Wendi Zhao†
a,
Kang Sun†a,
Jiayi Xua,
Zhongyuan Lina,
Qihui Chen*b,
Maochun Hong
b and
Hai-Long Jiang
*a
aHefei National Research Center for Physical Sciences at the Microscale, Department of Chemistry, University of Science and Technology of China, Hefei, Anhui 230026, P. R. China. E-mail: jianglab@ustc.edu.cn; Web: http://mof.ustc.edu.cn/
bState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, P. R. China. E-mail: chenqh@fjirsm.ac.cn
First published on 5th September 2025
While the rational fabrication of heterojunction photocatalysts with tunable interfaces and precise location control of cocatalysts holds great promise for enhanced photocatalysis, the synergistic integration of these parameters remains a substantial challenge. Herein, a series of metal–organic framework (MOF) composites with compact interfaces and customizable interface coverage are designed by epitaxial growth of ZIF-8 on the surface of MIL-125-NH2, yielding ZIF-8m/MIL-125-NH2 (m = 21, 35, 65, representing the coverage percentage of ZIF-8 on the MIL-125-NH2 surface). These composites are then converted into ZnO/TiOx heterojunctions through a two-step thermal treatment, termed ZTO-m, for photocatalytic CH4 oxidation. The results reveal that the interface coverage in ZTO-m plays critical roles in charge separation, where ZTO-65 gives the best activity. With ZTO-65 as a basis, the cocatalysts, Au clusters and CoOx species, are respectively positioned onto TiOx and ZnO. The targeted positioning of cocatalysts not only improves charge separation but also facilitates O2 activation. As a result, the resulting Au-Co-ZTO demonstrates excellent activity toward liquid oxygenate production, achieving 1723.5 μmol g−1 h−1 with a selectivity of 99%, in photocatalytic CH4 oxidation.
In heterojunction systems, interface interaction plays crucial roles in determining charge transfer efficiency.12 However, challenges such as lattice mismatch and chemical incompatibility often prevent the formation of tightly bonded heterojunction interfaces between different components.13 As a result, heterojunction photocatalysts are typically constructed by simple physical assembly, which tends to result in poor interface contact and inefficient charge transfer. Achieving a tight, well-structured interface and precisely regulated interface area/length between two semiconductor photocatalysts remains a significant challenge. In addition, cocatalysts are widely recognized for enhancing charge separation and substrate activation in photocatalysis, while their random distribution often diminishes these advantages.14 This limitation is particularly pronounced in heterojunction photocatalysts with complex components, where precise location control of cocatalysts becomes increasingly challenging. Therefore, the development of heterojunction systems with closely contacted interfaces and precisely positioned cocatalysts could significantly improve the photocatalytic efficiency. Unfortunately, such design has rarely been reported so far, and it remains a grand challenge to achieve using conventional methods.15,16
Metal–organic frameworks (MOFs), distinguished by their extraordinary structural versatility, have attracted widespread attention across diverse application fields.17–23 MOFs exhibit high structural compatibility and customizable skeletons, enabling the fabrication of composites with controllable components, sizes, and interfaces.24–26 Moreover, MOFs can serve as versatile precursors for creating different derivatives tailored to diverse applications.27–30 More importantly, these MOF-derived materials often inherit the original morphology and interface characteristics of their parent MOFs.31–33 By leveraging the inherent tunability of MOFs, it might be possible to fabricate MOF-based composites with tunable interfaces and precise cocatalyst placement as precursors, followed by thermal treatment to afford heterojunction photocatalysts with inherited structural features for improved photocatalysis.
In this work, binary MOF composites with controlled interface coverage were obtained by controlling the epitaxial growth of ZIF-8 on the surface of MIL-125-NH2, affording ZIF-8m/MIL-125-NH2 (m = 21, 35, 65, representing the coverage percentage of ZIF-8 on the MIL-125-NH2 surface). Subsequently, ZTO-m composites featuring varying heterojunction interface coverages between zinc oxide (ZnO) and titanium oxide (TiOx) were synthesized via a two-step thermal treatment of ZIF-8m/MIL-125-NH2 (Scheme 1a). The heterojunction interface area (with m as the descriptor; see SI, Discussion S1) is found to significantly influence charge separation, and as a result, ZTO-65 affords optimal performance toward photocatalytic CH4 oxidation. Further, Au clusters and CoOx species, serving as cocatalysts, were loaded onto ZIF-8 and MIL-125-NH2, respectively, to form the Au-Co-MOF composite with a ZIF-8 coverage of 65%, followed by thermal treatment to yield Au-Co-ZTO (Scheme 1b). Results demonstrate that the cocatalyst distribution not only suppresses the recombination of electrons and holes but also plays a crucial role in O2 activation. Remarkably, the optimized Au-Co-ZTO, featuring Au clusters on TiOx and CoOx on ZnO, achieves a production rate of 1723.5 μmol g−1 h−1 with 99% selectivity toward liquid oxygenate production in photocatalytic CH4 oxidation.
![]() | ||
Scheme 1 Illustration showing the stepwise synthesis route to the fabrication of (a) ZTO-m, (b) Au-Co-ZTO, and (c) MOF-derived ZnO + TiO2 mixture photocatalysts for CH4 oxidation. |
![]() | ||
Fig. 1 SEM images of (a) ZIF-865/MIL-125-NH2, and (b) Au-Co-ZTO. (c) and (d) EDS mapping of Au and Co elements for Au-Co-ZTO. |
The above ZIF-865/MIL-125-NH2 composites were transformed into ZTO-65 via pyrolysis in O2 at 425 °C, followed by N2 treatment at 600 °C (Scheme 1a). The first pyrolysis under an oxygen atmosphere produced metal oxides. During this process, the original morphology of MIL-125-NH2 is largely inherited in the resulting titanium oxide, while the external ZIF-8 is transformed into porous ZnO with a stacked structure, as previously reported.35 Powder XRD profiles reveal that, while ZIF-8 and MIL-125-NH2 are respectively converted to ZnO and TiO2 upon thermal treatment, for ZIF-865/MIL-125-NH2, only ZnO diffraction peaks can be observed, with the absence of TiO2 signals (Fig. S4). This absence of crystalline TiO2 might be attributed to the influence of Zn during pyrolysis, which inhibits the crystallization of TiO2, leading to the formation of amorphous TiOx.36 To enhance the interface interaction, higher pyrolysis temperatures were investigated.37 While temperatures above 500 °C promote the formation of tight ZnO/TiOx interfaces, too high temperature leads to structural collapse (Fig. S4 and S5). As a result of optimization, the ZIF-865/MIL-125-NH2 composite was first treated in an oxygen atmosphere at 425 °C to produce metal oxide products, followed by N2 treatment at 600 °C. The inert N2 atmosphere would facilitate compact interfacial contact and avoid structural collapse.38 Such a two-step process yields ZTO-65, in which not only a well-structured heterojunction is maintained but also a new Zn2TiO4 phase is created (Fig. S6 and S7).
The emergence of this new phase suggests a compact interface within the heterojunction, facilitating charge separation.39 As verification, photocurrent and electrochemical impedance spectroscopy (EIS) results confirm that ZTO-65 prepared via the two-step thermal treatment possesses superior charge separation compared to that produced by direct pyrolysis (Fig. S8). The results indicate that the two-step process not only largely preserves the morphology of the parent MOFs but also promotes the formation of a compact interface, improving charge separation. Additionally, ZTO-21 and ZTO-35, with different coverage levels, were also prepared as controls under identical transformation conditions (Fig. S9 and S10).
Moreover, two cocatalysts, Au clusters and CoOx, are introduced into the ZTO heterojunction system. The spatial arrangement of these cocatalysts is precisely controlled by varying the introduction sequence of cocatalyst precursors (Scheme 1b). Typically, Au clusters were first photodeposited on the surface of MIL-125-NH2, followed by the epitaxial growth of ZIF-8, during which the coverage percentage of ZIF-8 remained at the optimized value of 65%. Subsequently, Zn2+ within ZIF-8 was exchanged with Co2+ through ion exchange, resulting in a Au-Co-MOF with Au and Co localized on MIL-125-NH2 and ZIF-8, respectively. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images and the corresponding energy dispersive X-ray spectroscopy (EDS) mapping show that Au clusters are located on MIL-125-NH2 and the Co element is on ZIF-8 (Fig. S11). Control experiments confirmed that the Co element cannot be incorporated into MIL-125-NH2 by the ion exchange process, ensuring its incorporation into ZIF-8 only (Table S1).
The obtained Au-Co-MOF composite with well-defined Au/Co locations is then converted into the corresponding oxide, namely Au-Co-ZTO, with the above two-step thermal treatment procedure as that for ZTO-65. Electron microscopy observation and powder XRD results indicate that the crystallinity and morphology of Au-Co-ZTO are similar to those of ZTO-65 (Fig. 1b, S12 and S13). X-ray photoelectron spectroscopy (XPS) reveals that Au is in its metallic state, while Co exhibits a mixed oxidation state of +2 and +3, identified as CoOx (Fig. S14).40 Furthermore, the Au clusters are hardly observed, indicating their tiny sizes, and primarily located on the central TiOx based on EDS mapping results, whereas CoOx species are predominantly distributed on the peripheral ZnO (Fig. S15, 1c and d). For comparison, ZTO-65 with randomly distributed cocatalysts, referred to as Au+Co+ZTO, was also obtained by random impregnation of Au and Co species into ZTO-65 (Fig. S16 and S17).
Entry | Catalyst | Yield (μmol g−1 h−1) | Liquid product selectivity | |||||
---|---|---|---|---|---|---|---|---|
CO2 | CH3OH | CH3OOH | HCHO | HCOOH | Liquid products | |||
a Reaction conditions: catalyst (5 mg), O2 (1 bar), CH4 (20 bar), H2O (10 mL), 25 °C, 300 W Xe lamp (320–800 nm). | ||||||||
1 | ZIF-8-derived ZnO | 0 | 51.2 | 168.9 | 96.4 | 13.2 | 329.7 | 1.00 |
2 | MIL-125-NH2-derived TiO2 | 152.3 | 95.4 | 15.6 | 325.0 | 14.8 | 450.8 | 0.75 |
3 | ZTO-21 | 9.0 | 363.8 | 25.7 | 82.7 | 3.9 | 476.1 | 0.98 |
4 | ZTO-35 | 3.7 | 128.6 | 176.7 | 248.0 | 7.8 | 561.1 | 0.99 |
5 | ZTO-65 | 0 | 58.9 | 391.5 | 163.6 | 14.7 | 628.7 | 1.00 |
6 | Mixture of MOF-derived ZnO and TiO2 | 40.0 | 26.0 | 114.3 | 167.9 | 11.7 | 319.9 | 0.89 |
7 | Au-Co-ZTO | 18.9 | 145.5 | 1195.2 | 360.2 | 22.6 | 1723.5 | 0.99 |
8 | Au+Co+ZTO | 41.3 | 421.4 | 331.0 | 295.2 | 35.9 | 1083.5 | 0.96 |
Based on the high activity of MIL-125-NH2-derived TiO2 and high selectivity of ZIF-8-derived ZnO, it is assumed that ZTO-m composites could combine their respective advantages. Compared to their single-component counterparts, ZTO-21 presents slightly increased activity for liquid oxygenate production with high selectivity, highlighting the importance of the heterojunction in promoting CH4 oxidation (Table 1, entry 3). Furthermore, as the coverage increases, the photocatalytic activity improves in the order of ZTO-65 > ZTO-35 > ZTO-21 (Table 1, entry 3–5). Among these composites, ZTO-65 shows the best performance, achieving an activity of 628.7 μmol g−1 h−1, which is 1.9 times higher than that of ZIF-8-derived ZnO and 1.4 times higher than that of MIL-125-NH2-derived TiO2. Notably, ZTO-65 not only exhibits higher activity, but also gives 100% selectivity toward liquid oxygenate production. As a control, the physical mixture of the MOF-derived ZnO and TiO2 shows much lower activity and selectivity than the heterojunction composites (Table 1, entry 6; Fig. S20). These results unambiguously underscore the critical role of interface coverage for selective CH4 oxidation.
The above results reveal that the interface coverage within heterojunction photocatalysts can be engineered using highly designable MOF composites as precursors (see SI, Discussion S1), which significantly influences the photocatalytic performance in CH4 oxidation. To further improve CH4 oxidation activity, Au and CoOx cocatalysts have been further introduced. First, the photocatalytic performance of ZTO-65 with selectively loaded Au clusters on the TiOx component has been evaluated. Along with increased loading of Au clusters, the photocatalytic activity follows a volcano-type trend (Fig. S21a). At the optimal Au loading of 0.58 wt%, the CoOx cocatalyst was further introduced into the ZnO component, which indicates a similar activity trend as observed with Au loading change (Fig. S21b). When the Au and Co loadings reach 0.58 wt% and 0.39 wt%, respectively, the resulting Au-Co-ZTO exhibits the highest activity, achieving a liquid oxygenate production rate of 1723.5 μmol g−1 h−1 with 99% selectivity (Table 1, entry 7). Compared with previous reports,40,46–48 the excellent performance of the heterojunction system with oxygen as the oxidant is significant (Table S2). In contrast, the Au+Co+ZTO catalyst, where Au and CoOx cocatalysts are randomly distributed, showcases significantly lower activity than Au-Co-ZTO, highlighting the significance of precisely loaded and placed cocatalysts for enhanced CH4 oxidation (Table 1, entry 8).
Furthermore, an isotope labeling experiment has been conducted to trace the source of liquid oxygenate production (Fig. 2a). When pure 12CH4 is used as the feedstock, two distinct single peaks are observed at 3.18 ppm and 3.68 ppm in the 1H-nuclear magnetic resonance (1H-NMR) spectrum, corresponding to 12CH3OH and 12CH3OOH. In contrast, when a mixed gas of 13CH4:
12CH4 in a 1
:
2 ratio is employed to replace pure 12CH4, two split peaks appear due to the coupling splitting caused by the presence of 13C.49 Additionally, the ratio of 13CH3OH/12CH3OH is 1
:
2, consistent with the ratio of feeding gases. These results clearly confirm that the liquid oxygenate production originates from CH4, rather than the decomposition of residual carbon. The stability of Au-Co-ZTO during photocatalysis has also been examined. After 5 consecutive photocatalytic cycles, no significant changes are observed in both activity and selectivity (Fig. 2b). Characterization confirms that the crystallinity and integrity are well preserved, suggesting the stability of the Au-Co-ZTO photocatalyst (Fig. S22–S25 and Table S3).
Additionally, the large interface areas within this heterojunction structure facilitate charge separation, contributing to the enhanced photocatalysis. All ZTO-m samples display similar light absorption properties, confirming that the differences in photocatalytic activity are not due to variations in light harvesting (Fig. S27). Steady-state photoluminescence (PL) spectroscopy reveals that ZTO-65 exhibits the weakest fluorescence intensity (Fig. S28), suggesting the most efficient electron–hole separation among ZTO-m. Similarly, photocurrent measurements and EIS also suggest that ZTO-65 possesses the best charge separation efficiency and the lowest interface charge transfer resistance (Fig. S29 and S30). Furthermore, time-resolved transient photoluminescence reveals that ZTO-65 has the longest electron lifetime (0.35 ns), followed by ZTO-35 (0.30 ns) and ZTO-21 (0.29 ns) (Fig. S31 and S32). The above results collectively demonstrate that the superior photocatalytic performance in ZTO-65 arises from its large interface coverage in the heterojunction structure, which greatly enhances separation of electrons and holes, thereby benefiting the photocatalysis.
Moreover, the importance of precise cocatalyst placement in promoting charge separation has been investigated. Due to the overlap between Au 4f and Zn 3p XPS signals,45 the sample directly from Au/MIL-125-NH2 (similar thermal treatment to that for preparing Au-Co-ZTO) was employed instead to study the electronic state changes of Au under light irradiation. In situ XPS spectra reveal that, under light irradiation, the Au and Co peaks shift by −0.3 eV and +0.2 eV, respectively, indicating that Au clusters and CoOx function as reduction and oxidation cocatalysts, respectively (Fig. 2e and f). Therefore, the electron reduction reaction occurs on Au, while the hole oxidation reaction takes place on CoOx (Fig. S33).50 In the heterojunction system, the spatial location of Au and CoOx would greatly impact charge separation efficiency. All steady-state PL, photocurrent, and EIS measurements confirm that Au-Co-ZTO exhibits superior charge separation efficiency compared to Au+Co+ZTO with Au and CoOx randomly distributed (Fig. S34). Time-resolved transient PL spectroscopy further indicates that Au-Co-ZTO has a longer fluorescence lifetime than ZTO-m without cocatalysts and Au+Co+ZTO (Fig. S35 and S36). These observations support that the precise cocatalyst placement is critical for promoting charge separation in heterojunction photocatalysts.
In addition to charge separation, the CH4 activation process also plays a key role in photocatalytic CH4 oxidation. The oxygen source for liquid oxygenate production was determined to elucidate the reaction pathway. With Au-Co-ZTO as the catalyst, mass spectrometry (MS) analysis of the liquid oxygenate production gives the CH3OH signal only, as CH3OOH is unstable under the testing conditions.51 The signals of CH318OH and CH316OH can be observed when 18O2 and H216O are used (Fig. 3a), suggesting that the oxygen in the products originates from both water and oxygen.
Furthermore, electron paramagnetic resonance (EPR) spectroscopy was employed to identify the reactive oxygen species (ROS) involved in the photocatalytic process. To identify signals originating from oxidation reactions, the catalysts were dispersed in water, and AgNO3 was added to suppress interference from reduction-related ROS, such as ˙OOH.52 No signal is detected in the dark when DMPO is used as the radical trapping agent (Fig. 3b). Upon light irradiation, DMPO-˙OH signals are observed in the presence of Au-Co-ZTO, indicating the production of hydroxyl radicals (˙OH) derived from H2O oxidation.52 Similarly, the catalysts were dispersed in methanol (as a hole sacrificial agent) to investigate the ROS produced from reduction reactions. In this case, only DMPO-˙OOH signals are detected, while ˙OH signals are absent (Fig. 3c). The DMPO-˙OOH signals suggest that O2 reacts with an electron and H+ to form hydroperoxyl radical (˙OOH) species, which are milder ROS than superoxide anions (O2˙−).53 These EPR results demonstrate the formation of ˙OH and ˙OOH, where the former is produced from water oxidation by holes, while the latter results from the reduction of O2 by electrons.
To further elucidate the roles of ˙OH and ˙OOH, quenching experiments were conducted (Fig. S37). The addition of salicylic acid, a ˙OH scavenger, significantly suppresses product formation, indicating that CH4 activation is primarily driven by ˙OH. When Na2C2O4, the hole scavenger, is added, the activity decreases but is not completely suppressed. Introducing H2O2 partially restores the activity, suggesting that water oxidation might not be the sole pathway for the ˙OH production. A smaller portion is generated from the decomposition of H2O2, which is produced from the further conversion of partial ˙OOH.54 Unfortunately, ˙OH from H2O2 is hardly detected by EPR due to its low concentration.55
According to these results, a potential photocatalytic CH4 oxidation mechanism is proposed (Fig. 3d). During photocatalysis, photogenerated holes oxidize H2O to produce ˙OH, while electrons reduce O2 to form ˙OOH. A fraction of ˙OOH undergoes hydrogenation to yield H2O2, which subsequently decomposes to release additional ˙OH.56 The generated ˙OH abstracts a hydrogen atom from CH4, forming methyl radicals (˙CH3) and initiating the methane activation (step I).57 The ˙CH3 radicals then react with ˙OOH to produce CH3OOH, which undergoes further transformation to CH3OH (step II, III). In addition, CH3OH also can be formed through the direct reaction between ˙OH and ˙CH3 (step IV).57 Dehydration of CH3OOH or further oxidation of CH3OH results in HCHO, which can be subsequently oxidized to HCOOH and eventually CO2 (step V–VIII).
Based on the reaction mechanism, the production rates of ˙OH and ˙OOH were quantified to unveil the influence of cocatalyst placement on surface reactions. EPR spectra show similar ROS production over Au+Co+ZTO and Au-Co-ZTO, indicating that the reaction pathway is undisturbed by the cocatalyst placement (Fig. 3b and c). However, while the DMPO-˙OH signal intensity is comparable, the DMPO-˙OOH intensity is obviously weaker for Au+Co+ZTO than Au-Co-ZTO. This suggests that the cocatalyst placement significantly impacts the oxygen reduction to ˙OOH but has a minimal impact on H2O oxidation to ˙OH, which avoids over-oxidation of liquid oxygenate production by excess ˙OH. Furthermore, the production rate of different ROS was quantified by probe experiments for the two catalysts. Specifically, coumarin acted as the fluorescence probe for ˙OH detection, and the degradation rate of nitrotetrazolium blue chloride (NBT) was adopted to examine ˙OOH.43 The PL intensity growth rate of 7-hydroxycoumarin originating from the reaction between coumarin and ˙OH is nearly equal, supporting that the placement of cocatalysts does not affect the water oxidation to ˙OH (Fig. 4a and S38). However, the NBT degradation rate, indicative of ˙OOH production, is higher for Au-Co-ZTO, consistent with the EPR results (Fig. 4b and S39). These findings demonstrate that the cocatalyst placement not only improves the charge separation, but also accelerates the O2 activation to ˙OOH for boosting photocatalytic activity.
In addition, mechanisms regarding how the cocatalyst placement influences CH4 activation have been investigated. XPS analysis was employed to disclose the electronic states of Au and Co, which affect the activation of intermediates. The Au 4f7/2 signal for Au+Co+ZTO appears at 83.5 eV, while it shifts to 83.3 eV for Au-Co-ZTO, indicating an increased electronic density of Au in the latter (Fig. 4c). This increased electronic density of Au clusters in Au-Co-ZTO enables efficient electron donation to O2, thereby facilitating activation of O2.58 In contrast, no difference is observed in the Co XPS signals between the two catalysts (Fig. 4d). This indicates that the absorption and activation of H2O on CoOx are comparable, consistent with their similar abilities to produce ˙OH from water oxidation (Fig. 3b and 4a). These results suggest that the cocatalyst distribution modulates the electronic states of Au, further enhancing O2 activation for improved CH4 oxidation.
On the basis of ZTO-65 with optimized interface coverage, the cocatalysts, Au clusters and CoOx, are further deposited on TiOx and ZnO, respectively. The spatial distribution of cocatalysts is found to affect charge separation, and meanwhile, improve O2 activation, further optimizing the performance for photocatalytic CH4 oxidation. Accordingly, the optimized Au-Co-ZTO achieves a liquid oxygenate production yield of 1723.5 μmol g−1 h−1 with an impressive selectivity of 99%. This work not only highlights the great potential of MOFs toward photocatalysis, but also offers deep insights into the development of heterostructured photocatalytic materials.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |