-electrode MEA with 1 mol/L KHCO3 as the anolyte to understand the potential for industry application. Our A-site modulation also improved the ECRR performance of the catalysts for CH4 production under high current conditions in MEA (Fig. S12 in Supporting information). Commercial CuO still has ethylene as the main product, and CH4 production is negligible (Fig. S13 in Supporting information). The FECH of La2CuO4 was slightly increased to 8.8% ± 5.6% by introducing La in the A-site of perovskite structure (Fig. 3d). Then, by doping Ca to share the A-site that La possessed, LaCa0.4CuO3-δ reached a FECH of 38.7% ± 5.7% (Fig. 3e and Fig. S14 in Supporting information), corresponding to a jCH of −154.9 mA/cm2 (Fig. 3f), showing an explicit CH4 preference in MEA (Table S4 in Supporting information). In addition, the voltage of the catalysts under MEA electrochemical reaction also showed no extinguished difference (Figs. S15a-c in Supporting information).
The samples loaded on carbon paper after electroreduction were subjected to XRD analysis. The characteristic peaks at 31.2° and 33.5°, corresponding to (131) and (200) lattice planes, respectively, are more distinctly retained in LaCa0.4CuO3-δ than La2CuO4 (Fig. 4a). Peaks at 48.1° and 58.5° were observed in LaCa0.4CuO3-δ but nearly disappeared in La2CuO4. Meanwhile, the main peaks in commercial CuO totally disappeared. Overall, the characteristic peaks in LaCa0.4CuO3-δ remained largely intact compared to those in La2CuO4, demonstrating that the introduction of Ca enhanced the stability of the perovskite structure without causing structural transformation during reduction. In addition, the atomic ratio of La/Ca/Cu in LaCa0.4CuO3-δ determined by ICP-OES also remained almost unchanged compared to that before the electroreduction (Table S2 in Supporting information).
|
Download:
|
| Fig. 4. (a) XRD patterns of commercial CuO, La2CuO4, and LaCa0.4CuO3-δ after electroreduction on carbon paper. (b) HR-TEM image of LaCa0.4CuO3-δ after electroreduction. Cu 2p XPS spectra of (c) LaCa0.4CuO3-δ and (d) La2CuO4 before and after ECRR. Cyclic voltammograms of (e) La2CuO4 and (f) LaCa0.4CuO3-δ at a scan rate of 0.1 V/s in 0.1 mol/L KHCO3. | |
After electroreduction, the morphology of LaCa0.4CuO3-δ revealed by TEM remained similar to the original sample mentioned before (Fig. S16 in Supporting information and Fig. 2d). The lattice fringe attributed to (131) and (200) facets of LaCa0.4CuO3-δ can be observed in the TEM images of LaCa0.4CuO3-δ after ECRR (Fig. 4b and Fig. S17 in Supporting information). The elemental mapping also showed a uniform distribution of elements as before (Fig. S18 in Supporting information). X-ray photoelectron spectroscopy (XPS) showed that Cu 2p in LaCa0.4CuO3-δ exhibited a main peak at 933.6 eV corresponding to Cu2+ 2p3/2 (Fig. 4c). After electroreduction, this peak attributed to Cu2+ remained stable, whereas the peak corresponding to lower valence Cu species was detected in La2CuO4 (Fig. 4d), demonstrating the effect of Ca doping in preventing the electrochemical reduction of Cu2+.
Cyclic voltammetry (CV) was performed to determine the specific valence state change at negative potentials and to understand the stability of the catalysts. Commercial CuO exhibited typical peaks for the electrochemical reduction of Cu2+ (Fig. S19 in Supporting information), with the reduction peaks at 0.5 and 0.2 V attributed to Cu2+ to Cu+ and Cu+ to Cu0, respectively [38, 39]. CV for La2CuO4 showed smaller reduction peaks corresponding to Cu2+ to Cu+ and Cu+ to Cu0 (Fig. 4e) [40]. For LaCa0.4CuO3-δ, no significant reduction peak was observed in CV curves, indicating that the introduction of Ca can stabilize the structure of LaCa0.4CuO3-δ and maintain the valence state of Cu in +2 (Fig. 4f).
Subsequently, in situ Raman spectroscopy was performed to investigate the stability of Cu species and the CO2 to CH4 mechanism in LaCa0.4CuO3-δ during the ECRR process. As shown in the ex situ Raman of the catalyst, the peak at 606 cm-1 in commercial CuO is related to Cu-O (Fig. S20a in Supporting information) [41]. La2CuO4 displayed Raman peaks at 695, 904, 1165, 1214, and 1412 cm-1 corresponding to the structure of La2CuO4 (Fig. S20b in Supporting information) [36, 42]. The peaks at 530 and 606 cm-1 in LaCa0.4CuO3-δ can be attributed to the characteristic stretching vibrations of the Cu-O bond, and the later peaks at 1164 and 1214 cm-1 can be attributed to the typical perovskite structure (Fig. S20c in Supporting information) [36, 43].
In situ Raman spectroscopy of the three catalysts were collected at the negative potential of −0.5 V vs. RHE to investigate the structure stability during electroreduction (Fig. S21 in Supporting information) [44]. The commercial CuO was reduced to metallic Cu with the peak at 606 cm-1 fading (Fig. S21a). In contrast, the other two perovskite catalysts were more stable, with the characteristic peak still present (Figs. S21b and c) [41, 45]. With more negative ECRR potentials applied, the peak at 1412 cm-1 representing La2CuO4 also gradually disappeared, and the peak corresponding to Cu-O in LaCa0.4CuO3-δ remained stable in contrast, indicating a more stable structure of LaCa0.4CuO3-δ than La2CuO4 (Fig. 5a and Fig. S22 in Supporting information) [46, 47].
|
Download:
|
| Fig. 5. In situ Raman spectra of (a) LaCa0.4CuO3-δ at different potentials, that with further Raman range of (b) LaCa0.4CuO3-δ and (c) La2CuO4. The density of states of Cu d-band and O p-band of (d) La2CuO4 and (e) LaCa0.4CuO3-δ. (f) Gibbs free energy pathway of CH4 on La2CuO4 and LaCa0.4CuO3-δ models. | |
Moreover, in situ Raman spectroscopy of LaCa0.4CuO3-δ exhibited a peak at 1398 cm-1 [38] corresponding to the *CH3O intermediate during the ECRR process (Fig. 5b) [35, 48]. It suggested that the hydrogenation process and the intermediate pathway starting from *CHO for CH4 formation were more favorable on the surface of LaCa0.4CuO3-δ compared to La2CuO4 (Fig. 5c). Combined with where the Cu2+ stability was achieved, it can be assumed that the Cu2+ species in LaCa0.4CuO3-δ facilitates the hydrogenation process and stabilizes the *CH3O intermediates for higher CH4 selectivity.
In order to further elucidate the reaction mechanism of the catalyst in the reaction process, DFT calculations were performed on La2CuO4 and LaCa0.4CuO3-δ (Figs. S23-S27 in Supporting information). The density of states (DOS) of Cu d-band and O p-band of La2CuO4 and LaCa0.4CuO3-δ were analyzed for the electronic property changes due to the doping (Figs. 5d and e). Since the center of the Cu p-band of LaCa0.4CuO3-δ was more approaching to the Fermi energy, more absorbed electronics on the surface of LaCa0.4CuO3-δ achieved better catalytic performance with higher electron occupation number, forming a new electronic state for intrinsic activity [49, 50]. As displayed in Fig. 5f, the rate-determining step of the ECRR on La2CuO4 was the hydrogenation of *CH2O to *CH3O, requiring 1.8 eV to overcome the energy barrier. By contrast, the rate-determining step on LaCa0.4CuO3-δ was the hydrogenation of *CO to *CHO, only 1.2 eV was required to overcome the energy barrier for further processing, making it easier for LaCa0.4CuO3-δ to form the subsequent intermediates *CH2O and *CH3O, which was in agreement with in situ Raman spectroscopy. The DFT calculation results indicated that LaCa0.4CuO3-δ lowered the energy barrier of hydrogenation of *CH2O in the electroreduction process and facilitated the intermediate hydrogenation pathway to CH4.
In summary, we developed a strategy for A-site modulation of La2CuO4 to stabilize Cu2+ for efficient electrochemical CO2 reduction (ECRR) to produce CH4. Specifically, a perovskite catalyst of LaCa0.4CuO3-δ was prepared by a sol-gel procedure. The structure of LaCa0.4CuO3-δ remains unchanged during ECRR. The LaCa0.4CuO3-δ exhibits a FECH of 59.6% at −1.3 V vs. RHE in H-cell and a FECH of 38.7% with a large partial current density of −155.0 mA/cm2 in MEA. In situ Raman spectra and DFT calculations confirmed LaCa0.4CuO3-δ facilitates the hydrogenation from *CH2O to *CH3O to produce CH4. This study provides an efficient strategy to stabilize Cu2+ and guides the further study of perovskite in ECRR.
Declaration of competing interestThe authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
CRediT authorship contribution statementYuhan Zheng: Writing – original draft, Visualization, Validation, Methodology, Investigation, Conceptualization. Yunzhen Jia: Writing – original draft, Software. Xuelei Lang: Methodology. Dazhong Zhong: Writing – review & editing, Writing – original draft, Supervision, Methodology, Funding acquisition, Formal analysis. Jinping Li: Supervision, Resources. Qiang Zhao: Writing – review & editing, Supervision, Funding acquisition, Formal analysis, Data curation.
AcknowledgmentsThe authors gratefully acknowledge financial support from the National Natural Science Foundation of China (Nos. 22308246, 22478278), Central Government Guides the Special Fund Projects of Local Scientific and Technological Development (No. YDZJSX20231A015), and the Fundamental Research Program of Shanxi Province (No. 202203021212266).
Supplementary materialsSupplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111193.
| [1] |
R. Zhao, P. Ding, P. Wei, et al., Adv. Funct. Mater. 31 (2021) 2009449. |
| [2] |
A. Bagger, W. Ju, A.S. Varela, et al., ChemPhysChem 18 (2017) 3266-3273. DOI:10.1002/cphc.201700736 |
| [3] |
X. Jiang, Y. Zhao, Y. Kong, et al., Chin. Chem. Lett. 36 (2025) 109555. |
| [4] |
X. Wang, Y. Wang, L. Cui, et al., Chin. Chem. Lett. 35 (2024) 110031. |
| [5] |
Y. Wang, J. Liu, G. Zheng, Adv. Mater. 33 (2021) e2005798. |
| [6] |
W. Ma, X. He, W. Wang, et al., Chem. Soc. Rev. 50 (2021) 12897-12914. DOI:10.1039/d1cs00535a |
| [7] |
J. Zhu, L. Lv, S. Zaman, et al., Energy Environ. Sci. 16 (2023) 4812-4833. DOI:10.1039/d3ee02196c |
| [8] |
M. Huang, S. Gong, C. Wang, et al., Angew. Chem. Int. Ed. 60 (2021) 23002-23009. DOI:10.1002/anie.202110594 |
| [9] |
L.L. Zhuo, P. Chen, K. Zheng, et al., Angew. Chem. Int. Ed. 61 (2022) e202204967. |
| [10] |
J. Zhao, P. Zhang, T. Yuan, et al., J. Am. Chem. Soc. 145 (2023) 6622-6627. DOI:10.1021/jacs.2c12006 |
| [11] |
L. Xue, C. Zhang, J. Wu, et al., Appl. Catal. B 304 (2022) 120951. |
| [12] |
Y. Zhao, Q. Yuan, M. Fan, et al., Chin. Chem. Lett. 34 (2023) 108120. |
| [13] |
Y. Zhang, X. Zhang, W. Sun, ACS Catal. 13 (2023) 1545-1553. DOI:10.1021/acscatal.2c05538 |
| [14] |
X. Zhi, A. Vasileff, Y. Zheng, et al., Energy Environ. Sci. 14 (2021) 3912-3930. DOI:10.1039/d1ee00740h |
| [15] |
A. Vasileff, C. Xu, Y. Jiao, et al., Chem 4 (2018) 1809-1831. |
| [16] |
L. Zhan, Y. Wang, M. Liu, et al., Chin. Chem. Lett. 36 (2025) 109695. |
| [17] |
A. Bagger, W. Ju, A.S. Varela, et al., Catal. Today 288 (2017) 74-78. |
| [18] |
H. Sun, J. Liu, Chin. Chem. Lett. 34 (2023) 108018. |
| [19] |
Y. Zhang, Q. Zhou, X. Lu, et al., CCS Chem. 6 (2024) 2950-2960. |
| [20] |
Y. Shangguan, F. Li, X. Feng, et al., Chem. Commun. 60 (2024) 13726-13729. DOI:10.1039/d4cc05372a |
| [21] |
S. Deng, R. Wang, X. Feng, et al., Angew. Chem. 135 (2023) e202309625. |
| [22] |
X. Zhou, J. Shan, L. Chen, et al., J. Am. Chem. Soc. 144 (2022) 2079-2084. DOI:10.1021/jacs.1c12212 |
| [23] |
X. Chen, S. Jia, C. Chen, et al., Adv. Mater. 36 (2023) 2310273. |
| [24] |
Y. Li, Y. Zhang, L. Shi, et al., Small 20 (2024) 2402823. |
| [25] |
Q. Lei, H. Zhu, K. Song, et al., J. Am. Chem. Soc. 142 (2020) 4213-4222. DOI:10.1021/jacs.9b11790 |
| [26] |
X. Xu, W. Wang, W. Zhou, Z. Shao, Small Methods 2 (2018) 1800071. |
| [27] |
S. Khamgaonkar, Q. Chen, K. Musselman, V. Maheshwari, J. Mater. Chem. A 11 (2023) 20079-20088. DOI:10.1039/d3ta04226j |
| [28] |
H. Zhang, Y. Xu, M. Lu, et al., Adv. Funct. Mater. 31 (2021) 2101872. |
| [29] |
Y. Sun, R. Li, X. Chen, et al., Adv. Energy Mater. 11 (2021) 2003755. |
| [30] |
G.M. Dalpian, Q. Liu, C.C. Stoumpos, et al., Phys. Rev. Mater. 1 (2017) 025401. DOI:10.1103/PhysRevMaterials.1.025401 |
| [31] |
W. Hu, M. Zheng, H. Duan, et al., Chin. Chem. Lett. 33 (2022) 1412-1416. |
| [32] |
Z. Gu, J. Le, H. Wei, et al., Chin. Chem. Lett. 35 (2024) 108849. |
| [33] |
F. Wang, W. Sun, ACS Sustainable Chem. Eng. 12 (2024) 15651-15658. DOI:10.1021/acssuschemeng.4c06093 |
| [34] |
F. Wang, J. Chen, W. Sun, Chin. J. Inorg. Chem. 41 (2025) 97-104. |
| [35] |
J. Zhao, P. Zhang, T. Yuan, et al., J. Am. Chem. Soc. 145 (2023) 6622-6627. DOI:10.1021/jacs.2c12006 |
| [36] |
S. Chen, Y. Su, P. Deng, et al., ACS Catal. 10 (2020) 4640-4646. DOI:10.1021/acscatal.0c00847 |
| [37] |
Z. Xu, C. Peng, G. Luo, et al., Adv. Energy Mater. 13 (2023) 2204417. |
| [38] |
D. Zhong, D. Cheng, Q. Fang, et al., Chem. Eng. J. 470 (2023) 143907. |
| [39] |
D. Ren, N.T. Wong, A.D. Handoko, et al., J. Phys. Chem. Lett. 7 (2016) 20-24. DOI:10.1021/acs.jpclett.5b02554 |
| [40] |
W. Tang, A.A. Peterson, A.S. Varela, et al., Phys. Chem. Chem. Phys. 14 (2012) 76-81. |
| [41] |
Y. Deng, A.D. Handoko, Y. Du, et al., ACS Catal. 6 (2016) 2473-2481. DOI:10.1021/acscatal.6b00205 |
| [42] |
N. Ekthammathat, T. Thongtem, S. Thongtem, Appl. Surf. Sci. 277 (2013) 211-217. |
| [43] |
M. Balık, V. Bulut, I.Y. Erdogan, Int. J. Hydrogen Energy 44 (2019) 18744-18755. |
| [44] |
W.K. Han, J.W. Choi, G.H. Hwang, et al., Appl. Surf. Sci. 252 (2006) 2832-2838. |
| [45] |
D. Ren, Y. Deng, A.D. Handoko, et al., ACS Catal. 5 (2015) 2814-2821. DOI:10.1021/cs502128q |
| [46] |
G. Kliche, Z.V. Popovic, Phys. Rev. B 42 (1990) 10060-10066. |
| [47] |
S.Karthick Kumar, S. Murugesan, S. Suresh, Mater. Chem. Phys. 143 (2014) 1209-1214. |
| [48] |
H. Matsuura, K. Fukuhara, J. Polym. Sci. B: Polym. Phys. 24 (1986) 1383-1400. |
| [49] |
X.G. Zhang, S. Feng, C. Zhan, et al., J. Phys. Chem. Lett. 11 (2020) 6593-6599. DOI:10.1021/acs.jpclett.0c01970 |
| [50] |
Y. Zhang, Y. Zhao, C. Wang, et al., Phys. Chem. Chem. Phys. 21 (2019) 21341-21348. DOI:10.1039/c9cp03692j |
2025, Vol. 36 

