Open Access Article
Qingyun Zhana,
Xiaowei Mu*b,
Yuxiang Konga,
Zhenlu Lia,
Le Liua,
Yumeng Qiana,
Shuyan Song
*b and
Lu Li
*a
aState Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, People's Republic of China. E-mail: luli@jlu.edu.cn
bState Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China. E-mail: muxw@ciac.ac.cn; songsy@ciac.ac.cn
First published on 6th October 2025
Selective methane conversion is a promising low-carbon technology, yet developing catalysts capable of effectively activating inert C–H bonds under mild conditions remains challenging. Here, we designed Fe2+–□–Ti4+ dual-metal-sites on the high-activity {012} facets of defective ilmenite (Fe1−xTiO3−x-NS). Mechanistic studies revealed that Fe2+–□–Ti4+ can be excited to form a long-lived Fe3+–□–Ti3+ active state, while surface lattice oxygen stabilizes the reaction intermediates during multistep elementary reactions. This bimetal–oxygen synergistic strategy significantly reduces the activation barrier for C–H bond cleavage to just 0.15 eV, fully blocks the over-dehydrogenation of methyl intermediates, and facilitates C–C bond formation, thereby achieving a favorable non-oxidative methane conversion rate that even surpasses noble metal-supported photocatalysts, with nearly 100% C2+ selectivity. This bimetallic-center construction strategy not only provides an efficient and economical pathway for methane conversion but also expands the boundaries of traditional photocatalysts, exhibiting high catalytic activity and product selectivity.
Photocatalytic non-oxidative coupling of methane (NOCM) is an emerging green technology that utilizes only solar energy and methane as inputs. By harnessing photon-induced high-energy charge carriers to pre-activate the C–H bonds in CH4, this approach reduces the reaction barrier and enables thermodynamically unfavourable reactions under mild conditions.6,9–12 Previous studies have demonstrated that quantum dot catalysts (SiO2,13–15 Al2O3,13–15 and MgO16) and zeolite-based catalysts (H-MOR,17 MCM-41,18 Zn+-ZSM-5,19 and Nb-TS20) can activate methane under light irradiation. However, despite considerable research efforts, the improvement in CH4 conversion has been limited. Noble metals such as Au,11,21 Pt,22,23 Pd,11,24–26 and Ag27,28 have been shown to significantly enhance electron transfer, facilitating C–H bond cleavage and accelerating methane conversion, but their high cost remains a significant drawback.29–31 In recent years, constructing frustrated Lewis pairs (FLPs) on semiconductors has emerged as an effective strategy for C–H activation.32–34 Nevertheless, this method relies on abundant surface hydroxyl groups, which compromises the long-term stability of the catalysts and requires frequent regeneration treatments.
Ilmenite (FeTiO3) is a titanate mineral with a bandgap of 2.5–2.9 eV and antiferromagnetic semiconducting properties, commonly found in the Earth's crust.35 Its structure consists of alternating FeO6 and TiO6 octahedra sharing faces (Fig. S1), with the metal ion valences represented as Fe2+Ti4+O3.36 Strong interactions between metals and between metals and ligands facilitate electron transfer,37–39 and studies suggest that iron-titanium oxides can effectively harness solar energy for photocatalysis.40 Additionally, calculations by Ribeiro41 indicate that the Fe valence band maximum (VBM) has more occupied states, while the contribution of oxygen to the valence band near the surface is minimal, reducing excessive oxidation of reactants.25 Thus, FeTiO3 shows significant potential for photocatalytic non-oxidative methane conversion.
Here, we report a defect-state ilmenite catalyst, Fe1−xTiO3−x-NS, featuring high-activity {012} facets. Surface oxygen vacancies create Fe2+–□–Ti4+ dual-metal sites, which can be excited to form a long-lived Fe3+–□–Ti3+. This structure reduces the activation barrier for C–H cleavage and the surface lattice oxygen can stabilize the reaction intermediates. Under light irradiation, the methane conversion rate reached 5456.2 μmol g−1 h−1. This work provides guidance for the rational design and construction of photocatalysts for efficient methane conversion without additional heating.
Following hydrothermal treatment, scanning electron microscopy (SEM) demonstrated a morphological transformation from particulate to a layered nanosheet structure (Fe1−xTiO3−x-NS) (Fig. 1d), designed to reduce steric hindrance during methane adsorption. HR-TEM images of the layered catalyst showed an interlayer spacing of 7.10–7.13 Å. The TEM image shows that the sample exhibits a nanosheet-like structure. The HRTEM image of the selected area displays clear lattice fringes with an interplanar spacing of 0.374 nm, which matches well with the {012} planes of FeTiO3. The measured lattice angles of approximately 86° and 94° are also consistent with the features of the {012} facet. Fast Fourier Transform (FFT) analysis of the same region clearly identifies diffraction spots corresponding to the (012), (1,0,−2), and (−1,−1,0) planes, with a zone axis of [2,−2,1], indicating that the exposed surface of the nanosheets can be assigned to the {012} facet. EDX mapping confirmed the uniform elemental distribution of Fe, Ti, and O within the samples (Fig. 1e). Inductively coupled plasma (ICP) analysis clearly revealed a reduced Fe/Ti molar ratio in Fe1−xTiO3−x-NS compared to FeTiO3−x-NP (Fig. S2), which can be ascribed to the partial dissolution of iron species during the hydrothermal treatment. X-ray diffraction (XRD) patterns showed that both materials maintain the pure FeTiO3 phase, although Fe1−xTiO3−x-NS exhibited decreased crystallinity and weaker diffraction peaks due to the thin layered structure formed in the hydrothermal step (Fig. 1f).
Theoretical analysis showed that the {012} facet has a lower work function (4.02 eV) compared to the {104} facet (4.89 eV), enabling easier electron transfer and potentially promoting more effective methane activation. (Fig. S3). Ultraviolet photoelectron spectroscopy (UPS) measurements confirmed that the work function of FeTiO3−x-NP was 5.56 eV, while that of Fe1−xTiO3−x-NS was 4.41 eV (Fig. S4), consistent with theoretical calculations. By combining UPS, X-ray photoelectron spectroscopy (XPS) valence band spectra, and UV-visible spectroscopy, the band structure of Fe1−xTiO3−x-NS was determined, showing an excellent match with the redox potential required for methane activation (Fig. S5).
The chemical states of the elements in the catalysts were analysed using XPS. From the Fe 2p and Ti 2p spectra, it was evidenced that Fe exists predominantly as Fe2+ with a minor presence of Fe3+ in both FeTiO3−x-NP and Fe1−xTiO3−x-NS, while Ti is entirely present as Ti4+ (Fig. S6). In the O 1s spectrum of FeTiO3−x-NP (Fig. 1g), peaks at 532.7 eV, 531.1 eV, and 528.9 eV were attributed to surface-adsorbed oxygen (OC), oxygen vacancies (VO), and lattice oxygen (OL),42 respectively. A marked increase in the oxygen vacancy peak was observed in Fe1−xTiO3−x-NS.
Theoretical calculations provided further insight, indicating that the formation energy of oxygen vacancies on the {012} facet is 1.18 eV lower than that on the {104} facet, making oxygen vacancies more likely to form on the {012} facet (Fig. S7). Similar trends were observed for Fe vacancies, which also formed more readily on the {012} facet. Together, these experimental and theoretical findings demonstrate that FeTiO3−x-NP contains abundant oxygen vacancies, while Fe1−xTiO3−x-NS, with exposed {012} facets after hydrothermal treatment, features a higher concentration of both oxygen and iron vacancies.
Based on the structural characteristics summarized above, we constructed surface models of the {104} and {012} facets using density functional theory (DFT). The models containing oxygen vacancies were denoted as {104}-VO and {012}-VO, while the {012} facet containing both oxygen and Fe vacancies was denoted as {012}-VO–VFe. In the optimized models, we measured the distances between Fe and Ti atoms across the bridging oxygen vacancy. This bridging oxygen vacancy between Fe and Ti atoms defined the Fe–Ti dual-metal sites (designated as Fe2+–□–Ti4+). On the {104} facet without oxygen vacancies, the Fe–Ti distance was 3.89 Å (Fig. S8a); introducing oxygen vacancies ({104}-VO) slightly increased this distance to 3.96 Å (Fig. 1h). On the {012} facet without oxygen vacancies, the Fe–Ti distance was 3.46 Å (Fig. S8b); after introducing oxygen vacancies ({012}-VO), the distance decreased to 2.97 Å (Fig. S8c); further introducing Fe vacancies on this facet ({012}-VO–VFe) reduced the distance to 2.89 Å (Fig. 1i). The distance between active sites often correlates directly with catalytic activity, as optimal spacing can accelerate reaction rates. These results reveal substantial structural differences in Fe–Ti configurations across the two facets, which may lead to significant variations in methane adsorption and activation.
We investigated the interaction between the catalyst and CH4 using temperature-programmed desorption of methane (CH4-TPD). The desorption peaks around 375 K and 550 K corresponded to the weak and strong chemical adsorption of CH4 on the catalyst surface, respectively.43 Notably, Fe1−xTiO3−x-NS exhibited significantly greater strong adsorption of CH4 than FeTiO3−x-NP (Fig. 1j), with complete desorption requiring a temperature of 840 K, suggesting stronger CH4 binding to the Fe1−xTiO3−x-NS surface. This enhanced strong adsorption can be attributed to Fe1−xTiO3−x-NS being almost entirely composed of the highly active {012} facet, which has a higher concentration of oxygen and iron vacancies than the {104} facet.
The non-oxidative coupling of methane (NOCM) is thermodynamically unfavourable due to its positive Gibbs free energy. Fig. 2b illustrates the temperature dependence of the NOCM reaction, where the dashed line represents the maximum equilibrium conversion achievable by thermocatalysis, setting a limit above which no thermocatalyst can convert methane. Notably, below 600 K, the theoretical conversion of thermocatalytic methane remains negligible. In contrast, Fe1−xTiO3−x-NS achieves a CH4 conversion rate far exceeding this thermodynamic limit under illumination at near-room temperature. It is noteworthy that in the lower temperature range (<393 K), the photocatalytic conversion rate significantly increases with rising temperature. However, above 393 K, this temperature effect gradually diminishes, indicating that in a low-temperature gas–solid phase system, heat-limited migration of intermediates and product desorption restrict the turnover frequency of active sites. As the temperature rises, these heat-driven processes accelerate. Once the temperature exceeds 393 K, the rate-determining step shifts to the C–H bond activation process driven by light, rendering further temperature increases ineffective in enhancing the conversion rate.
Fig. 2c compares the performance of Fe1−xTiO3−x-NS with a range of reported non-noble metal-based photocatalysts for NOCM. Our catalyst demonstrates a remarkably higher ethane yield, surpassing that of other non-noble metal photocatalysts and even outperforming noble-metal-supported photocatalysts such as Au/ZnO and Pd1/TiO2 (Table S4). This exceptional photocatalytic efficiency positions Fe1−xTiO3−x-NS as a promising and low-cost candidate for advanced NOCM applications.
We also compared the product distribution over time under light irradiation for both catalysts (Fig. 2d). Apart from C2H6, H2, and C3H8, no other products were observed with extended irradiation. The dashed line represents the theoretical hydrogen yield. Fe1−xTiO3−x-NS produced hydrogen closely matching the theoretical value, while FeTiO3−x-NP produced significantly more hydrogen than expected due to side reactions involving excessive dehydrogenation of CH4. Raman spectroscopy on the samples after seven reaction cycles (Fig. S12) revealed distinct D and G bands at 1345 cm−1 and 1593 cm−1 over FeTiO3−x-NP, respectively, indicating carbon deposition. In contrast, Fe1−xTiO3−x-NS showed no significant D or G bands, suggesting excellent resistance to carbon deposition.
Next, we conducted cyclic testing on both catalysts, with each cycle lasting two hours. Notably, Fe1−xTiO3−x-NS maintained consistent activity across 20 cycles, with C2+ selectivity remaining above 99% (Fig. 2e), demonstrating not only excellent resistance to sintering but also eliminating the need for regeneration treatments typically required for conventional non-noble metal catalysts. In comparison, the ethane production rate on FeTiO3−x-NP dropped to one-fifth of its initial rate after seven cycles, along with a decline in selectivity (Fig. S13). Overall, Fe1−xTiO3−x-NS outperforms FeTiO3−x-NP in all aspects (Fig. S14). Post-reaction analysis (Fig. S15 and 16) confirmed that both photocatalysts retained their structural integrity, further underscoring their stability.
XPS revealed the oxidation state changes of Fe and Ti during photoexcitation. As shown in Fig. 3b and S18, under illumination, the Fe 2p binding energy in Fe1−xTiO3−x-NS shifted positively by 0.3 eV, indicating the accumulation of photogenerated holes at Fe sites by the formation of Fe3+. Concurrently, the Ti 2p binding energy shifted negatively by 0.2 eV, suggesting an accumulation of photogenerated electrons at Ti sites, resulting in the formation of Ti3+. These findings align with theoretical predictions. Additionally, the concentration of oxygen vacancies increased under illumination (Fig. 3c), indicating the generation of more Fe–Ti dual-metal sites by photoexcitation. In comparison, FeTiO3−x-NP exhibited similar but much smaller changes in Fe, Ti, and oxygen vacancies upon illumination than Fe1−xTiO3−x-NS (Fig. S19).
To further analyse electron transfer in Fe1−xTiO3−x-NS during the photocatalytic NOCM reaction, we used in situ electron paramagnetic resonance (EPR). The EPR signal with g = 2.13 measured at room temperature is attributed to Fe3+.47,48 The signals with g = 2.003 and g = 1.96, measured at 77 K, are assigned to oxygen vacancies (VO) and Ti3+, respectively.23,49,50 As shown in Fig. 3d, after 10 minutes of vacuum illumination, the signals for Fe3+ and oxygen vacancies (VO) increased significantly, and a new signal corresponding to Ti3+ emerged, fully consistent with the in situ XPS results. When methane was introduced, both the Fe3+ and Ti3+ signals decreased markedly, suggesting that the photogenerated Fe3+ and Ti3+ species actively participate in methane activation, facilitating C–H bond cleavage through charge transfer. Notably, after the illumination was removed, the generated Fe3+ species remain stable (Fig. S20), indicating that these dual-metal centers effectively inhibit charge transfer from Ti3+ to Fe3+. This feature enables the catalyst surface to capture and retain a substantial amount of photogenerated charge, forming a long-lived Fe3+–□–Ti3+ active state.
The “Pre-Exposure experiment” further confirmed that Fe1−xTiO3−x-NS underwent a photoinduced dynamic transition under illumination from a “resting state” to a long-lived “active state”. As shown in Fig. 3e, when the photocatalytic NOCM reaction began immediately after methane introduction, an induction period was observed, during which the catalytic activity of Fe1−xTiO3−x-NS increased progressively with extended light exposure, reaching a maximum after approximately 20 minutes before stabilizing (Fig. 3e, bottom). When the catalyst was pre-exposed to methane for 30 minutes before starting the photocatalytic reaction, the induction period persisted (Fig. 3e, middle), indicating that it was not due to CH4 diffusion or adsorption equilibrium. Conversely, when the catalyst was pre-exposed to light under vacuum for 30 minutes before methane introduction, the NOCM reaction began at the optimal rate without any induction period (Fig. 3e, top). These results collectively underscore that the formation of a long-lived Fe3+–□–Ti3+ active state is crucial for methane activation, enabling efficient and sustained photocatalytic performance of Fe1−xTiO3−x-NS.
To gain insights into the efficiency of photogenerated charge separation, we conducted photoelectrochemical measurements to investigate interfacial charge dynamics in Ar and CH4 atmospheres. In the absence of CH4, Fe1−xTiO3−x-NS exhibited a higher photocurrent than FeTiO3−x-NP (Fig. 3f), indicating more efficient charge separation and transfer in line with the theoretical calculations and XPS results. When CH4 was introduced, the photocurrent increased significantly, especially for Fe1−xTiO3−x-NS. This surge in photocurrent is attributed to the consumption of photogenerated holes by CH4, which enhances the transfer of photogenerated electrons to the electrode. These results demonstrate that Fe1−xTiO3−x-NS not only facilitates more efficient charge separation but also exhibits a stronger ability to activate methane through enhanced electron transfer dynamics.
As shown in Fig. 4b, we used DFT to calculate the adsorption energies of CH4 on the {012}-VO–VFe, {104}-VO, {012}, and {104} facets. The results indicate that on all tested facets, methane preferentially adsorbs at Fe sites. This preference arises from the interaction between the d orbitals of Fe and the σ orbitals of CH4, forming metal–molecule interactions that strengthen CH4 adsorption on Fe. The optimized adsorption configurations show that the adsorption energy (Eads) of CH4 on {012}-VO–VFe (−0.58 eV) is significantly lower than on {104}-VO (−0.42 eV), indicating stronger CH4 adsorption on the {012}-VO–VFe facet, consistent with the CH4-TPD results. Stronger adsorption also led to a lower reaction order for methane on Fe1−xTiO3−x-NS (0.44) compared to FeTiO3−x-NP (0.51) (Fig. S21).
Analysis of the optimal adsorption configurations indicates that the Fe–□–Ti active sites on the {012}-VO–VFe facet exhibit stronger polarization ability for CH4 adsorption compared to the {104}-VO facet. As shown in Fig. S22, on the {012}-VO–VFe facet, the distance between the Fe site and the carbon atom of methane is 2.29 Å, and the distance between the Ti site and the hydrogen species is 2.91 Å. The C–H bond of CH4 adsorbed on the catalyst surface is elongated from 1.099 Å in free methane to 1.17 Å. In contrast, on the {104}-VO facet, the distance between the Fe site and the carbon atom of methane is 2.50 Å, and the distance between the Ti site and the hydrogen species is 3.53 Å, with the C–H bond being elongated to 1.13 Å. We further analysed the projected density of states (PDOS) and projected crystal orbital Hamilton populations (COHP) after CH4 adsorption (Fig. 4c and S23). The negative integrated COHP (–ICOHP) values for the Fe–C (1.54 eV) and Ti–H (0.90 eV) interactions on the {012}-VO–VFe facet are more positive than the Fe–C (1.22 eV) and Ti–H (0.17 eV) interactions on the {104}-VO facet. All these computational results clearly demonstrate that the Fe–□–Ti active sites on the {012}-VO–VFe facet form stronger interactions with methane, which is beneficial for C–H bond cleavage.
To gain a comprehensive understanding of the coupling process in NOCM, we employed DFT to study the reaction pathway of CH4 on the catalyst surface (Fig. 4d). In the gas phase, C–H bond cleavage requires a large amount of energy (∼473 kJ mol−1). However, on the {012}-VO–VFe facet, the dissociation of the C–H bond required only 0.15 eV (∼14.46 kJ mol−1), significantly lowering the high activation energy barrier typically associated with C–H bond activation. In contrast, on the {104}-VO facet, the activation energy for the first C–H bond cleavage was much higher, at 0.52 eV, over three times that of the {012}-VO–VFe facet. Additionally, the energy required for methyl migration from the Fe site to the adjacent lattice oxygen atom to form the methoxy species intermediate on the {012}-VO–VFe surface was 0.76 eV, compared to 1.32 eV on the {104}-VO surface, indicating that methyl migration proceeds more smoothly on the {012}-VO–VFe facet.
Taking spatial effects into account, the activation energy for the second adsorbed CH4 molecule on the {012}-VO–VFe facet was 1.20 eV, higher than that for the first C–H cleavage, but still lower than the 1.44 eV required on the {104}-VO facet, further demonstrating the superior ability of the {012}-VO–VFe facet for C–H bond activation. For ethane formation in the NOCM process, the energy required on the {012}-VO–VFe facet was 0.53 eV, compared to 0.81 eV on the {104}-VO facet, confirming that methane-to-ethane conversion on the {012}-VO–VFe facet is more energetically favourable. In terms of hydrogen gas formation, the energy barrier on the {012}-VO–VFe facet was only 0.1 eV (Fig. S24), with product desorption requiring 0.33 eV, indicating that hydrogen can easily desorb from the surface, promoting the overall forward reaction. Moreover, calculations reveal that the cleavage of the C–H of the second adsorbed CH4 is the rate-determining step in the overall process.
The conversion mechanism of the methyl intermediate is crucial in determining reaction selectivity. We explored the potential reaction pathways of the in situ generated methyl intermediate on the catalyst surface. As shown in Fig. 4e, on the {012}-VO–VFe facet, the energy required for direct desorption of the methyl intermediates into the gas phase was 2.32 eV, while further dehydrogenation required 1.65 eV. In contrast, the migration of the methyl group from the Fe site to the neighbouring lattice oxygen atom required only 0.76 eV. This energetically favourable migration pathway ensures Fe site regeneration, enabling the adsorption and activation of a second methane molecule, leading to nearly exclusive production of coupling products and complete suppression of carbon deposition. On the {104}-VO facet, however, as shown in Fig. S25, the more favourable pathway for the methyl intermediate was further dehydrogenation (0.32 eV) rather than migration (1.32 eV), which leads to significant carbon deposition. These theoretical findings align well with the catalytic testing and Raman spectroscopy results. The FeTiO3−x-NP sample, composed of a mixture of {012}-VO–VFe and {104}-VO facets, produces excessive hydrogen and large amounts of carbon deposition. In contrast, the Fe1−xTiO3−x-NS sample, composed almost entirely of {012}-VO–VFe facets, demonstrates nearly 100% C2+ product selectivity and stoichiometric hydrogen production, with no significant carbon deposition observed, even after extended cycling.
The photocatalytic NOCM reaction occurring on Fe1−xTiO3−x-NS is illustrated in Fig. S26. Initially, Fe1−xTiO3−x-NS absorbs a photon of light with energy greater than or equal to their bandgap, leading to the generation of photogenerated electron–hole pairs. The photogenerated holes are captured by Fe in the valence band, while the photogenerated electrons are captured by Ti in the conduction band, forming a stable, long-lived activated state. Subsequently, methane molecules are firmly adsorbed onto the catalyst surface, establishing stable molecule–site interactions. This interaction significantly enhances the orbital coupling between methane and the active sites, thereby facilitating the interaction of photogenerated charge carriers with methane and enabling efficient C–H bond cleavage. The generated methyl group preferentially migrates to nearby lattice oxygen, forming a stable methoxy intermediate, which releases the active site and reduces the spatial hindrance for the adsorption of a second methane molecule. Finally, the two methyl groups couple to form ethane, while two hydrogen species couple to form hydrogen gas. The products then desorb from the catalyst surface, completing the catalytic cycle. Throughout the reaction, the bimetallic sites and surface lattice oxygen exhibit a synergistic effect that optimizes the pathways for each elementary step. This includes lowering the activation barriers for both C–H bond cleavages, facilitating methyl migration, reducing the energy barrier for C–C bond formation, and suppressing further dehydrogenation of the methyl group. Compared to conventional active centers, this bimetal–oxygen synergistic strategy significantly enhances both catalytic activity and product selectivity.
| This journal is © The Royal Society of Chemistry 2025 |