Hai-Tao
Wan
a,
Chang-Long
Tan
*b,
Ming-Yu
Qi
b,
Yin-Feng
Wang
a,
Zi-Rong
Tang
*ab and
Yi-Jun
Xu
*ab
aCollege of Chemistry, State Key Laboratory of Photocatalysis on Energy and Environment, Fuzhou University, Fuzhou 350116, P. R. China
bInstitute of Fundamental and Frontier Sciences & School of Materials and Energy, University of Electronic Science and Technology of China, Chengdu 611731, P. R. China. E-mail: cl_tan@uestc.edu.cn; zrtang@uestc.edu.cn; yjxu@uestc.edu.cn
First published on 29th October 2025
The photocatalytic conversion of methane (CH4) into high-value multicarbon (C2+) products under ambient conditions provides a highly promising approach for the transformation of the energy structure and environmental protection. However, the high C–H bond dissociation energy of CH4 and the overoxidation of methyl radical (˙CH3) intermediates greatly limit the conversion of CH4 to C2+ products. Herein, we demonstrate a metal–organic framework (MOF) crystal engineering strategy to synthesize a MOF-derived PdO/TiO2 nanocomposite for photocatalytic nonoxidative coupling of methane (NOCM), achieving high selectivity and activity in the conversion of CH4 to ethane (C2H6). Mechanistic investigations reveal that the spatially separated active sites for C–H bond cleavage and C–C coupling contribute to the efficient conversion of CH4 to C2H6. Specifically, the lattice oxygen captures the photogenerated holes, leading to the formation of oxygen radical anions (˙O−), which activate the C–H bond and generate ˙CH3 intermediates. PdO stabilizes ˙CH3 intermediates, effectively inhibiting the overoxidation of ˙CH3, and thereby promoting the C–C coupling process. This work opens a new avenue for the rational design of efficient MOF-derived photocatalysts for NOCM.
New conceptsMethane (CH4), the primary component of natural gas and one of the most abundant fossil energy sources, holds immense potential as a feedstock for producing high-value hydrocarbons. However, current industrial routes for methane transformation are highly energy-intensive, largely stemming from the high C–H bond energy and low polarizability of CH4. Herein, we report the design of a porous PdO/TiO2 nanocomposite through a MOF crystal engineering strategy for highly efficient and selective photocatalytic methane to ethane (C2H6). The optimal PdO/TiO2 photocatalyst achieves a C2H6 yield of 1196 μmol g−1 h−1 with a selectivity of 94%, outperforming most of the previously reported works on photocatalytic NOCM. Mechanistic investigations reveal that the spatially separated active sites (C–H cleavage on TiO2 and C–C coupling on PdO) contribute to the efficient conversion of CH4 to C2H6. This study provides a novel and efficient approach for CH4 conversion under mild conditions, highlighting the potential of MOF-derived photocatalysts in NOCM applications. |
The photocatalytic NOCM reaction involves multiple steps, including C–H activation, hydrogen (H) abstraction, and C–C coupling.8 These processes are closely linked to electron transfer, the redox reaction of photogenerated charges and the adsorption–desorption of reaction intermediates or products in the reaction on catalysts.9 However, previous NOCM studies typically rely on single metal sites to control the C–H bond activation and the C–C coupling, which inevitably limits the efficiency of CH4 conversion.10 To overcome these limitations, the rational design of photocatalysts with multiple active sites to facilitate both C–H activation and C–C coupling is crucial for improving the CH4 conversion efficiency.11,12 Metal–organic frameworks (MOFs), an emerging class of porous materials, have gained significant attention due to their well-defined and customizable structures, making them ideal for incorporating diverse active sites.13 Using a MOF as a precursor, the derived nanomaterials not only retain pre-designed structures but also exhibit enhanced photogenerated carrier separation performance,14 thereby promoting CH4 adsorption and activation. Moreover, the unique coordination environment provided by the MOF structure can further optimize the distribution and stability of the active sites,15 which can effectively ensure the stability of methyl (˙CH3) intermediates to facilitate C–C coupling.
Herein, we report the design of a porous PdO/TiO2 nanocomposite through a MOF crystal engineering strategy for highly efficient and selective photocatalytic NOCM. The optimal PdO/TiO2 photocatalyst achieves a C2H6 yield of 1196 μmol g−1 h−1 with a selectivity of 94%, outperforming most of the previously reported works on photocatalytic NOCM. Mechanistic investigations reveal that the spatially separated active sites for C–H cleavage and C–C coupling contribute to the efficient conversion of CH4 to C2H6. Electron paramagnetic resonance (EPR) spectra demonstrate that oxygen radical anions (˙O−) produced by photoexcitation promote the activation of the C–H bond of CH4 to form the ˙CH3 intermediates. PdO improves the adsorption of ˙CH3, effectively suppressing the overoxidation of ˙CH3 intermediates and lattice oxygen, and facilitating the C–C coupling process. This study provides a novel and efficient approach for CH4 conversion under mild conditions, highlighting the potential of MOF-derived photocatalysts in NOCM applications.
X-ray photoelectron spectroscopy (XPS) has been applied to determine the surface electronic state of the elements. The survey spectra (Fig. S5, SI), together with the high-resolution XPS spectra (Fig. 2a–c) of the PdO/TiO2 composites, confirm the presence of all elements (Pd, O, and Ti) associated with PdO and TiO2, which aligns well with the results of element mapping. The Pd 3d XPS spectrum is presented in Fig. 2a, and the peaks of 336.49 eV and 341.79 eV for PdO/TiO2 are attributed to Pd 3d5/2 and Pd 3d3/2 of Pd2+, respectively.20,21Fig. 2b displays the Ti 2p XPS spectra of the samples, where the peaks of 458.59 eV and 464.06 eV for TiO2 are ascribed to Ti4+.22 Furthermore, the peak observed at 458.04 eV is attributed to Ti3+, whose presence is typically accompanied by the formation of oxygen vacancies.23 In the O 1s XPS spectra shown in Fig. 2c, the sharp peak of 529.74 eV and the broad peak of 530.79 eV are assigned to lattice oxygen and surface-adsorbed oxygen.24 Notably, the characteristic peaks of Ti 2p and O 1s in PdO/TiO2 shift to higher binding energies after PdO modification, indicating the transfer of electrons from TiO2 to PdO.25
The XRD patterns (Fig. 2d) reveal that all samples exhibit peaks at 2θ values of 25.8°, 37.8°, and 48.0°, corresponding to the (101), (004), and (200) crystallographic planes of tetragonal anatase TiO2 (PDF# 21-1272). Additionally, the peak appearing at 33.6° in the PdO/TiO2 corresponds to the (002) crystallographic plane of PdO (PDF# 75-0584), indicating that PdO has a certain degree of crystallinity on the TiO2 surface; however, due to the low content or high dispersion of PdO, signals from other crystallographic planes are not detected.18 Nitrogen adsorption–desorption isotherms, shown in Fig. 2e and Fig. S6 (SI), demonstrate that the samples exhibit type IV isotherms with H3 hysteresis loops, indicating the presence of an irregular mesoporous structure.26,27 The specific surface area, pore size, and pore volume are detailed in Table S1 (SI). PdO/TiO2 exhibits a well-developed mesoporous structure with a large surface area (240.1 m2 g−1), which is beneficial for CH4 adsorption.28
Electron paramagnetic resonance (EPR) spectra (Fig. 2f) indicate a signal peak at g = 2.003 belonging to the oxygen vacancy,29 which is consistent with the XPS results. The introduction of PdO alters the local electronic structure of TiO2, facilitating electron transfer from TiO2 to PdO, which stabilizes Ti3+ species and promotes the formation of oxygen vacancies to maintain charge neutrality.30 Moreover, Ultraviolet-visible (UV-Vis) diffuse reflectance spectroscopy (DRS) has been used to measure the properties of optical absorption of the samples. As shown in Fig. S7a (SI), pristine TiO2 exhibits weak light absorption. In contrast, the modification with PdO significantly enhances the light absorption intensity of TiO2, which may be due to the interfacial interaction between PdO and TiO2.31 Moreover, on the basis of Tauc plots of (αhν)2versus photo energy (hν),32 the band gap energy (Eg) of TiO2 is derived to be 3.3 eV (Fig. S7b, SI).
Upon obtaining the structural information of the catalysts, we then conducted a systematic evaluation of the photocatalytic activity of TiO2 and PdO/TiO2 for NOCM. As shown in Fig. 3a, blank TiO2 exhibits a low C2H6 production rate (233 μmol g−1 h−1). After PdO modification, the C2H6 production rate of the PdO/TiO2 composite significantly increased, with a normal distribution trend observed in relation to varying PdO content (Table S2, SI). Notably, the 1PdO/TiO2 sample shows the optimal photocatalytic performance, with a C2H6 production rate increased to 565 μmol g−1 h−1. In addition, the MOF precursor (Ti-BPDC-Pd) was also evaluated under identical conditions and displayed negligible C2H6 formation (Fig. S8, SI). Considering the exceptional performance of noble metals such as Au and Pt in NOCM reaction,33,34 we prepared a series of TiO2 samples modified with the same proportion (1%) of different metals using the same method, and compared their photocatalytic activities for NOCM (Fig. S9, SI). As is seen clearly in Fig. S10 (SI), PdO/TiO2 demonstrates the highest C2H6 production rate among them. Additionally, we investigated the performance of other oxide supports (P25, ZnO, Al2O3 and CeO2) in place of TiO2 under the same reaction conditions (Fig. S11, SI). The results indicate that PdO/TiO2 exhibits a significantly higher C2H6 production rate than PdO/P25, PdO/Al2O3, PdO/ZnO, and PdO/CeO2 (Fig. 3b), suggesting that MOF-derived TiO2 is a highly promising support for NOCM reaction. Further analysis reveals a positive correlation between the photocatalytic activity of PdO/TiO2 and light intensity, whose C2H6 production rate can reach 1196 μmol g−1 h−1 at a light intensity of 300 mW cm−2 with a selectivity of 94% (Fig. S12 and S13, SI). The apparent quantum yield (AQY) for the NOCM reaction was also evaluated. The PdO/TiO2 catalyst exhibited an AQY of 1.1% under 360 nm monochromatic light irradiation. Under AM 1.5G simulated sunlight, the AQY decreased to 0.26%, which is significantly lower than that under monochromatic illumination. This reduction can be attributed to the fact that the catalyst mainly absorbs ultraviolet light, resulting in limited utilization of the visible-light portion of the solar spectrum.
The C2H6 yield of all samples increases and then stabilizes over time (Fig. 3c), and PdO/TiO2 reaches a C2H6 production of 701.5 μmol g−1 in the first hour, but the rate gradually decreases due to the consumption of lattice oxygen.8 This phenomenon is also further observed in the cycling experiment. As shown in Fig. S14 (SI), the photocatalytic activity of PdO/TiO2 decreases slightly with the increase of the cycle period. However, by heat treatment in air, the lattice oxygen on PdO/TiO2 is replenished, and the catalytic activity is successfully restored to its original level in subsequent cycles (Fig. 3d).10 To further investigate whether the decline in activity during cycling is associated with changes in the Pd oxidation state, XPS analysis was performed on the PdO/TiO2 catalyst after reaction. As shown in the Pd 3d spectra (Fig. S15, SI), the binding energies remain at 336.24 eV and 341.51 eV, which are characteristic of Pd2+. These results confirm that the valence state of Pd remains unchanged after the reaction. Therefore, the observed decrease in activity is more likely due to the temporary depletion of lattice oxygen, rather than changes in the Pd electronic structure. Furthermore, no CH4-derived products are detected for control experiments conducted in the absence of reactant, photocatalyst, or light irradiation (Fig. S16, SI), confirming that this is a light-driven process and CH4 is the sole source of carbon. Isotopic labelling tests using 13CH4 as the feedstock further confirm the origin of the resulting C2H6. As shown in Fig. 3e, dominant peaks ascribed to 13C2H6 and its molecular fragments can be observed (28 = 13C2H2, 29 = 13C2H3, 30 = 13C2H4, 31 = 13C2H5).35 These results collectively confirm that the generated C2H6 originates solely from the feedstock CH4. By virtue of the high activity and selectivity of C2H6, our optimal PdO/TiO2 catalyst demonstrates a competitive performance in NOCM compared to other recently reported alternatives (Fig. 3f, details are listed in Table S3, SI).
The photocatalytic NOCM primarily involves CH4 adsorption, C–H bond activation, and subsequent C–C coupling.6 To figure out the functional orientation of each part of PdO/TiO2 in photocatalytic NOCM, CH4 temperature-programmed-desorption (TPD) was performed to investigate the adsorption of CH4 on the samples. As disclosed in Fig. 4a, TiO2 exhibits two distinct desorption peaks at 211 °C and 379 °C, with the more intense peak appearing at 211 °C, indicating its strong adsorption capacity for CH4 at low temperatures. This enhanced adsorption performance can be ascribed to the well-developed mesoporous structure of MOF-derived TiO2, which provides abundant surface area and accessible channels for CH4 molecules. In contrast, the PdO/TiO2 composite displays a single dominant desorption peak at 251 °C with a notably reduced overall desorption intensity, suggesting a reduced CH4 adsorption capacity. This decrease is likely attributable to the PdO loading, which reduces the specific surface area and induces partial collapse of the pore structure, thereby hindering CH4 adsorption. Furthermore, the electrochemical impedance spectroscopy (EIS) and transient photocurrent response spectra (Fig. S17a and b, SI) show that the modification of PdO significantly improves the photogenerated charge separation efficiency. To gain deeper insight into the charge transfer process, in situ XPS measurements were conducted. As illustrated in Fig. 4b and c, upon light irradiation, the Pd 3d spectrogram shows clear peak splitting, with a peak at 343 eV, indicating that it has undergone partial reduction due to the gain of electrons. Concurrently, the O 1s peak shifts positively by 0.28 eV, implying that photogenerated holes tend to accumulate at the oxygen sites of TiO2 for CH4 activation.34In situ EPR measurements of the samples further support this conclusion. As disclosed in Fig. 4d, under an Ar atmosphere and dark conditions, no signal peak is detected. Upon light irradiation, the newly appeared signal peaks at gx = 2.031 and gy = 2.002 are attributed to ˙O− generated by surface lattice oxygen through capturing photogenerated holes.36 When CH4 was subsequently introduced into the EPR chamber, the intensity of the ˙O− clearly lowered, demonstrating that the ˙O− reacted with CH4.37 It is noteworthy that the similar changes observed in the EPR spectra of PdO/TiO2 and TiO2 suggest that CH4 activation primarily occurs on the surface of TiO2. Additionally, due to the electron-rich nature of the oxygen vacancies (Ov),38 Ov may be a favorable position for CH4 adsorption by promoting the formation of ˙O− on adjacent lattice oxygen sites to activate the C–H bond. To gain insight into the intermediates of photocatalytic NOCM on PdO/TiO2, under liquid-phase reaction conditions, EPR measurements were carried out using 5,5′-dimethyl-1-pyrroline N-oxide (DMPO) as a radical scavenger to further investigate the potential intermediates. As shown in Fig. 4e, distinct characteristic signals of ˙CH3 and ˙OH are detected on PdO/TiO2. Given that ˙OH originates from the water in the test conditions,39,40 this finding indicates that the surface ˙CH3 is the key intermediate for CH4 coupling.41 In comparison, the ˙CH3 signal on the surface of TiO2 was weak, suggesting the high efficiency of CH4 dissociation on PdO/TiO2.
To elucidate the detailed reaction mechanism underlying the photocatalytic NOCM, we employed an in situ Fourier transform infrared spectroscopy (FTIR) approach to track the dynamic evolution of surface-adsorbed species and key intermediates during the reaction. During the dark adsorption phase, characteristic peaks of C–H deformation vibration of CH4 appear around 1300 cm−1 as well as 3000 cm−1 (Fig. S18, SI).42 As shown in Fig. 4f, under light irradiation, a new peak around 1558 cm−1 is assigned to the C–H symmetry vibration of adsorbed CH4.43 Simultaneously, the ˙CH3 deformation vibration mode observed at 1469 cm−1 confirms that CH4 undergoes rapid dissociation by photogenerated charge carriers on the PdO/TiO2 surface.10 Additionally, the peak at 3528 cm−1 is attributed to the stretching and bending vibrations of the O–H bonds of H2O, indicating that the H atoms from CH4 dissociation combine with lattice oxygen to form H2O.8 Meanwhile, the slight growth of C
O (1671 cm−1) stretching vibrational modes is assigned to the species for CO2 formation, suggesting the sluggish overoxidation of CH4 on PdO/TiO2. In contrast, an intense C
O stretching vibration signal can be observed for TiO2 (Fig. S19, SI). Taken as a whole, the modification with PdO enhances the separation efficiency of photogenerated electron–hole pairs, thus promoting the dissociation of CH4 on the TiO2 surface. Additionally, PdO effectively stabilizes the ˙CH3 intermediates and diminishes the overoxidation of CH4 under light irradiation, facilitating the C–C coupling process to produce C2H6.
Based on the experimental findings, we propose a mechanistic pathway for the photocatalytic NOCM process over PdO/TiO2, as illustrated in Fig. 4g. Under photoexcitation, PdO/TiO2 nanocomposites generate electron–hole pairs, CH4 molecules are adsorbed at the Ov site, and the photogenerated holes interact with the nearby lattice oxygen, resulting in the formation of oxygen radical anions (˙O−). These radicals extract hydrogen from the C–H bond, initiating CH4 activation and producing the ˙CH3 intermediate.44 The PdO component plays a dual role by capturing and stabilizing ˙CH3 radicals, thus preventing their overoxidation and facilitating C–C coupling to form C2H6. Meanwhile, the photo-generated electrons participate in the reaction between the H atoms produced by the dissociation of CH4 and the lattice oxygen, generating H2O as a by-product. This cooperative interaction between TiO2 and PdO ensures efficient charge separation and controlled radical transformation, ultimately enabling the selective photocatalytic conversion of CH4 to C2H6 under mild conditions.45
| This journal is © The Royal Society of Chemistry 2026 |