Dongxing Tan†
*,
Xianfang Yin†,
Hengrui Kang,
Jing Wang,
Dan Zhang and
Yuanyuan Feng*
Key Laboratory of Catalytic Conversion and Clean Energy in Universities of Shandong Province, School of Chemistry and Chemical Engineering, Qufu Normal University, Qufu, Shandong 273165, P. R. China. E-mail: tandx@qfnu.edu.cn; fengyy@qfnu.edu.cn
First published on 11th August 2025
We employed a Lewis acid doping strategy to create stable In/In2O3 interfaces and oxygen vacancies during the electrocatalytic reduction of CO2, thereby enhancing the selectivity of formate. The optimized catalyst achieves a high faradaic efficiency for formate (exceeding 90%), across a broad range of potentials.
Beyond conventional strategies (e.g., morphology control, composition optimization, and defect engineering), precise regulation of surface oxygen vacancies and construction of In/In2O3 heterostructures are pivotal for improving the selectivity of formate.9–11 Oxygen vacancies could reduce the energy barrier and stabilize the critical OCHO* intermediate. Concurrently, the Schottky effect at In/In2O3 heterointerfaces modulates the electronic structure of the active sites, further promoting OCHO* formation and catalytic activity.12 Therefore, integrating abundant, stable oxygen vacancies with well-defined In/In2O3 heterostructures effectively activates CO2 and improves formate production activity.
Herein, we constructed a catalyst with abundant oxygen vacancies and stabilized In/In2O3 interfaces by a Lewis acidic metal-doping strategy. Lewis acid doping not only effectively suppresses the excessive reduction of In2O3 during the electrocatalytic process and maintains the stability of the In/In2O3 interfacial structure, but also increases the content of oxygen vacancies on the catalyst surface. Benefiting from the synergistic effect of In/In2O3 heterointerfaces and oxygen vacancies, the catalyst demonstrates superior formate selectivity. When paired with a commercial solar cell, the solar-driven synchronous CO2 reduction and methanol oxidation electrochemical cell demonstrates a sustainable coelectrolysis process for formate production.
The target catalyst was synthesized via a three-step procedure (Fig. 1a). Initially, Al-doped MIL-68(In) metal–organic frameworks (Al–In MOFs) were synthesized solvothermally. Subsequently, Al-doped In2O3 was obtained by calcining the Al–In MOFs in air. Finally, Al-doped In2O3 underwent pre-reduction at −1.0 V for 1 h to stabilize its structure and composition. A reference catalyst was synthesized identically without Al(NO3)3·9H3O addition. Scanning electron microscopy (SEM) revealed that the Al–In MOFs possess a well-defined hexagonal morphology (Fig. S1). Calcination decomposed the organic framework, yielding a porous Al-doped In2O3 structure assembled from nanoparticles (Fig. 1b), with Al doping exhibiting negligible morphological influence (Fig. S2). Transmission electron microscopy (TEM) confirmed nanoparticles with an average size of about 30 nm (Fig. 1c). The XRD patterns showed the absence of distinct Al-related peaks, suggesting the presence of amorphous Al species (Fig. S3).13 The slight shift of the diffraction lines toward higher angles confirms the substitution of smaller-radius Al3+ ions for In3+ within the In2O3 lattice. After pre-reduction at −1.0 V for 1 h, Al-doped In2O3 was partially reduced to metallic In, yielding coexisting In and In2O3 phases (denoted Al–In/In2O3) (Fig. S4). HRTEM revealed lattice spacings of 0.247 and 0.253 nm, corresponding to the In (002) and In2O3 (400) planes, respectively, and a clear grain boundary, indicating the formation of an In/In2O3 interface (Fig. 1d).14 This interface enhances CO2 activation and facilitates rapid electron transfer, optimizing the electrocatalytic conversion.15 In contrast to Al-doped In2O3, pre-reduction almost completely converted In2O3 to metallic In (denoted R–In) (Fig. S5). Energy dispersive X-ray spectroscopy (EDS) elemental mapping of Al–In/In2O3 showed the homogeneous distribution of In, Al, O, and C (Fig. 1e).
XPS was employed to investigate the surface electronic configurations and oxidation states of the catalysts. As shown in Fig. S6a, the high-resolution In 3d spectra of Al–In2O3 exhibit a 0.3 eV negative shift relative to In2O3. The high-resolution O 1s XPS spectra indicated that the Al–In2O3 surface possesses a higher content of oxygen vacancies compared to In2O3 (Fig. S6b). Furthermore, the high-resolution Al 2p spectrum indicated the presence of Al3+ species (Fig. S6c). These XPS results demonstrate that the introduction of Lewis acidic Al species effectively modulates the surface electronic structure of the catalyst.16 Following pre-reduction, significant changes occur in the surface chemical valence states. Compared to R–In, Al–In/In2O3 can maintain a higher ratio of In3+ species and oxygen vacancy content (Fig. S7).17,18 EPR spectroscopy further confirmed that Lewis acidic metal doping significantly enhanced the concentration of surface oxygen vacancies on Al–In/In2O3 (Fig. S8). Semi in situ XRD patterns were used to track the derivation of the crystal structure of the catalyst during the pre-reduction process. Fig. S9 reveals the persistent coexistence of metallic In0 and In2O3 phases within Al–In/In2O3 after 1 h of reduction. These results indicate that the Lewis acid metal doping strategy effectively maintains the stability of the In/In2O3 heterointerfaces and preserves a high oxygen vacancy content during the electrocatalytic CO2 reduction process. The In/In2O3 heterointerfaces and abundance of oxygen vacancies in the Al–In/In2O3 catalyst can enhance the activation of CO2 and improve the selectivity of formate.19
The CO2RR performances of the Al–In/In2O3 and R–In catalysts were systematically evaluated in a H-cell with a 0.1 M K2SO4 electrolyte. The linear scan voltammetry (LSV) curves revealed enhanced current densities in a CO2-saturated electrolyte compared to an Ar-saturated electrolyte for both catalysts. Notably, the Al–In/In2O3 catalyst demonstrated a higher current density in the CO2-saturated electrolyte than that of R–In, indicating enhanced CO2RR activity (Fig. 2a). Gas chromatography (GC) and nuclear magnetic resonance (NMR) were employed to analyze the gas-phase and liquid-phase products of the CO2RR, respectively. The reduction products of the electrocatalytic CO2RR for Al–In/In2O3 and R–In catalysts were formate, CO, and H2 from the competitive HER (Fig. S10). As expected, the Al–In/In2O3 catalyst exhibited higher selectivity for formate. The maximum FE for the Al–In/In2O3 catalyst is up to 94% at −1.0 V, and remains above 90% over a potential range of −0.8 to −1.3 V (Fig. 2b). The FE of H2 for Al–In/In2O3 is only 3% at −1.0 V, which is significantly lower than that of R–In (14%) (Fig. S11). This observation suggests that the incorporation of Lewis acidic Al species could enhance the formate selectivity. Based on the high current density and FE, the Al–In/In2O3 catalyst exhibits higher partial current density (Fig. 2c).
The electrochemical active surface areas (ECSAs) of Al–In/In2O3 and R–In were further evaluated (Fig. S12). As shown in Fig. 2d, the double-layer capacitance of Al–In/In2O3 is 1.2 times that of R–In. This suggests that abundant oxygen vacancies and In/In2O3 interfaces can provide more active sites to enhance the CO2 catalytic activity.20 The smaller charge transfer impedance of Al–In/In2O3 also further indicates the rapid charge transfer kinetics (Fig. 2e).21 Hydroxyl (OH−) adsorption was utilized as an alternative test to investigate the binding affinity of the catalysts to the *CO2•− intermediate. Al–In/In2O3 exhibited a lower OH− adsorption potential, indicating strong adsorption capacity for *CO2•− (Fig. 2f). This indicates that Al–In/In2O3 could stabilize the intermediate and further converts into formate.22 Benefiting from robust In/In2O3 heterointerfaces and abundant oxygen vacancies, Al–In/In2O3 achieves formate selectivity on par with the state-of-the-art catalysts reported (Fig. 2g and Table S1).10,11,15,23–36 Moreover, stability tests revealed no significant decay in current density and FE of formate during 30 h of continuous electrolysis, highlighting the good stability of the active sites (Fig. 2h). A series of Al-doped In2O3 catalysts with varying Al molar ratios demonstrate that Lewis acid metal doping effectively enhances the selectivity towards formate, with the selectivity exhibiting an initial increase followed by a decrease as the doping concentration increases (Fig. S13–S15).
Although Al–In/In2O3 showed excellent FE of formate in the H-cell, the low CO2 solubility in the neutral electrolyte greatly limited the current density of the catalytic reaction. To overcome the mass transport bottleneck, the electrocatalytic CO2RR activity of the catalysts was further investigated in a flow cell equipped with gas diffusion electrodes (GDE) (Fig. S16). This advanced reactor allows for the direct supply of gaseous CO2 to the catalyst layer, effectively bypassing the limitations of CO2 solubility. As anticipated, the current density of Al–In/In2O3 was dramatically enhanced in a flow cell, reaching 264 mA cm−2 at −1.0 V, which is 6 times that of the H-cell (Fig. 3a). Moreover, the alkaline electrolyte further increases the catalytic activity (Fig. 3b, c and Fig. S17). Additionally, long-term stability tests revealed no significant decay in current density and FE of formate, confirming the exceptional stability of the catalyst's active sites (Fig. 3d).
![]() | ||
Fig. 3 (a) LSV curves, (b) FE of formate, (c) partial current density of formate for Al–In/In2O3 and R–In in a flow cell. (d) Stability test of the Al–In/In2O3 catalyst at −0.8 V in a flow cell. |
Mn, Cd, and Ce-doped In2O3 catalysts were synthesized to elucidate the Lewis acid species doping in modulating catalytic performance. XRD patterns and SEM images confirm that the Mn, Cd and Ce-doped In2O3 catalysts exhibit morphology and crystal structure similar to Al–In2O3 (Fig. S18 and S19). The high-resolution O 1s spectra revealed significantly enhanced surface oxygen vacancy concentrations compared to In2O3 (Fig. S20). Furthermore, the positive correlation between oxygen vacancy content and catalytic activity demonstrates that Lewis acid doping enhances CO2RR performance by facilitating oxygen vacancy formation (Fig. S21).
Although Al–In/In2O3 can effectively reduce CO2 into formate, the slow kinetics of the anodic oxygen evolution reaction (OER) increases the energy input of the entire catalytic reaction. To address this limitation, we propose a coupled electrolysis strategy integrating the CO2RR with the methanol oxidation reaction (MOR) to reduce energy consumption. The MOR is characterized by a low thermodynamic equilibrium potential (0.103 V) for efficient conversion of methanol into formate. According to the literature reported method, the Ni0.2Mo0.8N/F,N–C@NF heterogeneous catalyst was prepared (Fig. S22 and S23).37 Ni0.2Mo0.8N/F,N–C@NF exhibited excellent MOR activity. The potentials required for the Ni0.2Mo0.8N/F,N–C@NF catalyst to achieve the same current density in the MOR are significantly lower compared to the OER (Fig. S24 and Table S2).
A two-electrode electrolytic cell was constructed with Al–In/In2O3 as the cathode and Ni0.2Mo0.8N/F,N–C@NF as the anode to realize the CO2RR-coupled MOR or OER. As shown in Fig. 4a, when the current density reaches 200 mA cm−2, the required potential for the CO2RR coupled MOR system decreases by 0.8 V compared to the CO2RR coupled OER system, indicating that integrating MOR with CO2RR can significantly reduce the input voltage. More importantly, the CO2RR coupled MOR simultaneously produces high value-added formate at the cathode and anode. FE of formate at both electrodes is maintained above 70% at current densities ranging from 50 to 250 mA cm−2 (Fig. 4b). This demonstrates that it is feasible to construct a two-electrode electrolytic cell for the simultaneous electrosynthesis of formate. Furthermore, utilizing the photovoltaic cell to drive the two-electrode electrolytic cell can construct a green carbon cycle system and realize the conversion of solar energy to chemical energy (Fig. 4c). The open-circuit voltage and short-circuit currents of the photovoltaic cell under simulated sunlight are 3.33 V and 163.8 mA, respectively. The intersection point of the LSV curves for the electrolytic cell and photovoltaic cell curves are 2.83 V and 159.1 mA, near the maximum power point (MPP, 2.85 V, 158.4 mA) of the photovoltaic cell (Fig. 4d). The solar driven two-electrode electrolytic cell exhibits rapid and uniform current responses when cycling light on and light off. Under prolonged illumination, the electrochemical cell demonstrates good stability with negligible current density decay and consistently maintained FE of formate, thereby robustly confirming the superior operational durability and product selectivity of the CO2RR coupled MOR system (Fig. 4e and Fig. S25).
In summary, we engineered the surface structure of In2O3 through a Lewis acid doping strategy, achieving stabilized In/In2O3 heterointerfaces and improved oxygen vacancy concentration. The experimental results clearly demonstrated that the Lewis acid doped catalyst has shown significant potential for electroreduction of CO2 into formate. Al–In/In2O3 achieves a maximum FE of 97% and maintains above 90% within a wide range of potentials. By further integrating the MOR, a two-electrode electrochemical cell for synchronous electrocatalytic generation of formate was constructed, which exhibits good catalytic efficiency and operational stability. This study demonstrated a rational approach to modulate the interfacial properties of catalysts by a Lewis acid doping strategy to enhance the generation of formate for the CO2RR.
We acknowledge the financial support from the Natural Science Foundation of Shandong Province (ZR2024QB034 and ZR2024MB078).
Footnote |
† D. T. and X. Y. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |