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

Light-induced electronic structure modulation in perovskite ferrite for efficient photothermal dry reforming of methane

Jilong Li ab, Xiang Hao ab, Jiwu Zhao ab, Jinyu Li c, Bo Su ab, Zhengxin Ding *ab, Meirong Huang ab, Zhi-An Lan ab, Min-Quan Yang *c and Sibo Wang *ab
aState Key Laboratory of Chemistry for NBC Hazards Protection, College of Chemistry, Fuzhou University, Fuzhou 350116, China
bState Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou 350116, P. R. China. E-mail: zxding@fzu.edu.cn; sibowang@fzu.edu.cn
cCollege of Environmental and Resource Sciences, College of Carbon Neutral Modern Industry, Fujian Normal University, Fuzhou, 350117, P. R. China. E-mail: yangmq@fjnu.edu.cn

Received 30th July 2025 , Accepted 24th November 2025

First published on 24th November 2025


Abstract

Solar-driven dry reforming of methane (DRM) offers a sustainable pathway to convert CH4 and CO2 into valuable syngas feedstock, yet the efficiency is hindered by the sluggish lattice oxygen (LO) migration of the catalyst and the incomplete understanding of light-enhanced redox cycling. Here, we demonstrate that Ru/LaFeO3 functions as a highly efficient and durable photothermal catalyst for DRM. The incorporation of Ru not only serves as an electron trap but also modulates the catalyst's electronic structure. Notably, under light irradiation, photoinduced charge redistribution further intensifies this electronic modulation, leading to electron enrichment at Ru, La, and Fe sites, and hole accumulation at LO sites. This interfacial charge dynamics weakens La–O and Fe–O bonds and facilitates LO migration, enabling efficient CH4 activation and oxidation at Ru sites, accompanied by the generation of oxygen vacancies (OVs). Simultaneously, the in situ generated OVs promote CO2 adsorption and activation, facilitating its cleavage into CO and replenishing the OVs, thereby sustaining the redox cycle for continuous catalysis. This study provides key mechanistic insights into photoinduced LO dynamics driven by charge redistribution, and offers valuable guidance for the rational design of advanced photothermal systems that leverage both thermal and photonic effects of solar energy for enhanced catalysis via the LO-mediated pathways.


Introduction

Dry reforming of methane (DRM) (CH4 + CO2 → 2CO + 2H2, ΔH298K = 247 kJ mol−1) offers a promising route for converting two major greenhouse gases into syngas, a valuable feedstock for various chemical process.1–4 However, due to the high bond dissociation energies of the C–H bond (434 kJ mol−1) in CH4 and the C[double bond, length as m-dash]O bond (805 kJ mol−1) in CO2,5–10 conventional thermal catalytic DRM typically proceeds at elevated temperatures of 700–1000 °C, which is highly energy-intensive and carbon emissive.11 Moreover, harsh conditions often lead to undesirable side reactions such as CH4 decomposition and CO disproportionation, resulting in carbon deposition, catalyst sintering, and deactivation.12,13

To overcome these limitations, solar-driven DRM has emerged as a sustainable and potentially low-temperature alternative.14–20 This approach is generally understood to follow a light-enhanced Mars–van Krevelen (MvK) mechanism in metal oxide catalysts.15,20–23 The key reaction process typically involves CH4 splitting into image file: d5sc05708f-t1.tif (0 ≤ x ≤ 3) and H* species, and lattice oxygen (LO) migration to react with image file: d5sc05708f-t2.tif, ultimately forming CO and suppressing coke formation.14,15,18 However, strong metal–oxygen interactions in normal metal oxides significantly limit LO mobility, restricting overall catalytic activity.24 In addition, the chemical inertness of the reactants and the multistep nature of DRM introduce further challenges in achieving efficient activation of both the C–H and C[double bond, length as m-dash]O bonds under mild conditions while controlling reaction pathways.25,26

Perovskite oxides (ABO3) have recently garnered increasing attention as promising platforms for photothermal DRM catalysis.14–16,18,20,21 These materials feature corner-sharing BO6 octahedra formed by A-site (typically rare-earth) and B-site (transition metal) ions, offering enhanced light absorption and efficient charge separation.27–32 Their compositional tunability and intrinsic LO activity enable dynamic modulation of catalytic properties.20,33 Indeed, several studies have confirmed the feasibility of light-driven DRM over perovskite-based systems.14–16,21,34 Nonetheless, developing photothermal catalysts with high efficiency and long-term stability remains a critical goal. Moreover, unlike conventional thermal catalysis that occurs in the electronic ground state, photothermal DRM operates under photoexcited states, in which the electronic structure of the catalyst is continuously modulated by photoexcited charge carriers.15,18,20 Yet, the fundamental understanding of how these photoinduced carriers influence catalytic performance remains incomplete and warrants further investigation.

Herein, we present a novel photothermal catalyst composed of Ru NPs supported on LaFeO3, which manifests excellent DRM activity and durability. Under a light intensity of 6.7 W cm−2, the catalyst delivers CO and H2 production rates of 34.14 and 27.92 mol gRu−1 h−1, respectively, with an outstanding CH4 turnover frequency (TOFCH4) of 0.64 s−1 and an excellent light-to-chemical energy efficiency (LTCEE) of 16.3%. Combined experimental and theoretical studies reveal that strong interfacial interaction between Ru and LaFeO3 enhances electron distribution at La and Fe sites, while light irradiation further strengthens such an effect. The charge redistribution of Ru/LaFeO3 results in electron enrichment on Ru, La and Fe sites, while holes are accumulated on LOs. The synergistic electronic modulation weakens La/Fe–O bonds and facilitates LO migration, enabling rapid LO-involved CH4 activation and oxidation at Ru sites, accompanied by the formation of oxygen vacancies (OVs). Concurrently, the in situ generated OVs efficiently adsorb and activate CO2, promoting its cleavage into CO, coupled with O to refill OVs, completing the catalytic cycle.

Results and discussion

The LaFeO3 perovskite catalyst was synthesized via a sol–gel method, followed by calcination to enhance the crystallinity.28 Fig. S1 displays the powder X-ray diffraction (XRD) pattern of LaFeO3, which matches well with the standard card of the orthorhombic structure (JCPDS no. 37-1493).28 Transmission electron microscopy (TEM) shows an irregular nanoparticle morphology for LaFeO3 (Fig. S2a). The high-resolution TEM (HRTEM) image presents clear lattice fringes with a d-spacing of 0.28 nm (Fig. S2b) that corresponds to the (121) plane, confirming the high crystallinity of the synthesized LaFeO3.27 Subsequently, Ru nanoparticles (NPs) were deposited onto the LaFeO3 surface by an impregnation–reduction method. Energy dispersive spectroscopy (EDS) analysis confirms the successful loading of Ru NPs (Fig. S3). Inductively coupled plasma optical emission spectrometry (ICP-OES) reveals that the actual contents of Ru loading closely match the theoretical values (Table S1). Notably, the catalyst used in all subsequent experiments is the 1.0% Ru/LaFeO3 sample unless otherwise stated, which is referred to as Ru/LaFeO3 for brevity.

TEM analysis of the Ru/LaFeO3 catalyst reveals that the introduction of Ru has minimal impact on the morphology of LaFeO3 (Fig. S4). The Ru species are dispersed as ultra-small nanoparticles with an average size of 2.3 nm on the surface of LaFeO3 (Fig. 1 and S5). In the HRTEM image (Fig. 1b), two sets of lattice fringes with d-spacings of 0.21 and 0.28 nm are observed, corresponding to the (101) plane of Ru NPs and (121) plane of LaFeO3, respectively. The EDS elemental mapping of Ru/LaFeO3 shows that Ru, La, Fe, and O are homogeneously distributed across the individual particles (Fig. 1c), suggesting a high degree of Ru dispersion. Moreover, the XRD pattern of Ru/LaFeO3 closely resembles that of pristine LaFeO3. No noticeable diffraction peaks attributable to Ru species are detected, further confirming the high dispersion and ultra-small nature of the loaded Ru NPs (Fig. 1d and S5).


image file: d5sc05708f-f1.tif
Fig. 1 (a) TEM image, (b) HRTEM image and (c) EDS elemental maps of Ru/LaFeO3. (d) XRD patterns and (e) DRS of LaFeO3 and Ru/LaFeO3 catalysts. (f) Surface local temperature evolution of LaFeO3 and Ru/LaFeO3 catalysts under 6.7 W cm−2 light irradiation.

X-ray photoelectron spectroscopy (XPS) was applied to check the chemical states of elements in the samples (Fig. S6). The Ru 3p spectrum of Ru/LaFeO3 reveals that Ru species are mainly present as metallic Ru0 (462.8 and 484.9 eV), along with a minor portion of Ru4+ (465.3 and 487.4 eV) (Fig. S6a).28 In the Fe 2p region (Fig. S6b), the peaks at 711.1 and 724.5 eV correspond to Fe3+, whereas those at 709.6 and 723.1 eV are assigned to Fe2+.29 The La 3d spectra exhibit characteristic doublets (Fig. S6c), which can be deconvoluted into La 3d5/2 (833.59, 835.70, 837.72, and 838.75 eV) and La 3d3/2 (850.42, 851.90, 854.55, and 855.59 eV). In the La 3d5/2 region, the spectrum can be deconvoluted into two distinct doublets. The first doublet corresponds to the lattice La–O species in the perovskite framework, with the main peak at 833.59 eV and a satellite at 837.72 eV. The second doublet arises from La(OH)3, featuring a main peak at 835.70 eV and its associated satellite at 838.75 eV.15 The O 1s spectrum consists of two components at 529.3 and 531.4 eV (Fig. S6d), assigned to LO and adsorbed oxygen (AO), respectively. The relative intensity of the AO component is generally correlated with the density of surface OV, providing valuable information on the defect chemistry of the catalyst.

Notably, XPS reveals strong electronic interactions between Ru NPs and LaFeO3, which change the electron density of LaFeO3. After Ru loading, the Fe species in Ru/LaFeO3 exhibit a more electron-rich state compared to those in bare LaFeO3.15 Meanwhile, the surface La–OH species undergo a dehydration process to form La–O bonds, accompanied by the generation of OVs, as evidenced by changes in the O 1s spectrum. The formation of these oxygen vacancies facilitates the activation of the C[double bond, length as m-dash]O bond in CO2 and thus promotes the DRM reaction.35,36

The light harvesting properties of the catalysts have been investigated using UV-Vis diffuse reflectance spectroscopy (DRS). Pristine LaFeO3 exhibits an intrinsic interband transition below 550 nm, with a bandgap energy of 2.12 eV as determined from the Tauc plot (Fig. S7a). The Mott–Schottky curves indicate that the conduction band (CB) edge of LaFeO3 is positioned at −0.83 V (vs. NHE) (Fig. S7b). Combined with the optical band gap, the valence band (VB) edge is estimated to be 1.29 V (Fig. S7c). Compared with bare LaFeO3, Ru loading greatly enhances light absorption across a broad spectral range from the ultraviolet to the visible-near-IR region (200–800 nm), endowing Ru/LaFeO3 with excellent “nano-heater” functionality (Fig. 1e).

Comparison of the VB potentials of Ru/LaFeO3 and LaFeO3 indicates that Ru incorporation noticeably modulates the electronic structure of LaFeO3 (Fig. S7d–f), which strengthens its oxidation capability. This modulation is further corroborated by the distinct binding energy shifts observed in the O 2p region of the VB XPS spectra (Fig. S8). On this basis, the photothermal conversion performances of the catalysts were evaluated under same light intensities using a thermocouple placed in close contact with the catalyst surface. At a light intensity of 6.7 W cm−2, the surface temperature of the Ru/LaFeO3 catalyst reaches 520 °C, which is much higher than that of bare LaFeO3 (Fig. 1f). The results highlight the strong potential of using Ru/LaFeO3 as an efficient light harvesting platform to drive DRM under relatively mild conditions.

Using a fixed-bed continuous flow system without external heating, the light-driven DRM performance of the catalysts was evaluated. As shown in Fig. S9, the optimal Ru loading is determined to be 1%, at which the Ru/LaFeO3 sample shows the best catalytic performance. The production rates of CO and H2 reach 34.14 and 27.92 mol gRu−1 h−1, respectively, which are comparable or better than those of state-of-the-art catalysts.14–18,21,25,37 Isotope labelling experiments confirm that the generated CO comes from the conversion of CH4 and CO2 (Fig. S10). Notably, pristine LaFeO3 exhibits no activity under light irradiation (Fig. 2a), indicating the essential role of Ru species in this catalytic process. For comparison, the catalytic performances of Ru/Fe2O3 and Ru/La2O3 were also tested (Fig. 2a), which exhibit markedly weakened performance compared to Ru/LaFeO3. These results highlight that the perovskite structure of LaFeO3 plays a critical role in enabling high activity for photothermal DRM.15,31


image file: d5sc05708f-f2.tif
Fig. 2 (a) DRM performances of different catalysts under light-driven conditions (catalyst: 80 mg; light intensity: 6.7 W cm−2; feedstocks: CO2/CH4/He = 47/47/6, GHSV = 22.5 L g−1 h−1). (b) DRM activities of Ru/LaFeO3 under photothermal and thermal conditions. (c) CH4 conversion under photothermal and thermal conditions at different temperatures. (d) H2/CO ratios of DRM at different temperatures powered by pure light irradiation and external heating. (e) LTCEEs of Ru/LaFeO3 under different incident intensities. (f) Arrhenius plots of CH4 in photothermocatalysis and thermocatalysis by Ru/LaFeO3. (g) Stability test of Ru/LaFeO3 in light-driven DRM.

Moreover, the thermocatalytic DRM performance of the Ru/LaFeO3 catalyst driven by external heating under equivalent temperature as that of photothermal conditions was also evaluated. Compared to the thermal catalytic process (Fig. 2b), H2 and CO production of Ru/LaFeO3 under photothermal conditions increased by about 2.4 and 1.7 times, respectively. In particular, in the temperature range of 350–550 °C, the efficiencies of photothermal DRM consistently outperform those of thermocatalysis (Fig. S11). At a catalyst surface temperature of 550 °C, CH4 conversion reaches 38.7%, corresponding to a TOFCH4 0.64 s−1,2 denoting excellent CH4 conversion capability. Importantly, the CH4 conversions of light-driven DRM significantly exceed the thermal equilibrium limits within the examined temperature range (Fig. 2c). This implies that the light-driven process integrates thermocatalytic and photocatalytic processes, that is, involving solar heating and photo-excitation.19,38,39 Under a constant light intensity of 6.7 W cm−2 by adjusting the distance between the light source and the catalyst, narrowing the light spectrum leads to a gradual decline in the catalytic activity of Ru/LaFeO3 (Fig. S12). Nevertheless, the photothermal DRM efficiencies remain much higher than those achieved under purely thermal conditions, confirming the essential contribution of photoexcited charges in promoting the reaction.

In the DRM process, the water–gas conversion reaction is an unavoidable reverse reaction, which generally results in a low ratio of H2/CO.19,40 However, under light-driven photothermal conditions, the H2/CO ratios are significantly improved compared to those under thermal-driven process (Fig. 2d). The finding highlights the key role of the light-induced effects in regulating the reaction pathway. LTCEEs measured under different light intensities reveal that the Ru/LaFeO3 catalyst exhibits increased LTCEEs with enhancing light intensity, reaching an impressive 16.3% at 7.2 W cm−2 (Fig. 2e), which is higher than the latest reported value.15,18,20 Furthermore, the Arrhenius plot analysis reveals that the apparent activation energy (Ea) for thermal catalysis (84.0 kJ mol−1) on the Ru/LaFeO3 catalyst is much higher than that for photothermal DRM (59.1 kJ mol−1) (Fig. 2f), indicating that light irradiation alters the reaction pathway, thereby lowering the apparent reaction energy barrier.33 Finally, the durability test of the Ru/LaFeO3 catalyst reveals outstanding stability, which retains stable photothermal DRM activity over 100 h, delivering a turnover number of 3.54 × 105 calculated by the H2 production.2 XPS analysis reveals an increased proportion of lower-valent Ru and Fe species in Ru/LaFeO3 after the reaction (Fig. S13 and S14). This can be ascribed to the strongly reducing atmosphere, which drives the metal sites toward lower oxidation states.19,36 However, the XRD pattern of the post-reaction Ru/LaFeO3 catalyst shows no detectable structural changes (Fig. S15). Together with the stability tests, these results suggest that the reduction process does not compromise the catalyst's performance significantly, which may point to the involvement of reversible structural dynamics. In sharp contrast, XRD analysis shows that Ru/Fe2O3 is transformed into Ru/Fe3O4 after only 10 min of reaction (Fig. S16). These results further verify that the perovskite structure of LaFeO3 ensures both high activity and robust stability in photothermal DRM.15,31

To investigate the origin of the high performance of Ru/LaFeO3 in the light-driven DRM reaction, temperature-programmed desorption (TPD) tests were conducted to assess the interactions between reactant molecules (CH4 and CO2) and the catalysts. In the CH4-TPD profiles (Fig. S17), the desorption peaks of Ru/LaFeO3 appeared at higher temperatures and with stronger intensities compared with those of bare LaFeO3, suggesting that the loading of Ru provides stronger active sites and promotes the chemisorption and activation of CH4.15 In contrast, the CO2-TPD results show that upon Ru loading, the physical adsorption of CO2 becomes weaker while chemisorption is significantly enhanced (Fig. 3a).17,41 This can be attributed to the electron transfer from loaded Ru NPs to LaFeO3 (as evidenced by XPS analysis), which increases the surface electron density and strengthens the CO2-catalyst interactions, thus shifting CO2 adsorption from physisorption to chemisorption.35,36 The chemisorbed CO2 is more likely to participate in the subsequent catalytic reactions. Fig. 3b shows the linear scanning voltammetry (LSV) measurement, which shows an increase in current density and a decrease in onset potential for Ru/LaFeO3, providing further evidence of its enhanced CO2 activation capacity.42


image file: d5sc05708f-f3.tif
Fig. 3 (a) CO2-TPD, (b) LSV curves, (c) steady-state PL spectra, (d) TRPL spectra, (e) transient photocurrent responses, and (f) EIS Nyquist plots of LaFeO3 and Ru/LaFeO3.

Moreover, the separation and migration kinetics of photoinduced carriers were systematically investigated using photoluminescence (PL) and electrochemical techniques. Steady-state PL and time-resolved PL (TRPL) analyses indicate that Ru/LaFeO3 effectively inhibits the recombination of charge carriers and promotes their separation compared to bare LaFeO3 (Fig. 3c and d).43–47 In addition, the photocurrent response of Ru/LaFeO3 is markedly higher than that of LaFeO3 (Fig. 3e), which is consistent with the electrochemical impedance spectroscopy (EIS) results showing a lower charge-transfer resistance of Ru/LaFeO3 (Fig. 3f).48–52 These findings collectively verify that Ru modification facilitates the separation and utilization of photoinduced charges, thereby contributing to enhancing the DRM activity under light irradiation.

To further identify the active sites in Ru/LaFeO3 for CH4 and CO2 conversion, light-driven control experiments were conducted under pure CH4 and CO2 atmospheres, respectively.20 The results show that CO and H2 are detected on Ru/LaFeO3 in a CH4 atmosphere, whereas no products are observed on bare LaFeO3 (Fig. S18). Moreover, no detectable products generate on LaFeO3 under a pure CO2 atmosphere, regardless of the presence or absence of Ru species (Fig. S19). These findings indicate that Ru NPs serve as the active sites for CH4 dissociation. Notably, during the reaction under pure CH4, CO production on Ru/LaFeO3 rapidly ceases, while the production of H2 gradually decreases (Fig. S20). Meanwhile, the electron paramagnetic resonance (EPR) analysis of Ru/LaFeO3 after reaction in pure CH4 shows an obviously enhanced OV signal (Fig. S21). This phenomenon can be attributed to the oxidation of CH4 by surface-active LOs; as LOs become depleted, deep dehydrogenation of CH4 leads to carbon deposition that blocks the active site (Fig. S22). Furthermore, when the deactivated Ru/LaFeO3 catalyst (which, after reaction in CH4, is subsequently exposed to pure CO2 under light radiation), CO generation is clearly detected (Fig. S23). Correspondingly, the EPR analysis shows diminished OVs signal on the Ru/LaFeO3, indicating the replenishment of oxygen atoms into the vacancies (Fig. S21). These observations, combined with the absence of CO production over LaFeO3 under identical conditions, indicate that the OVs generated in LaFeO3 serve as the active centers for CO2 reduction to CO. This mechanism is further supported by the 18O-labelling experiments, which confirm the involvement of LOs in photothermal DRM (Fig. S24). In brief, the LOs in Ru/LaFeO3 facilitate the oxidation of CH4 to generate CO and H2, while simultaneously creating OVs. These vacancies then dissociate CO2 to produce CO and replenish O atoms, thus regenerating the LOs.14,20,53

To gain deeper insight into the photothermal catalytic mechanism of DRM over Ru/LaFeO3, a series of in situ characterization experiments and theoretical calculations were performed. Under light irradiation, the in situ XPS (ISI-XPS) measurement for the Ru/LaFeO3 catalyst exhibits negative shifts in the binding energies of Ru 3p, Fe 2p, and La 3d, along with a positive shift in O 1s. These shifts indicate the accumulation of photoexcited electrons on the metal sites of Ru, Fe, and La, and the localization of photogenerated holes on LOs (Fig. 4a–c and S25).15,20 Under the reducing influence of photogenerated electrons, the catalyst undergoes electronic rearrangement, leading to a shift of Ru and Fe species toward lower oxidation states (Fig. S26). Meanwhile, the decreased proportion of La–OH can be ascribed to light-induced desorption of surface-adsorbed species, a phenomenon that is also reflected in the O 1s spectrum. This light-induced electronic restructuring in the catalyst is pivotal for enhancing both the efficiency and stability of the DRM reaction, as discussed below. The photoexcited electrons captured at the Ru site facilitate the H* to H2 transition. Simultaneously, electron accumulation on Fe and La species enhances their electron densities, weakening their bonding with oxygen atoms and promoting the mobility of LOs.18 The hole accumulation on the LOs enhances their oxidative reactivity, thereby improving CH4 conversion.54 Additionally, the elevated electron densities at Fe and La sites reduce the reactive energy barriers for CO2 molecules trapped by OVs, thus promoting both CO2 activation and the regeneration of LOs.35,36


image file: d5sc05708f-f4.tif
Fig. 4 ISI-XPS spectra of (a) Fe 2p, (b) La 3d and (c) O 1s of Ru/LaFeO3. (d and e) In situ DRIFTS spectra of Ru/LaFeO3 catalysts in CO2/CH4-saturated atmosphere under (d) external heating and (e) light irradiation.

To track the key reaction intermediates, in situ DRIFTS was performed on the catalysts in a DRM atmosphere. The DRIFTS spectra of Ru/LaFeO3 recorded in the dark show effective adsorption of CO2 (2300–2400 cm−1) and CH4 (3016 and 1242–1350 cm−1) (Fig. S27).55–58 After reaching adsorption equilibrium, the IR spectrum was recorded as a baseline. Upon heating, inverted peaks of CO2 and CH4 emerged under thermal-driven conditions (Fig. S28), indicating the consumption of adsorbed CH4 and CO2 molecules. Meanwhile, signals ascribed to bidentate carbonate (b-CO32−, 1230 cm−1)17 and CO (2112 and 2184 cm−1) appear,59 indicating CO2 activation and CO formation (Fig. 4d). In addition, the appearance of the image file: d5sc05708f-t3.tif signal (1480 cm−1) can be attributed to the dehydrogenation of adsorbed CH4.60 Under photothermal conditions, the DRIFTS spectra of the Ru/LaFeO3 catalyst are significantly different. The signals for b-CO32−, CO(g), and image file: d5sc05708f-t4.tif are gradually enhanced with continuous light irradiation (Fig. 4e). Meanwhile, a distinct signal of image file: d5sc05708f-t5.tif at 1370 cm−1 appears,17 indicating further dehydrogenation of image file: d5sc05708f-t6.tif. This intermediate is subsequently converted to CO and H2 through additional dehydrogenation steps. These findings suggest that light irradiation significantly boosts activation and dissociation of CH4 and CO2, aligning well with the improved photothermal DRM activity.

Finally, DFT calculations were performed to theoretically validate the activation and transformation processes of CH4 and CO2 on Ru/LaFeO3 (Fig. 5). Initially, CH4 adsorbs onto the Ru site with an adsorption energy of 0.7 eV (step I). Subsequently, the C–H bond undergoes cleavage, yielding image file: d5sc05708f-t7.tif and H* species (step II). After that, a deep dehydrogenation of image file: d5sc05708f-t8.tif proceeds, leading to the formation of C* and the stepwise release of all H* (steps III–VI), forming two H2 molecules at the Ru site (steps IV–VII). Concurrently, migration of LO on the Ru/LaFeO3 surface proceeds to generate reactive O* and OV (steps IX–XI). This oxygen transfer process encounters the highest energy barrier of 1.47 eV in the transition state (TS), identifying it as the rate-determining step. The migrated O* then reacts with C* to form CO*. Meanwhile, the generated OV facilitates the adsorption of CO2 with an adsorption energy of −0.76 eV (step XII). After adsorption, CO2 dissociates with an energy barrier of −0.29 eV to produce CO* (step XIII) and releases O atoms that replenish the OVs. Finally, CO* desorbs from the catalyst surface, completing the CO generation process (steps XIV and XV). These results highlight the high effectiveness of Ru/LaFeO3 in activation and conversion of CH4 and CO2, explaining its excellent performance in light-driven DRM.


image file: d5sc05708f-f5.tif
Fig. 5 Relative energy diagrams of CH4 and CO2 conversion on Ru/LaFeO3 simulated by DFT calculation and some corresponding structures of the key reaction intermediates during the CH4 dissociation and CO2 conversion processes.

Based on both experimental findings and theoretical results, a light-driven DRM mechanism over the Ru/LaFeO3 catalyst is proposed (Fig. 6). Upon light irradiation, photoexcited electrons migrate directionally to the Ru NPs, as well as the La and Fe sites, while photogenerated holes accumulated on the surface LOs (step I). Following the synergistic activation of light-induced heat and charge carriers, CH4 molecules undergo cleavage at Ru sites and produce C* and H* species via sequential dehydrogenation (step II). The accumulated H* species are reduced to H2 at the electron-rich Ru sites. Simultaneously, photoexcited electrons weaken the Fe–O or La–O bonds, promoting the migration of LOs and the generation of OVs. The migrated LOs combine with C* to form CO* (step II), while the CO2 molecules adsorb on the OVs-rich LaFeO3 surface and dissociate in the presence of photoexcited electrons into CO* and provides O to refill the OVs (step III). Finally, the adsorbed CO* desorbs from the catalyst surface as CO gas, completing the catalytic cycle (step IV).


image file: d5sc05708f-f6.tif
Fig. 6 Illustration of the proposed photothermal DRM mechanism over Ru/LaFeO3.

Conclusions

In summary, we have developed Ru/LaFeO3 as an efficient photothermal catalyst that exhibits exceptional light-driven DRM performance. This system achieves a methane TOF of 0.64 s−1 and a LTCEE of 16.3%. The high catalytic activity is attributed to the synergistic modulation of the catalyst's electronic structure induced by Ru loading and light excitation, in which electrons accumulate at the Ru, La, and Fe sites, and holes gather at lattice oxygens (LOs). This charge redistribution weakens the La/Fe–O bonds and promotes LO migration, enabling efficient CH4 activation and oxidation at Ru sites, while in situ generating OVs. These OVs reduce the activation energy barrier of CO2 dissociation, promoting its cleavage into CO and replenishing OVs through O* species, thereby completing the catalytic cycle. This work deciphers the feasibility of activating LOs through the dual action of cocatalysts and photoelectrons in modulating the catalyst's electronic structure, offering a strategic blueprint to fabricate high-performance catalysts toward solar-powered DRM.

Author contributions

J. Li and S. Wang conceived the project. J. Li designed and carried out all the experiments. B. Su, J. Li and X. Hao provided support with the theoretical concepts and experiment design software operation. J. Zhao assisted with theoretical calculations. J. Li prepared the initial manuscript. Z. Ding, M. Huang, M. Yang, Z. Lan and S. Wang revised the manuscript. All the authors discussed the results and provided feedback on the manuscript.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The datasets supporting this article have been uploaded as part of the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5sc05708f.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (22372035, 22472026, and 22302039) and the 111 Project (D16008). The authors acknowledge the facility resources from the Electron Microscopy Center of Fuzhou University.

Notes and references

  1. Q. Zhu, H. Zhou, L. Wang, L. Wang, C. Wang, H. Wang, W. Fang, M. He, Q. Wu and F.-S. Xiao, Nat. Catal., 2022, 5, 1030–1037 CrossRef CAS.
  2. X. Li, C. Wang and J. Tang, Nat. Rev. Mater., 2022, 7, 617–632 CrossRef CAS.
  3. Q. Hao, Z. Li, Y. Zhu, Y. Shi, M. Huo, H. Yuan, S. Ouyang and T. Zhang, Adv. Funct. Mater., 2024, 34, 2403848 CrossRef CAS.
  4. S. Chu and Q. Wang, Front. Energy, 2024, 18, 717–726 CrossRef CAS.
  5. M. Zhou, H. Wang, R. Liu, Z. Liu, X. Xiao, W. Li, C. Gao, Z. Lu, Z. Jiang, W. Shi and Y. Xiong, Angew. Chem., Int. Ed., 2024, 63, e202407468 CrossRef CAS.
  6. J. Wang, R. Li, D. Zeng, W. Wang, Y. Zhang, L. Zhang and W. Wang, Chem. Eng. J., 2023, 452, 139505 CrossRef CAS.
  7. B. Su, Y. Kong, S. Wang, S. Zuo, W. Lin, Y. Fang, Y. Hou, G. Zhang, H. Zhang and X. Wang, J. Am. Chem. Soc., 2023, 145, 27415–27423 CrossRef CAS.
  8. X. Xu, J. Lu, B. Su, J. Chen, X. Chen and S. Wang, Acta Phys.–Chim. Sin., 2025, 41, 100153 CrossRef.
  9. Q. Huang, J. Cai, F. Wei, Y. Fan, Z. Liang, K. Liu, X. F. Lu, Z. Ding and S. Wang, J. Mater. Chem. A, 2024, 12, 21334–21340 RSC.
  10. B. Su, S. Wang, W. Xing, K. Liu, S.-F. Hung, X. Chen, Y. Fang, G. Zhang, H. Zhang and X. Wang, Angew. Chem., Int. Ed., 2025, 64, e02505453 Search PubMed.
  11. H. Wang, G. Cui, H. Lu, Z. Li, L. Wang, H. Meng, J. Li, H. Yan, Y. Yang and M. Wei, Nat. Commun., 2024, 15, 3765 CrossRef CAS PubMed.
  12. J. Sasson Bitters, T. He, E. Nestler, S. D. Senanayake, J. G. Chen and C. Zhang, J. Energy Chem., 2022, 68, 124–142 CrossRef CAS.
  13. J. S. Bitters, T. N. He, E. Nestler, S. D. Senanayake, J. G. G. Chen and C. Zhang, J. Energy Chem., 2022, 68, 124–142 CrossRef.
  14. H. Xiong, Y. Dong, C. Hu, Y. Chen, H. Liu, R. Long, T. Kong and Y. Xiong, J. Am. Chem. Soc., 2024, 146, 9465–9475 CrossRef CAS PubMed.
  15. Y. Yao, B. Li, X. Gao, Y. Yang, J. Yu, J. Lei, Q. Li, X. Meng, L. Chen and D. Xu, Adv. Mater., 2023, 35, e2303654 CrossRef PubMed.
  16. Y. Tang, Y. Li, W. Bao, W. Yan, J. Zhang, Y. Huang, H. Li, Z. Wang, M. Liu and F. Yu, Appl. Catal., B, 2023, 338, 123054 CrossRef CAS.
  17. Q. Li, H. Wang, M. Zhang, G. Li, J. Chen and H. Jia, Angew. Chem., Int. Ed., 2023, 135, e202300129 CrossRef.
  18. Y. Yang, Z. Chai, X. Qin, Z. Zhang, A. Muhetaer, C. Wang, H. Huang, C. Yang, D. Ma, Q. Li and D. Xu, Angew. Chem., Int. Ed., 2022, 61, e202200567 CrossRef CAS PubMed.
  19. Z. Q. Rao, Y. H. Cao, Z. A. Huang, Z. H. Yin, W. C. Wan, M. Z. Ma, Y. X. Wu, J. B. Wang, G. D. Yang, Y. Cui, Z. M. Gong and Y. Zhou, ACS Catal., 2021, 11, 4730–4738 CrossRef CAS.
  20. J. Li, J. Zhao, S. Wang, K.-S. Peng, B. Su, K. Liu, S.-F. Hung, M. Huang, G. Zhang, H. Zhang and X. Wang, J. Am. Chem. Soc., 2025, 147, 14705–14714 CrossRef CAS PubMed.
  21. S. Shoji, X. B. Peng, A. Yamaguchi, R. Watanabe, C. Fukuhara, Y. Cho, T. Yamamoto, S. Matsumura, M. W. Yu, S. Ishii, T. Fujita, H. Abe and M. Miyauchi, Nat. Catal., 2020, 3, 148–153 CrossRef CAS.
  22. F. Wei, J. Zhao, Y.-C. Liu, Y.-H. Hsu, S.-F. Hung, J. Fu, K. Liu, W. Lin, Z. Yu, L. Tan, X. F. Lu, C. Feng, H. Zhang and S. Wang, Nat. Commun., 2025, 16, 6586 CrossRef CAS.
  23. Y. Liu, W. Xue, X. Liu, F. Wei, X. Lin, X. F. Lu, W. Lin, Y. Hou, G. Zhang and S. Wang, Small, 2024, 20, 2402004 CrossRef CAS.
  24. S. Chen, L. Zeng, H. Tian, X. Li and J. Gong, ACS Catal., 2017, 7, 3548–3559 CrossRef CAS.
  25. C. He, S. Wu, Q. Li, M. Li, J. Li, L. Wang and J. Zhang, Chem, 2023, 9, 3224–3244 CAS.
  26. Y. Li, Z. Li, N. Wang, Y. Zha, K. Zheng, Y. Xu, B. Liu and X. Liu, Chem Catal., 2025, 5, 101189 CAS.
  27. J. He, T. Wang, X. Bi, Y. Tian, C. Huang, W. Xu, Y. Hu, Z. Wang, B. Jiang, Y. Gao, Y. Zhu and X. Wang, Nat. Commun., 2024, 15, 5422 CrossRef CAS PubMed.
  28. Z. Shi, H. Li, L. Zhang and Y. Cao, J. Power Sources, 2022, 519, 230738 CrossRef CAS.
  29. X. Zhang, C. Pei, X. Chang, S. Chen, R. Liu, Z.-J. Zhao, R. Mu and J. Gong, J. Am. Chem. Soc., 2020, 142, 11540–11549 CrossRef CAS.
  30. R. Maity, M. S. Sheikh, A. Dutta and T. P. Sinha, J. Electron. Mater., 2019, 48, 4856–4865 CrossRef CAS.
  31. A. G. Bhavani, W. Y. Kim and J. S. Lee, ACS Catal., 2013, 3, 1537–1544 CrossRef CAS.
  32. B. Yang, K. Liu, Y. Ma, J.-J. Ma, Y.-Y. Chen, M. Huang, C. Yang, Y. Hou, S.-F. Hung, J. C. Yu, J. Zhang and X. Wang, Angew. Chem., Int. Ed., 2024, 63, e202410394 CAS.
  33. R. Song, G. Zhao, J. M. Restrepo-Flórez, C. J. Viasus Pérez, Z. Chen, C. Ai, A. Wang, D. Jing, A. A. Tountas, J. Guo, C. Mao, C. Li, J. Shen, G. Cai, C. Qiu, J. Ye, Y. Fu, C. T. Maravelias, L. Wang, J. Sun, Y.-F. Xu, Z. Li, J. Y. Y. Loh, N. T. Nguyen, L. He, X. Zhang and G. A. Ozin, Nat. Energy, 2024, 9, 750–760 CrossRef CAS.
  34. A. Shahnazi and S. Firoozi, J. CO2 Util., 2021, 45, 101455 CrossRef CAS.
  35. Q. Lin, J. Zhao, P. Zhang, S. Wang, Y. Wang, Z. Zhang, N. Wen, Z. Ding, R. Yuan, X. Wang and J. Long, Carbon Energy, 2024, 6, e435 CrossRef CAS.
  36. Y. Cao, R. Guo, M. Ma, Z. Huang and Y. Zhou, Acta Phys.–Chim. Sin., 2024, 40, 2303029 CrossRef.
  37. L. Zhou, J. M. P. Martirez, J. Finzel, C. Zhang, D. F. Swearer, S. Tian, H. Robatjazi, M. Lou, L. Dong, L. Henderson, P. Christopher, E. A. Carter, P. Nordlander and N. J. Halas, Nat. Energy, 2020, 5, 61–70 CrossRef CAS.
  38. J. Zhang, K. Xie, Y. Jiang, M. Li, X. Tan, Y. Yang, X. Zhao, L. Wang, Y. Wang, X. Wang, Y. Zhu, H. Chen, M. Wu, H. Sun and S. Wang, ACS Catal., 2023, 13, 10855–10865 CrossRef CAS.
  39. J. Zhang, Y. Li, J. Sun, H. Chen, Y. Zhu, X. Zhao, L.-C. Zhang, S. Wang, H. Zhang, X. Duan, L. Shi, S. Zhang, P. Zhang, G. Shao, M. Wu, S. Wang and H. Sun, Appl. Catal., B, 2022, 309, 121263 CrossRef CAS.
  40. H. Liu, X. Meng, T. D. Dao, H. Zhang, P. Li, K. Chang, T. Wang, M. Li, T. Nagao and J. Ye, Angew. Chem., Int. Ed., 2015, 54, 11545–11549 CrossRef CAS PubMed.
  41. Y. Wang, Y. Liu, L. Tan, X. Lin, Y. Fang, X. F. Lu, Y. Hou, G. Zhang and S. Wang, J. Mater. Chem. A, 2023, 11, 26804–26811 RSC.
  42. B. Su, M. Zheng, W. Lin, X. F. Lu, D. Luan, S. Wang and X. W. Lou, Adv. Energy Mater., 2023, 13, 2203290 CrossRef CAS.
  43. J. Cai, X. Li, B. Su, B. Guo, X. Lin, W. Xing, X. F. Lu and S. Wang, J. Mater. Sci. Technol., 2025, 234, 82–89 CrossRef CAS.
  44. C. Zhang, X. Tao, W. Jiang, J. Guo, P. Zhang, C. Li and R. Li, Acta Phys.–Chim. Sin., 2024, 40, 2303034 CrossRef.
  45. E. Lu, J. Tao, C. Yang, Y. Hou, J. Zhang, X. Wang and X. Fu, Acta Phys.–Chim. Sin., 2023, 39, 2211029 Search PubMed.
  46. B. Li, W. Wang, J. Zhao, Z. Wang, B. Su, Y. Hou, Z. Ding, W. J. Ong and S. Wang, J. Mater. Chem. A, 2021, 9, 10270–10276 RSC.
  47. M. Liu, G. Zhang, X. Liang, Z. Pan, D. Zheng, S. Wang, Z. Yu, Y. Hou and X. Wang, Angew. Chem., Int. Ed., 2023, 62, e202304694 CrossRef CAS PubMed.
  48. X. Xu, C. Shao, J. Zhang, Z. Wang and K. Dai, Acta Phys.–Chim. Sin., 2024, 40, 2309031 CrossRef.
  49. J. Su, J. Zhang, S. Chai, Y. Wang, S. Wang and Y. Fang, Acta Phys.–Chim. Sin., 2024, 40, 2408012 CrossRef.
  50. F. Liu, J. Deng, B. Su, K.-S. Peng, K. Liu, X. Lin, S.-F. Hung, X. Chen, X. F. Lu, Y. Fang, G. Zhang and S. Wang, ACS Catal., 2025, 15, 1018–1026 CrossRef CAS.
  51. Y. Zou, H. Chen, Y. Hou, W. Xing, Z. Pan, O. Savateev, M. Anpo and G. Zhang, Adv. Funct. Mater., 2025, e16479 CrossRef.
  52. L. Huang, B. Li, B. Su, Z. Xiong, C. Zhang, Y. Hou, Z. Ding and S. Wang, J. Mater. Chem. A, 2020, 8, 7177–7183 RSC.
  53. C. Guo, Y. Cui, W. Zhang, X. Du, X. Peng, Y. Yu, J. Li, Y. Wu, Y. Huang, T. Kong and Y. Xiong, Adv. Mater., 2025, 37, 2500928 CrossRef CAS.
  54. Y.-F. Wang, M.-Y. Qi, M. Conte, Z.-R. Tang and Y.-J. Xu, Angew. Chem., Int. Ed., 2024, 63, e202407791 CrossRef CAS PubMed.
  55. W. Zhang, C. Fu, J. Low, D. Duan, J. Ma, W. Jiang, Y. Chen, H. Liu, Z. Qi, R. Long, Y. Yao, X. Li, H. Zhang, Z. Liu, J. Yang, Z. Zou and Y. Xiong, Nat. Commun., 2022, 13, 2806 CrossRef CAS PubMed.
  56. W. Jiang, J. Low, K. Mao, D. Duan, S. Chen, W. Liu, C.-W. Pao, J. Ma, S. Sang, C. Shu, X. Zhan, Z. Qi, H. Zhang, Z. Liu, X. Wu, R. Long, L. Song and Y. Xiong, J. Am. Chem. Soc., 2021, 143, 269–278 CrossRef CAS PubMed.
  57. W. Zhou, B. H. Wang, L. Tang, L. Chen, J. K. Guo, J. B. Pan, B. Lei, B. Hu, Z. J. Bai, M. Tulu, Z. X. Li, X. Wang, C. T. Au and S. F. Yin, Adv. Funct. Mater., 2023, 33, 2214068 CrossRef CAS.
  58. S. Luo, H. Song, F. Ichihara, M. Oshikiri, W. Lu, D.-M. Tang, S. Li, Y. Li, Y. Li, P. Davin, T. Kako, H. Lin and J. Ye, J. Am. Chem. Soc., 2023, 145, 20530–20538 CrossRef CAS PubMed.
  59. S. Singh, R. Verma, N. Kaul, J. Sa, A. Punjal, S. Prabhu and V. Polshettiwar, Nat. Commun., 2023, 14, 2551 CrossRef CAS PubMed.
  60. Q. Cheng, X. Yao, L. Ou, Z. Hu, L. Zheng, G. Li, N. Morlanes, J. L. Cerrillo, P. Castaño, X. Li, J. Gascon and Y. Han, J. Am. Chem. Soc., 2023, 145, 25109–25119 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.