Previous Article in Journal
The Mutual Influence of Elemental S and Cs on the Ni(100) Surface at Room and Elevated Temperatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Conversion of Methane to Methanol on Fe-Cu/ZSM-5 Under Mild Conditions: Effective Utilization of Free Radicals by Favorable Valence Ratios

1
School of Environment and Energy, South China University of Technology, Guangzhou 510006, China
2
Guangdong Provincial Key Laboratory of Atmospheric Environment and Pollution Control, South China University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Surfaces 2025, 8(4), 69; https://doi.org/10.3390/surfaces8040069
Submission received: 4 August 2025 / Revised: 7 September 2025 / Accepted: 19 September 2025 / Published: 23 September 2025
(This article belongs to the Special Issue Surface and Interface Science in Energy Materials)

Abstract

The selective oxidation of methane to methanol under mild conditions remains a significant challenge due to its stable C-H bond and the propensity for overoxidation of products. Herein, we investigated the Fe- and Cu-modified ZSM-5 catalysts using H2O2 as an oxidant for the selective oxidation of methane. It was found that the Fe/Cu ratio had a great impact on methanol yield. The Fe3Cu1 displayed the highest methanol yield of 29.7 mmol gcat−1 h−1 with a selectivity of 80.9% at 70 °C. Further analysis revealed that Fe3Cu1 showed the highest Fe3+ and Cu+ contents. The optimal dual valence cycle not only facilitates the efficient utilization of H2O2, promoting the activation of methane to •CH3 at the Fe site, but also suppresses the deep oxidation caused by the Fenton-like effect of Fe/H2O2, thus maintaining the high yield and high selectivity of methanol.

1. Introduction

Methane is the main component of natural gas and combustible ice [1,2]. However, most natural gas reserves are located in remote or offshore areas, making direct utilization challenging [3]. Converting methane into liquid fuels like methanol offers a promising solution for transportation-related issues. Currently, industrial methane utilization relies heavily on indirect conversion processes, involving methane reforming followed by Fischer–Tropsch synthesis, which typically requires high temperatures, high pressures, and significant energy inputs [4,5,6]. As a result, there is a pressing need for direct oxidative conversion of methane under mild conditions [7]. However, the stable C-H bond of methane (438.8 kJ mol−1) makes it difficult to activate under mild conditions. Furthermore, the desired oxidation products (C1 oxygenates) are often more reactive than methane itself, leading to over-oxidation and the formation of CO2 [8,9]. Thus, achieving efficient methane conversion while maintaining high selectivity for C1 oxygenates remains a major challenge.
Direct catalytic conversion of methane offers an environmentally friendly and energy-efficient method, enabling methane-to-methanol transformation at room temperature with the presence of a catalyst, which plays a critical role in facilitating the process. By simulating the structure of natural methane monooxygenase (MMO), loading noble metals [10,11] and transition metals [12,13,14] on zeolites shows great promise in methane conversion [15]. The binuclear iron center formed on Fe/ZSM-5 resembles the active center of the diiron species in soluble methane monooxygenase (sMMO) [16], and the extra-framework Fe3+ species are key species for methane transformation [14,16], which promote the dissociation of C-H bonds and effectively activate methane. Meanwhile, the Fenton-like reaction of Fe with H2O2 can generate a significant amount of free radicals, which promote the cleavage of the C-H bond in methane and enhance its conversion efficiency [17]. However, the large number of free radicals also leads to the over-oxidation of C1 oxygenates, resulting in reduced selectivity of methanol. Consequently, a key research focus is to simultaneously improve methane conversion activity and enhance methanol selectivity.
Studies have demonstrated that Cu-based catalysts can enhance methanol yield in methane conversion. The presence of Cu species consumes •OH radicals, thereby preventing the over-oxidation of methanol and suppressing the formation of formic acid and CO2 [1,18]. CuO-loaded zeolite can significantly improve methanol selectivity by decomposing excess H2O2 on CuO species during the reaction, which inhibits further oxidation of methanol [10]. Additionally, research reports that Cu species promote the generation of •OH radicals rather than •OOH radicals from H2O2, facilitating methanol production. These •OH radicals further react with •CH3 intermediates to form methanol, thus improving the selectivity [19].
Given the ability of Cu to effectively suppress the over-oxidation of methane, this study focuses on enhancing methane activation by loading Fe onto ZSM-5 while introducing Cu to mitigate the over-oxidation effects associated with Fenton-like reactions. In this work, a series of high-loading Fe-Cu/ZSM-5 catalysts were prepared using the incipient wetness impregnation method, and the effects of varying Fe/Cu ratios on methane oxidation activity and selectivity were systematically investigated. On one hand, the addition of Cu to Fe-based catalysts alters the structural and surface properties [20]. On the other hand, the synergistic interaction between Fe and Cu bimetallic species on ZSM-5 facilitates redox cycling and promotes the activation of H2O2, enabling the continuous generation of active radicals [21,22], thereby achieving efficient methane conversion. The metal loaded on Fe3Cu1 showed good dispersion with the highest Fe3+ and Cu+ content, and its bimetallic valence cycle efficiently utilized H2O2, exhibited the best methane conversion activity, achieving a methanol yield of 29.7 mmol gcat−1 h−1 with a selectivity of 80.9%, and good cyclic stability.

2. Experimental Sections

2.1. Materials and Reagents

Fe(NO3)2 9H2O, Cu(NO3)2 3H2O, sulfuric acid (H2SO4), acetic acid, ammonium acetate, acetylacetone, isopropanol (IPA), p-benzoquinone (p-BQ), and 5,5-dimethyl-1-pyrroline N-oxide (DMPO) were purchased from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China) and were analytical grade without further purification. HZSM-5 (25) was purchased from the Nankai University Catalyst Plant. Deionized water was used throughout the whole experiment.

2.2. Synthesis of Catalysts

A series of Fe-Cu/ZSM-5 catalysts with a fixed metal content of 10 wt.% and variable Fe and Cu loading amounts were prepared by incipient wetness impregnation. In a typical synthesis, the support of 4 g HZSM-5 was calcined in air at 550 °C for 4 h as pretreatment. Then, 2 mL of aqueous solution with different Fe(NO3)3 and Cu(NO3)2 concentrations were slowly added to the support with thorough stirring at room temperature. After ultrasonic treatment for 10 min, the support was dried at 30 °C for 48 h and then calcined in air at 550 °C for 4 h. The Fe and Cu loading amounts were determined by ICP-OES (Table S1) and labeled as Fe/ZSM-5, Cu/ZSM-5, Fe3Cu1, Fe1Cu1, and Fe1Cu3.

2.3. Characterizations

X-ray diffraction (XRD) patterns were determined by a Bruker D8 Advance X-ray instrument equipped with Cu-Kα radiation at 2θ of 20–60° with a scanning speed of 5°/min. High-resolution transmission electron microscope (HRTEM) images were captured on the Talos F200X instrument (FEI Co., Ltd. Shanghai, China). The N2 adsorption–desorption isotherms were obtained at 77 K using a Micromeritics ASAP 2460 instrument (Shanghai, China). X-ray photoelectron spectroscopy (XPS) spectra were collected on the Thermo Scientific EscaLab Xi+ (Waltham, MA, USA) equipped with an Mg-Kα radiation source (1253.6 eV). Fourier transform–infrared spectroscopy (FT-IR) was carried out on a Nicolet iS50 spectrometer (Thermo Scientific, Waltham, MA, USA). The absorbance of the liquid mixture was recorded on visible spectroscopy (722N, Shanghai Jinghua Technology Instrument Co., Ltd. Shanghai, China). CH4 temperature-programmed desorption (CH4-TPD) experiments were performed on a Micromeritics AutoChem II 2920 chemisorption analyzer (Shanghai, China). UV–vis diffuse reflection spectra (UV–vis DRS) were recorded on a Shimadzu UV-3600 instrument (Shimadzu Instruments, Suzhou, China). Reactive free radicals, including methyl radicals (•CH3), hydroxyl radicals (•OH), and hydroperoxyl (•OOH), were monitored by electron paramagnetic resonance (EPR, Bruker E500 spectrometer, Billerica, MA, USA) instrument by using DMPO as the radical trap reagent.

2.4. Catalytic Activity Test

The methane conversion and determination process are illustrated in Figure S1. Methane oxidation was conducted in an 80 mL stainless-steel autoclave with a Teflon vessel. Typically, 25 mg of catalyst and 20 mL of 0.5 M H2O2 were added to the reactor, which was then purged with pure CH4 (>99.9%) three times and pressurized to 30 bars. The autoclave was heated to 70 °C and stirred at 800 rpm. After the reaction, the products were rapidly cooled to room temperature using ice water. The gas components (CH4 and CO2) were determined by gas chromatograph equipped with a flame ionization detector (FID) and a thermal conductivity detector (TCD) (GC9720 Plus, Zhejiang Fuli Analytical Instrument Co., Ltd. Wenling, China). After centrifugation, the liquid products, including CH3OOH and CH3OH, were quantitatively analyzed with 1H-NMR and gas chromatography with an FID (GC9720 Plus, Zhejiang Fuli Analytical Instrument Co., Ltd.). For 1H-NMR, typically, 0.5 mL of liquid product was mixed with 0.1 mL of deuterium oxide (D2O), and 0.05 μL of dimethyl sulfoxide (DMSO, 99.99%, Shanghai Sinopharm, Shanghai, China) was added as an internal standard. Calibration curves were constructed between 1H-NMR area and concentrations of standard product solutions. The concentration of HCHO was determined using a colorimetric method, which relies on the reaction of acetylacetone with HCHO in the presence of acetic acid and ammonium acetate [23,24]. Meanwhile, the H2O2 concentration was quantified through the conventional titanium oxalate potassium method and analyzed using visible spectroscopy [11,25].
The yields of catalysts and selectivity were calculated using the following equation:
Yield of C1 oxygenates (mmol gcat−1 h−1):
Y p r o d u c t s = n p r o d u c t s m c a t a l y s t × t
Selectivity of CH3OH (%):
S C H 3 O H = n C H 3 O H n C H 3 O H + n C H 3 O O H + n H C H O + n C O 2 × 100 %

3. Results and Discussion

3.1. Structural and Morphological Characteristics

The states of Fe and Cu on ZSM-5 were analyzed by HRTEM. The hexagonal prism of ZSM-5 remained unchanged, with Cu and Fe particles dispersed on its surface. Lattice fringes with lattice spacings of 0.22 nm and 0.36 nm, corresponding to the (113) and (012) planes of Fe2O3, respectively (Figure S2a), were observed on Fe/ZSM-5. Similarly, lattice fringes with lattice spacing of 0.36 nm and 0.25 nm identified on Fe3Cu1 and Fe1Cu3 were assigned to the (012) and (110) planes of α-Fe2O3, respectively (Figure 1a,b). The energy dispersive spectrometer (EDS) shows that Cu was more uniformly distributed on the ZSM-5 surface, as well as better dispersion of Fe on Fe3Cu1, compared to Fe1Cu3, where Fe was more aggregated on the ZSM-5 exterior.
The structure of the synthesized samples was characterized by XRD (Figure 1c). The XRD patterns after loading Fe were similar to that of pure HZSM-5, implying that both the Fe- and Cu-loaded catalysts maintain the structure of the ZSM-5. When the content of Cu or Fe increased, peaks of Fe2O3 and CuO were observed on Fe/ZSM-5 and Cu/ZSM-5, respectively. The peaks located at 2θ of 33.16, 35.63, 49.46, and 54.07 correspond to Fe2O3 (JCPDS 99-0060) [26]; 2θ of 35.25, 38.54, and 48.71 are attributed to CuO (JCPDS 89-2530); and 2θ of 30.17 and 35.6 are attributed to Cuprospinel (JCPDS 25-0283 CuFe2O4). The bimetallic catalysts showed no obvious oxide peaks, suggesting that Fe and Cu species are finely dispersed on ZSM-5 as amorphous oxides or nanocomposites. It can be inferred that the Fe-Cu interaction favored the dispersion of its oxides.
Additionally, an increase in the Fe/Cu ratio promoted the dispersion of Cu species and reduction of both CuO and Fe2O3 [27]. Figure 1d shows the N2 adsorption–desorption isotherms, where distinct type IV adsorption–desorption isotherms were identified, suggesting that these catalysts are predominantly microporous in structure [28]. The number of micropores decreased significantly after metal loading, indicating that the active component blocks some of the pores of zeolite. At a relatively high pressure (P/P0) of 0.5–1.0, the N2 adsorption–desorption isotherm showed an H4 hysteresis loop, reflecting a mesoporous structure [29]. The microporous pore sizes of the catalysts were mainly distributed in 1.5–2 nm, with a minor presence of mesopores between 3.5 and 4.5 nm (Figure S3). Table 1 shows that the specific surface area of the bimetallic catalyst increased with the Cu loading ratio, likely due to the better dispersion of Cu on the zeolite [27]. Additionally, the decrease in Fe loading contributes to mitigating the agglomeration phenomenon.
As shown in Figure 2a, the CH4-TPD tests revealed that the introduction of Cu and Fe had a significant effect on methane adsorption. The desorption temperature decreased after loading Cu alone, while a broad desorption band was observed for Fe3Cu1. The higher desorption temperature and significantly larger desorption peak area indicate that Fe3Cu1 exhibits a stronger capacity for methane adsorption, which is beneficial for subsequent methane conversion [30]. Subsequently, the valence states of Fe and Cu on the catalyst surface were analyzed by XPS. Figure 2b shows the XPS spectrum of Fe 2p3/2 and 2p1/2, signifying the presence of both Fe2+ and Fe3+ species on the surface of ZSM-5. The peak around 710.7 eV is assigned to Fe2+ species, while the peak around 712.8 eV corresponds to Fe3+ species [31]. The relative contents and ratios of Fe2+ and Fe3+ are summarized in Figure 2d. As the Fe proportion increased, the relative content of Fe3+ also rose. There are studies [32] using Density Functional Theory (DFT) calculations that found that Fe(IV) clusters generated by the reaction of Fe(III) and H2O2 are able to promote homolytic activation of methane, leading to the formation of •CH3, which means Fe3+ plays an important role in methane conversion. Figure 2c shows the XPS spectra of Cu 2p3/2 and Cu 2p1/2, with the shakeup satellites centered around 944.5 eV [33]. The absorption peaks of Cu 2p3/2 and Cu 2p1/2 can be divided into two peaks with Cu2+ located at 934.8 eV and 933.5 eV for Cu+ species [34,35]. The relative contents of Cu2+ and Cu+ are displayed in Figure S4. As the Cu proportion increased, the Cu+ content decreased, implying a higher presence of CuO [36], consistent with the CuO peaks observed in the XRD pattern of Cu/ZSM-5. Cu+ can effectively promote the selectivity of methanol [37], and Fe3Cu1 presented the highest Fe3+/Cu+ ratio among the bimetallic catalysts.
UV–vis DRS spectroscopy was applied to understand the nature and coordination of copper and iron species in the samples. The absorption bands related to the ligand geometry and environment are related to the ligand-to-metal charge-transfer (LMCT), where the framework Fe adsorption bands are in the range of 200–250 nm, 250–350 nm for isolated and oligomerized extra-framework Fe in zeolite channels, 350–450 nm for larger Fe clusters, and >450 nm for bulk FexOy particles [12]. The UV–vis DRS spectra were normalized, and it can be seen from the magnified portion of Figure 2e that the isolated Fe3+ at the cationic positions on the surface of Cu/ZSM-5 was replaced by Cu compared to Fe/ZSM-5, and therefore the peak of isolated Cu2+ appeared (213 nm) [28]. The peaks maintained high intensity over a broad range, suggesting that different Fe species existed [38], which is consistent with the results of the TEM. The absorption bands in the spectral range of 300–450 nm indicate the oligomeric Fe/Cu species. Fe3Cu1 displays the highest isolated Fe3+ and oligomeric clusters band in the <300 nm range and 300–400 nm range, respectively, and the extra-framework binuclear Fe is usually considered to be the active site for selective oxidation of methane [12,39]. In addition, the band at >450 nm is significantly lower than that of other catalysts, indicating the least amount of bulk Fe2O3 particles, which accelerates the catalytic decomposition of H2O2 as well as promotes the further oxidation of the C1 oxygenates, unfavorable for methane to methanol [40]. Appropriate proportions of Cu can modulate the ratios of different Fe species and thus affect their catalytic properties.

3.2. Catalytic Performance Evaluation

The direct oxidation reaction of methane was carried out in a stainless-steel autoclave as described in the experimental section. As shown in Figure 3a, pure HZSM-5 performed poorly with a CH3OH yield of 2.1 mmol gcat−1 h−1. After introducing Fe, the Fe/ZSM-5 significantly increased the CH3OH yield to 15.4 mmol gcat−1 h−1. Although Cu/ZSM-5 showed a similar CH3OH yield to HZSM-5, more HCHO was produced, demonstrating a stronger methane conversion capability. Fe/ZSM-5 achieved the highest methane conversion with a total C1 oxygenates and CO2 yield of 38.1 mmol gcat−1 h−1, but its selectivity of C1 oxygenates was low at 44.4%. The optimal Fe3Cu1 catalyst demonstrated a CH3OH yield of 29.7 mmol gcat−1 h−1 with a selectivity of 80.9%. In summary, Cu incorporation significantly reduces CO2 formation, and an optimal Fe/Cu ratio enhances both CH3OH and selectivity. However, increasing Cu content led to higher HCHO production and lower CH3OH yield. Fortunately, CO2 yield can still remain low. Excess Fe tends to cause excessive oxidation of C1 oxygenates, which can be inhibited by the presence of Cu.
The catalytic performance was evaluated by varying the reaction temperature with respect to Fe3Cu1. Methane oxidation was observed at room temperature (Figure 3b). Within 30–60 °C, the CH3OH yield increased with temperature, indicating that the appropriate high temperature is favorable for the activation of methane. At 60 °C, the yield of CH3OH reached its maximum of 89.1 mmol gcat−1 h−1, with a selectivity of 91.3%. Further temperature increases led to a decline in CH3OH yield, likely due to the high temperature that promoted the deep oxidation of CH3OH to formic acid and CO2. Additionally, the decomposition of H2O2 to O2 and H2O at higher temperatures may contribute to the reduced CH3OH yield. The effect of reaction time on oxidation products was also investigated. As shown in Figure 3c,d, the CH3OH yield decreased while the HCHO yield increased over time on Fe/ZSM-5, accompanied by a continuous rise in CO2 selectivity. This suggests that CH3OH is continuously oxidized over the Fe/ZSM-5 catalyst to HCHO and CO2. In contrast, CH3OH yield stayed up on Fe3Cu1 and remained high after 5 h reaction (61.0 mmol gcat−1), with little HCHO production and stable CO2 selectivity (<25%).
Figure S5 shows the effect of different total pressures on the methane conversion performance of Fe3Cu1. Methane conversion exhibited a positive correlation with increasing pressure, with the highest CH3OH yield observed at 30 bars. Additionally, the effect of H2O2 concentration on methane conversion was evaluated (Figure 4a). It was found that CH3OH yield increased and then decreased with rising H2O2 concentration, while CO2 selectivity continuously increased, suggesting that excessive H2O2 tends to lead to deep oxidation of methane. Stability test results (Figure 4b) showed that the Fe3Cu1 maintained promising initial stability for oxidation of methane to methanol, with a slight decrease in CH3OH and a minor increase in CO2 yield after four cycles. This decline may be due to the leaching of loaded Fe and Cu during the reaction and centrifugation. XPS analysis of Fe3Cu1 before and after the reaction (Figure 4c,d) revealed a slight shift of Fe 2p to lower binding energies and Cu 2p to higher binding energies, showing good valence cycle stability.

3.3. Mechanism Investigation

To investigate the roles of different radicals in methane conversion and C1 oxygenate formation, tests were conducted by adding 0.5 mM of p-BQ and IPA, respectively (quenchers for •OH and •OOH, respectively). As shown in Figure 5a, methanol yield decreased significantly, along with a slight increase in CO2 production. CH3OOH yield rose after IPA addition, which was thought to be an intermediate in the formation of methanol [39]. These results demonstrate that both •OH and •OOH are crucial radicals in the methane-to-methanol conversion process. The H2O2 utilization efficiency is an important indicator of the catalyst on methane partial oxidation performance, which can be assessed by the “gain factor” and defined as the molar ratio of the produced C1 oxygenates to the consumed H2O2 [41]. The gain factor for bimetallic catalysts positively correlated with the Fe content, while monometallic Fe/ZSM-5 deviated from this trend (Figure 5b), suggesting synergistic interactions between Fe and Cu that enhance methane conversion. The Fe3Cu1 exhibited the highest gain factor (0.065), indicating superior H2O2 utilization. In contrast, Cu/ZSM-5 nearly consumed all the H2O2 but yielded the lowest C1 oxygenates, which may be due to CuO clusters on Cu/ZSM-5 directly catalyzing the decomposition of the excess H2O2 into H2O and O2, resulting in a decrease in its gain factor [10].
DMPO was used as a trapping agent to explore the differences and effects of free radicals under different catalyst systems. From Figure 5c, a weak DMPO-OH signal was observed on HZSM-5 but was significantly enhanced after Cu loading. Due to insufficient activation of methane, •CH3 was not enough to form CH3OH. At the same time, excess •OH also over-oxidized CH3OH to produce more HCHO. In contrast to the monometallic Cu/ZSM-5, •CH3 and •OH were detected on Fe/ZSM-5, and Fe-Cu interactions enhanced the •CH3 radical formation. The Fe/Cu ratio strongly influenced radical generation: higher Cu content increased •OH levels, while Fe-Cu synergy and dual redox cycling significantly raised •OH production. Therefore, Fe1Cu3 in the bimetallic catalyst presented the highest •OH content, and an insignificant •OOH signal was also observed. Its CH3OH and HCHO yields resembled those of Cu/ZSM-5. The •CH3 and •OH can quickly combine to form CH3OH, and since Fe3Cu1 had the highest •CH3 and appropriate •OH, it showed the highest CH3OH yield.
To investigate the mechanism of direct catalytic oxidation of methane, in situ CH4-DRIFTS was performed. The catalysts were exposed to H2O2 and CH4 at 70 °C, and the spectra were recorded over 60 min. Figure 6a shows the time-dependent changes in the in situ CH4-DRIFTS spectrum of Fe3Cu1. The peaks at 3614, 3401, and 3214 cm−1 are assigned to *OH2, *OH, and *OOH [4,42]; the peaks at 3016 cm−1, 1305 cm−1, and 1353 cm−1 are attributed to CH4, stretching of *CH3, and *OCH3 [42,43,44]; the peak at 1643 cm−1 is assigned to adsorbed H2O* [44,45]; and 1837cm−1 is assigned to Fe(Ⅲ)-(OH) species [46], respectively. After introducing methane, an *OH vibration peak and a gradually increasing *CH3 peak appeared with reaction time, indicating that *OH formed after the decomposition of H2O2 can activate the C-H bond and thus generate the adsorbed *CH3 species. The source of *OH2 may be formed by the combination of *OH and H decomposed by CH4. The appearance of *OCH3, *OH, and *OOH peaks indicates the formation of CH3OH and CH3OOH. The intensities of *OH and *CH3 peaks increased over time, while the HCOO- peak intensity remained constant, demonstrating its good methane oxidation capability and effective suppression of deep methanol oxidation. Figure 6b shows the time-dependent changes in the in situ CH4-DRIFTS spectrum of Cu/ZSM-5. The weaker *OH2, *OH, and *OOH peaks on Cu/ZSM-5 were observed, consistent with their lower methane oxidation performance. Formic acid species are the oxidation products of methanol, but the stretching vibration peak of formic acid species (HCOO-) at 1620 cm−1 was not observed at Cu/ZSM-5 [47], indicating a weak oxidizing capacity.
CO-DRIFTS was employed to analyze the reduction process of Cu2+ on the catalyst. CO can be adsorbed on Cu+ sites to form stable Cu+-CO species without binding to Cu2+. The catalyst was treated with Ar at 300 °C, cooled to 50 °C, exposed to CO for 30 min, and then purged with Ar for 30 min. Figure S6 illustrates the spectra of Cu-containing catalysts after Ar purging. The peak at 2157 cm−1 is assigned to the Cu+-CO species on the isolated Cu+ site, while the weaker absorption peak at 2137 cm−1 is assigned to the Cu+-CO-Cu+ species on the dimeric Cu+ site. Fe3Cu1 exhibited the weakest peak at 2137 cm−1 among the catalysts, implying the lowest degree of agglomeration of CuOx [48]. This is consistent with the XPS results and shows that different Fe/Cu ratios affect the valence and state of Cu species. Figure S7 shows the time-dependent CO-DRIFTS spectrum of Fe1Cu3. As the Ar purge time increased, the peak at 2177 cm−1, caused by the polarization of CO by acidic hydroxyls of the zeolites [49], gradually disappeared. In contrast, the peaks at 2157 cm−1 and 2137 cm−1 gradually increased.
Based on the above results and analysis, the reaction path of methane on as-prepared catalysts is shown in Scheme 1. Fe(Ⅲ) species were oxidized by H2O2 to [Fe(Ⅳ)-O•] species, and the C-H bond broke by a homolytic mechanism on the Fe(IV) species to form •CH3, and the [Fe(IV)-O•] species were reduced to [Fe(III)-OH] species (Equation (1)) [32]. Then, •OOH and •OH were generated by the Fenton-like reaction of Fe2+/Fe3+ cycling and H2O2 (Equations (2) and (3)), and similarly on Cu+/Cu2+ (Equations (4) and (5)). The •CH3 could also be generated by the •OH and CH4 reaction (Equation (6)); then, it reacted with •OH and •OOH to produce CH3OH and CH3OOH, respectively (Equations (7) and (8)). CH3OH and CH3OOH were deeply oxidized by •OH/•OOH/H2O2 to form HCHO, CO2, and H2O (Equation (9)).
Fe(Ⅲ) + H2O2 + CH4 → Fe(Ⅲ)-OH + •CH3 + H2O
Fe(Ⅱ) + H2O2 → Fe(Ⅲ) + •OH + OH
Fe(Ⅲ) + H2O2 → Fe(Ⅱ) + •OOH + H+
Cu(Ⅱ) + OH + H2O2 → •OOH + Cu(Ⅰ) + H2O
Cu(Ⅰ) + H+ + H2O2 → •OH + Cu(Ⅱ) + H2O
•OH + CH4 → •CH3 + H2O
•CH3 + •OH → CH3OH
•CH3 + •OOH → CH3OOH
CH3OH/CH3OOH + •OH/H2O2 → HCHO → CO2 + H2O

4. Conclusions

In summary, Fe- and Cu-modified ZSM-5 catalysts were synthesized via incipient wetness impregnation and evaluated for the direct oxidation of methane to methanol with H2O2 as oxidant. The Fe3Cu1 exhibited the highest H2O2 utilization efficiency, demonstrating optimal performance on methanol yield of 29.7 mmol gcat−1 h−1 with a selectivity of 80.9% at 70 °C. In comparison with state-of-the-art catalysts for methane-to-methanol conversion (Table S2), the Fe3Cu1 catalyst demonstrates competitive performance, highlighting its exceptional balance of activity and selectivity among non-noble metal catalysts. Characterization results revealed that the superior performance of Fe3Cu1 stems from several factors: I. abundant active sites due to high metal loading. II. Fe-Cu interactions preventing oxide cluster aggregation, which is detrimental to methane activation. III. an optimal Fe/Cu ratio maximizing Fe3+ and Cu+ content, facilitating a robust dual redox cycle. This cycle promotes methane activation to •CH3 at Fe sites while suppressing deep oxidation induced by the Fenton-like effect of Fe/H2O2, thereby maintaining high methanol selectivity. This work provides valuable insights for designing efficient catalysts for methane conversion under mild conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/surfaces8040069/s1. Figure S1: The CH4 conversion process and determination procedure; Figure S2: HRTEM images and elemental mappings of Fe/ZSM-5 (a), SEM image of HZSM-5 (b). Figure S3: Pore size distribution curves of catalysts; Figure S4: Proportion of Cu+/Cu2+ by XPS analysis; Figure S5: Effect of reaction pressure on the catalytic performance of Fe3Cu1; Figure S6: FTIR spectra of CO at 50 °C on different catalysts after desorbed by Ar for 30 min; Figure S7: FTIR spectra of CO at 50 °C on Fe3Cu1 at different time; Table S1: Iron and copper load obtained by ICP-OES; Table S2: Comparison of batch CH4 conversion performance with reported literature. Refs. [10,11,40,50,51,52,53,54,55,56,57] are cited in the supporting information.

Author Contributions

H.Z.: writing—original draft, investigation, and data curation. Y.P.: data curation. Y.L.: data curation. M.F.: writing—review and editing, validation, and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers: 51878293 and 22476054.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Hammond, C.; Forde, M.M.; Ab Rahim, M.H.; Thetford, A.; He, Q.; Jenkins, R.L.; Dimitratos, N.; Lopez-Sanchez, J.A.; Dummer, N.F.; Murphy, D.M.; et al. Direct Catalytic Conversion of Methane to Methanol in an Aqueous Medium by using Copper-Promoted Fe-ZSM-5. Angew. Chem. Int. Ed. 2012, 51, 5129–5133. [Google Scholar] [CrossRef]
  2. Tang, X.; Wang, L.; Yang, B.; Fei, C.; Yao, T.Y.; Liu, W.; Lou, Y.; Dai, Q.G.; Cai, Y.F.; Cao, X.M.; et al. Direct oxidation of methane to oxygenates on supported single Cu atom catalyst. Appl. Catal. B Environ. 2021, 285, 119827. [Google Scholar] [CrossRef]
  3. Mao, Y.; Hu, P. Identification of the active sites and mechanism for partial methane oxidation to methanol over copper-exchanged CHA zeolites. Sci. China Chem. 2020, 63, 850–859. [Google Scholar] [CrossRef]
  4. Yu, X.; Wu, B.; Huang, M.; Lu, Z.X.; Li, J.; Zhong, L.S.; Sun, Y.H. IrFe/ZSM-5 Synergistic Catalyst for Selective Oxidation of Methane to Formic Acid. Energy Fuel 2021, 35, 4418–4427. [Google Scholar] [CrossRef]
  5. Wang, B.; Albarracín-Suazo, S.; Pagán-Torres, Y.; Nikolla, E. Advances in methane conversion processes. Catal. Today 2017, 285, 147–158. [Google Scholar] [CrossRef]
  6. Li, R.J.; Yu, C.C.; Dai, X.P.; Shen, S.K. Selective oxidation of methane to synthesis gas using lattice oxygen from perovskite LaSrFeO catalyst. Chin. J. Catal. 2002, 23, 549–554. [Google Scholar] [CrossRef]
  7. Meng, X.G.; Cui, X.J.; Rajan, N.P.; Yu, L.; Deng, D.H.; Bao, X.H. Direct Methane Conversion under Mild Condition by Thermo-, Electro-, or Photocatalysis. Chem 2019, 5, 2296–2325. [Google Scholar] [CrossRef]
  8. Liu, C.-C.; Mou, C.-Y.; Yu, S.S.F.; Chan, S.I. Heterogeneous formulation of the tricopper complex for efficient catalytic conversion of methane into methanol at ambient temperature and pressure. Energy Environ. Sci. 2016, 9, 1361–1374. [Google Scholar] [CrossRef]
  9. Hammond, C.; Dimitratos, N.; Lopez-Sanchez, J.A.; Jenkins, R.L.; Whiting, G.; Kondratt, S.A.; ab Rahim, M.H.; Forde, M.M.; Thetford, A.; Hagen, H.; et al. Aqueous-Phase Methane Oxidation over Fe-MFI Zeolites; Promotion through Isomorphous Framework Substitution. ACS Catal. 2013, 3, 1835–1844. [Google Scholar] [CrossRef]
  10. Huang, W.; Zhang, S.; Tang, Y.; Li, Y.; Nguyen, L.; Li, Y.; Shan, J.; Xiao, D.; Gagne, R.; Frenkel, A.I.; et al. Low-Temperature Transformation of Methane to Methanol on Pd1 O4 Single Sites Anchored on the Internal Surface of Microporous Silicate. Angew. Chem. Int. Ed. 2016, 55, 13441–13445. [Google Scholar] [CrossRef] [PubMed]
  11. Agarwal, N.; Freakley, S.J.; McVicker, R.U.; Althahban, S.M.; Dimitratos, N.; He, Q.; Morgan, D.J.; Jenkins, R.L.; Willock, D.J.; Taylor, S.H.; et al. Aqueous Au-Pd colloids catalyze selective CH4 oxidation to CHOH with O2 under mild conditions. Science 2017, 358, 223–226. [Google Scholar] [CrossRef]
  12. Cheng, Q.P.; Li, G.N.; Yao, X.L.; Zheng, L.R.; Wang, J.H.; Emwas, A.H.; Castaño, P.; Ruiz-Martínez, J.; Han, Y. Maximizing Active Fe Species in ZSM-5 Zeolite Using Organic-Template-Free Synthesis for Efficient Selective Methane Oxidation. J. Am. Chem. Soc. 2023, 145, 5888–5898. [Google Scholar] [CrossRef]
  13. He, M.; Zhang, J.; Sun, X.L.; Chen, B.H.; Wang, Y.G. Theoretical Study on Methane Oxidation Catalyzed by Fe/ZSM-5: The Significant Role of Water on Binuclear Iron Active Sites. J. Phys. Chem. C 2016, 120, 27422–27429. [Google Scholar] [CrossRef]
  14. Yu, T.; Li, Z.; Jones, W.; Liu, Y.; He, Q.; Song, W.; Du, P.; Yang, B.; An, H.; Farmer, D.M.; et al. Identifying key mononuclear Fe species for low-temperature methane oxidation. Chem. Sci. 2021, 12, 3152–3160. [Google Scholar] [CrossRef]
  15. Dinh, K.T.; Sullivan, M.M.; Serna, P.; Meyer, R.J.; Dinca, M.; Román-Leshkov, Y. Viewpoint on the Partial Oxidation of Methane to Methanol Using Cu- and Fe-Exchanged Zeolites. ACS Catal. 2018, 8, 8306–8313. [Google Scholar] [CrossRef]
  16. Battiston, A.A.; Bitter, J.H.; de Groot, F.M.F.; Overweg, A.R.; Stephan, O.; van Bokhoven, J.A.; Kooyman, P.J.; van der Spek, C.; Vankó, G.; Koningsberger, D.C. Evolution of Fe species during the synthesis of over-exchanged Fe/ZSM5 obtained by chemical vapor deposition of FeCl3. J. Catal. 2003, 213, 251–271. [Google Scholar] [CrossRef]
  17. Freakley, S.J.; Dimitratos, N.; Willock, D.J.; Taylor, S.H.; Kiely, C.J.; Hutchings, G.J. Methane Oxidation to Methanol in Water. Acc. Chem. Res. 2021, 54, 2614–2623. [Google Scholar] [CrossRef]
  18. Dinh, K.T.; Sullivan, M.M.; Narsimhan, K.; Serna, P.; Meyer, R.J.; Dinca, M.; Román-Leshkov, Y. Continuous Partial Oxidation of Methane to Methanol Catalyzed by Diffusion-Paired Copper Dimers in Copper-Exchanged Zeolites. J. Am. Chem. Soc. 2019, 141, 11641–11650. [Google Scholar] [CrossRef] [PubMed]
  19. Yu, T.; Li, Z.; Lin, L.; Chu, S.Q.; Su, Y.; Song, W.Y.; Wang, A.Q.; Weckhuysen, B.M.; Luo, W.H. Highly Selective Oxidation of Methane into Methanol over Cu-Promoted Monomeric Fe/ZSM-5. ACS Catal. 2021, 11, 6684–6691. [Google Scholar] [CrossRef]
  20. Yue, H.; Zhao, Y.; Zhao, S.; Wang, B.; Gong, J. A copper-phyllosilicate core-sheath nanoreactor for carbon–oxygen hydrogenolysis reactions. Nat. Commun. 2013, 4, 2339. [Google Scholar] [CrossRef] [PubMed]
  21. Lee, H.; Seong, J.; Lee, K.-M.; Kim, H.-H.; Choi, J.; Kim, J.-H.; Lee, C. Chloride-enhanced oxidation of organic contaminants by Cu(II)-catalyzed Fenton-like reaction at neutral pH. J. Hazard. Mater. 2018, 344, 1174–1180. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, J.; Zhao, W.; Wu, S.; Yin, R.; Zhu, M. Surface dual redox cycles of Mn(III)/Mn(IV) and Cu(I)/Cu(II) for heterogeneous peroxymonosulfate activation to degrade diclofenac: Performance, mechanism and toxicity assessment. J. Hazard. Mater. 2021, 410, 124623. [Google Scholar] [CrossRef]
  23. Luo, L.; Han, X.; Wang, K.; Xu, Y.; Xiong, L.; Ma, J.; Guo, Z.; Tang, J. Nearly 100% selective and visible-light-driven methane conversion to formaldehyde via. single-atom Cu and Wδ+. Nat. Commun. 2023, 14, 2690. [Google Scholar] [CrossRef] [PubMed]
  24. Tian, J.; Tan, J.; Zhang, Z.; Han, P.; Yin, M.; Wan, S.; Lin, J.; Wang, S.; Wang, Y. Direct conversion of methane to formaldehyde and CO on B2O3 catalysts. Nat. Commun. 2020, 11, 5693. [Google Scholar] [CrossRef]
  25. Zhang, L.; Lin, Y. Facile synthesis of N-doped carbon supported iron species for highly efficient methane conversion with H2O2 at ambient temperature. Appl. Catal. A Gen. 2021, 615, 118052. [Google Scholar] [CrossRef]
  26. Delahay, G.; Valade, D.; Guzmán-Vargas, A.; Coq, B. Selective catalytic reduction of nitric oxide with ammonia on Fe-ZSM-5 catalysts prepared by different methods. Appl. Catal. B Environ. 2005, 55, 149–155. [Google Scholar] [CrossRef]
  27. Jianguo, L.; Mingyue, D.; Tiejun, W.; Longlong, M. Structure and Performance of Cu-Fe Bimodal Support for Higher Alcohol Syntheses. Acta Phys.-Chim. Sin. 2012, 28, 1964–1970. [Google Scholar] [CrossRef]
  28. Zhang, Y.-B.; Wang, P.; Yu, D.; Zhao, H.-Y.; Lyu, X.-L.; Lei, L.-L. Evolution mechanism of active sites for selective catalytic reduction of NOx with NH3 over Fe-ZSM-5 catalysts doped by Ce/Cu. J. Cent. South Univ. 2022, 29, 2239–2252. [Google Scholar] [CrossRef]
  29. Hsu, Y.Y.; Suen, N.T.; Chang, C.C.; Hung, S.F.; Chen, C.L.; Chan, T.S.; Dong, C.L.; Chan, C.C.; Chen, S.Y.; Chen, H.M. Heterojunction of Zinc Blende/Wurtzite in ZnCdS Solid Solution for Efficient Solar Hydrogen Generation: X-ray Absorption/Diffraction Approaches. Acs Appl. Mater. Interfaces 2015, 7, 22558–22569. [Google Scholar] [CrossRef] [PubMed]
  30. Bang, Y.; Han, S.J.; Yoo, J.; Choi, J.H.; Lee, J.K.; Song, J.H.; Lee, J.; Song, I.K. Hydrogen production by steam reforming of simulated liquefied natural gas (LNG) over nickel catalyst supported on mesoporous phosphorus-modified alumina xerogel. Appl. Catal. B Environ. 2014, 148–149, 269–280. [Google Scholar] [CrossRef]
  31. Grosvenor, A.P.; Kobe, B.A.; Biesinger, M.C.; McIntyre, N.S. Investigation of multiplet splitting of Fe 2p XPS spectra and bonding in iron compounds. Surf. Interface Anal. 2004, 36, 1564–1574. [Google Scholar] [CrossRef]
  32. Szécsényi, Á.; Li, G.; Gascon, J.; Pidko, E.A. Mechanistic Complexity of Methane Oxidation with H2O2 by Single-Site Fe/ZSM-5 Catalyst. ACS Catal. 2018, 8, 7961–7972. [Google Scholar] [CrossRef]
  33. Xing, X.; Li, N.; Cheng, J.; Sun, Y.G.; Zhang, Z.S.; Zhang, X.; Hao, Z.P. Synergistic effects of Cu species and acidity of Cu-ZSM-5 on catalytic performance for selective catalytic oxidation of butylamine. J. Environ. Sci. 2020, 96, 55–63. [Google Scholar] [CrossRef]
  34. Dou, B.J.; Li, S.M.; Liu, D.L.; Zhao, R.Z.; Liu, J.G.; Hao, Q.L.; Bin, F. Catalytic oxidation of ethyl acetate and toluene over Cu-Ce-Zr supported ZSM-5/TiO2 catalysts. RSC Adv. 2016, 6, 53852–53859. [Google Scholar] [CrossRef]
  35. Li, S.M.; Hao, Q.L.; Zhao, R.Z.; Liu, D.L.; Duan, H.Z.; Dou, B.J. Highly efficient catalytic removal of ethyl acetate over Ce/Zr promoted copper/ZSM-5 catalysts. Chem. Eng. J. 2016, 285, 536–543. [Google Scholar] [CrossRef]
  36. Zhu, Z.; Liu, Z.; Liu, S.; Niu, H.; Hu, T.; Liu, T.; Xie, Y. NO reduction with NH3 over an activated carbon-supported copper oxide catalysts at low temperatures. Appl. Catal. B Environ. 2000, 26, 25–35. [Google Scholar] [CrossRef]
  37. Dasireddy, V.D.B.C.; Likozar, B. The role of copper oxidation state in Cu/ZnO/Al2O3 catalysts in CO2 hydrogenation and methanol productivity. Renew. Energy 2019, 140, 452–460. [Google Scholar] [CrossRef]
  38. Forde, M.M.; Armstrong, R.D.; McVicker, R.; Wells, P.P.; Dimitratos, N.; He, Q.; Lu, L.; Jenkins, R.L.; Hammond, C.; Lopez-Sanchez, J.A.; et al. Light alkane oxidation using catalysts prepared by chemical vapour impregnation: Tuning alcohol selectivity through catalyst pre-treatment. Chem. Sci. 2014, 5, 3603–3616. [Google Scholar] [CrossRef]
  39. Taran, O.P.; Yashnik, S.A.; Boltenkov, V.V.; Parkhomchuk, E.V.; Sashkina, K.A.; Ayusheev, A.B.; Babushkin, D.E.; Parmon, V.N. Formic Acid Production Via Methane Peroxide Oxidation Over Oxalic Acid Activated Fe-MFI Catalysts. Top. Catal. 2019, 62, 491–507. [Google Scholar] [CrossRef]
  40. Oda, A.; Aono, K.; Murata, N.; Murata, K.; Yasumoto, M.; Tsunoji, N.; Sawabe, K.; Satsuma, A. Rational design of ZSM-5 zeolite containing a high concentration of single Fe sites capable of catalyzing the partial oxidation of methane with high turnover frequency. Catal. Sci. Technol. 2022, 12, 542–550. [Google Scholar] [CrossRef]
  41. Xie, P.F.; Ding, J.; Yao, Z.H.; Pu, T.C.; Zhang, P.; Huang, Z.N.; Wang, C.H.; Zhang, J.L.; Zecher-Freeman, N.; Zong, H.; et al. Oxo dicopper anchored on carbon nitride for selective oxidation of methane. Nat. Commun. 2022, 13, 1375. [Google Scholar] [CrossRef]
  42. Bai, S.X.; Liu, F.F.; Huang, B.L.; Li, F.; Lin, H.P.; Wu, T.; Sun, M.Z.; Wu, J.B.; Shao, Q.; Xu, Y.; et al. High-efficiency direct methane conversion to oxygenates on a cerium dioxide nanowires supported rhodium single-atom catalyst. Nat. Commun. 2020, 11, 954. [Google Scholar] [CrossRef] [PubMed]
  43. Li, B.; Song, X.G.; Feng, S.Q.; Yuan, Q.; Jiang, M.; Yan, L.; Ding, Y.J. Direct conversion of methane to oxygenates on porous organic polymers supported Rh mononuclear complex catalyst under mild conditions. Appl. Catal. B Environ. 2021, 293, 120208. [Google Scholar] [CrossRef]
  44. Liu, B.; Huang, M.Y.; Fang, Z.H.; Kong, L.; Xu, Y.B.; Li, Z.J.; Liu, X.H. Breaking the scaling relationship in selective oxidation of methane via dynamic Metal-Intermediate Coordination-Induced modulation of reactivity descriptors on an atomically dispersed Rh/ZrO catalyst. J. Catal. 2022, 416, 68–84. [Google Scholar] [CrossRef]
  45. Artsiusheuski, M.A.; Verel, R.; van Bokhoven, J.A.; Sushkevich, V.L. Methane Transformation over Copper-Exchanged Zeolites: From Partial Oxidation to C-C Coupling and Formation of Hydrocarbons. ACS Catal. 2021, 11, 12543–12556. [Google Scholar] [CrossRef]
  46. Fellah, M.F. CO and NO Adsorptions on Different Iron Sites of Fe-ZSM-5 Clusters: A Density Functional Theory Study. J. Phys. Chem. C 2011, 115, 1940–1951. [Google Scholar] [CrossRef]
  47. Marwood, M.; Doepper, R.; Renken, A. In-situ surface and gas phase analysis for kinetic studies under transient conditions—The catalytic hydrogenation of CO2. Appl. Catal. A Gen. 1997, 151, 223–246. [Google Scholar] [CrossRef]
  48. Wang, L.; Li, W.; Qi, G.S.; Weng, D. Location and nature of Cu species in Cu/SAPO-34 for selective catalytic reduction of NO with NH3. J. Catal. 2012, 289, 21–29. [Google Scholar] [CrossRef]
  49. Mihaylov, M.; Ivanova, E.; Chakarova, K.; Novachka, P.; Hadjiivanov, K. Reduced iron sites in Fe–BEA and Fe–ZSM-5 zeolites: FTIR study of CO adsorption and 12C16O–13C18O co-adsorption. Appl. Catal. A Gen. 2011, 391, 3–10. [Google Scholar] [CrossRef]
  50. Ab Rahim, M.H.; Forde, M.M.; Jenkins, R.L.; Hammond, C.; He, Q.; Dimitratos, N.; Lopez-Sanchez, J.A.; Carley, A.F.; Taylor, S.H.; Willock, D.J.; et al. Oxidation of Methane to Methanol with Hydrogen Peroxide Using Supported Gold–Palladium Alloy Nanoparticles. Angew. Chem. Int. Ed. 2013, 52, 1280–1284. [Google Scholar] [CrossRef]
  51. Williams, C.; Carter, J.H.; Dummer, N.F.; Chow, Y.K.; Morgan, D.J.; Yacob, S.; Serna, P.; Willock, D.J.; Meyer, R.J.; Taylor, S.H.; et al. Selective Oxidation of Methane to Methanol Using Supported AuPd Catalysts Prepared by Stabilizer-Free Sol-Immobilization. ACS Catal. 2018, 8, 2567–2576. [Google Scholar] [CrossRef]
  52. Kwon, Y.; Kim, T.Y.; Kwon, G.; Yi, J.; Lee, H. Selective Activation of Methane on Single-Atom Catalyst of Rhodium Dispersed on Zirconia for Direct Conversion. J. Am. Chem. Soc. 2017, 139, 17694–17699. [Google Scholar] [CrossRef]
  53. Fang, Z.; Murayama, H.; Zhao, Q.; Liu, B.; Jiang, F.; Xu, Y.; Tokunaga, M.; Liu, X. Selective mild oxidation of methane to methanol or formic acid on Fe–MOR catalysts. Catal. Sci. Technol. 2019, 9, 6946–6956. [Google Scholar] [CrossRef]
  54. Wang, L.; Jin, J.; Li, W.; Li, C.; Zhu, L.; Zhou, Z.; Zhang, L.; Zhang, X.; Yuan, L. Highly selective catalytic oxidation of methane to methanol using Cu–Pd/anatase. Energy Environ. Sci. 2024, 17, 9122–9133. [Google Scholar] [CrossRef]
  55. Yu, B.; Cheng, L.; Dai, S.; Jiang, Y.; Yang, B.; Li, H.; Zhao, Y.; Xu, J.; Zhang, Y.; Pan, C.; et al. Silver and Copper Dual Single Atoms Boosting Direct Oxidation of Methane to Methanol via Synergistic Catalysis. Adv. Sci. 2023, 10, 2302143. [Google Scholar] [CrossRef]
  56. Antil, N.; Chauhan, M.; Akhtar, N.; Newar, R.; Begum, W.; Malik, J.; Manna, K. Metal–Organic Framework-Encaged Monomeric Cobalt(III) Hydroperoxides Enable Chemoselective Methane Oxidation to Methanol. ACS Catal. 2022, 12, 11159–11168. [Google Scholar] [CrossRef]
  57. Cui, X.; Li, H.; Wang, Y.; Hu, Y.; Hua, L.; Li, H.; Han, X.; Liu, Q.; Yang, F.; He, L.; et al. Room-Temperature Methane Conversion by Graphene-Confined Single Iron Atoms. Chem 2018, 4, 1902–1910. [Google Scholar] [CrossRef]
Figure 1. HRTEM images and elemental mappings of Fe3Cu1 (a), Fe1Cu1 (b), XRD patterns of different catalysts (c), and N2 adsorption–desorption isotherms of catalysts (d).
Figure 1. HRTEM images and elemental mappings of Fe3Cu1 (a), Fe1Cu1 (b), XRD patterns of different catalysts (c), and N2 adsorption–desorption isotherms of catalysts (d).
Surfaces 08 00069 g001
Figure 2. CH4-TPD curves of different catalysts, monitored by a thermal conductivity detector (TCD) (a), XPS spectrum of Fe 2p (b), Cu 2p (c), surface proportion of Fe2+ and Fe3+ by XPS analysis (d), and UV-Vis DRS spectra (e).
Figure 2. CH4-TPD curves of different catalysts, monitored by a thermal conductivity detector (TCD) (a), XPS spectrum of Fe 2p (b), Cu 2p (c), surface proportion of Fe2+ and Fe3+ by XPS analysis (d), and UV-Vis DRS spectra (e).
Surfaces 08 00069 g002
Figure 3. Catalytic performance of different catalysts (a), effect of reaction temperature on the catalytic performance of Fe3Cu1 (b), effect of reaction time on the catalytic performance of Fe/ZSM-5 (c), and Fe3Cu1 (d). Reactions in (ad) were conducted in 20 mL H2O with 0.5 M H2O2 and 25 mg catalyst at 30 bars for 30 min (a,b), under a temperature of 70 °C (a,c,d).
Figure 3. Catalytic performance of different catalysts (a), effect of reaction temperature on the catalytic performance of Fe3Cu1 (b), effect of reaction time on the catalytic performance of Fe/ZSM-5 (c), and Fe3Cu1 (d). Reactions in (ad) were conducted in 20 mL H2O with 0.5 M H2O2 and 25 mg catalyst at 30 bars for 30 min (a,b), under a temperature of 70 °C (a,c,d).
Surfaces 08 00069 g003
Figure 4. Effect of H2O2 concentration on the catalytic performance of Fe3Cu1 (a), cycling catalytic performance of Fe3Cu1 (b), XPS analysis of Fe 2p (c), and Cu 2p (d) on the fresh and used Fe3Cu1.
Figure 4. Effect of H2O2 concentration on the catalytic performance of Fe3Cu1 (a), cycling catalytic performance of Fe3Cu1 (b), XPS analysis of Fe 2p (c), and Cu 2p (d) on the fresh and used Fe3Cu1.
Surfaces 08 00069 g004
Figure 5. Effects of IPA and p-BQ on catalytic performance of Fe3Cu1 (a), correlation between productivity of C1 oxygenates and gain factor for different catalysts (b), and EPR spectra of different catalysts (c).
Figure 5. Effects of IPA and p-BQ on catalytic performance of Fe3Cu1 (a), correlation between productivity of C1 oxygenates and gain factor for different catalysts (b), and EPR spectra of different catalysts (c).
Surfaces 08 00069 g005
Figure 6. CH4-DRIFTS spectra of Fe3Cu1 at 70 °C after exposure to CH4 and H2O2 for 60 min (a) and Cu/ZSM-5 (b).
Figure 6. CH4-DRIFTS spectra of Fe3Cu1 at 70 °C after exposure to CH4 and H2O2 for 60 min (a) and Cu/ZSM-5 (b).
Surfaces 08 00069 g006
Scheme 1. Illustrated mechanism for the selective oxidation of methane over Fe-Cu/ZSM-5.
Scheme 1. Illustrated mechanism for the selective oxidation of methane over Fe-Cu/ZSM-5.
Surfaces 08 00069 sch001
Table 1. Specific surface area and pore volume of catalysts.
Table 1. Specific surface area and pore volume of catalysts.
CatalystSurface Area (m2/g)t-Plot Micropore Volume (cm3/g)Pore Diameter (nm)
Fe/ZSM-5310.200.111.82
Fe3Cu1261.640.101.92
Fe1Cu1272.620.101.87
Fe1Cu3286.020.111.85
Cu/ZSM-5298.930.121.88
HZSM-5391.170.131.87
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Pu, Y.; Li, Y.; Fu, M. Highly Efficient Conversion of Methane to Methanol on Fe-Cu/ZSM-5 Under Mild Conditions: Effective Utilization of Free Radicals by Favorable Valence Ratios. Surfaces 2025, 8, 69. https://doi.org/10.3390/surfaces8040069

AMA Style

Zhang H, Pu Y, Li Y, Fu M. Highly Efficient Conversion of Methane to Methanol on Fe-Cu/ZSM-5 Under Mild Conditions: Effective Utilization of Free Radicals by Favorable Valence Ratios. Surfaces. 2025; 8(4):69. https://doi.org/10.3390/surfaces8040069

Chicago/Turabian Style

Zhang, Huajie, Yunhan Pu, Yanjun Li, and Mingli Fu. 2025. "Highly Efficient Conversion of Methane to Methanol on Fe-Cu/ZSM-5 Under Mild Conditions: Effective Utilization of Free Radicals by Favorable Valence Ratios" Surfaces 8, no. 4: 69. https://doi.org/10.3390/surfaces8040069

APA Style

Zhang, H., Pu, Y., Li, Y., & Fu, M. (2025). Highly Efficient Conversion of Methane to Methanol on Fe-Cu/ZSM-5 Under Mild Conditions: Effective Utilization of Free Radicals by Favorable Valence Ratios. Surfaces, 8(4), 69. https://doi.org/10.3390/surfaces8040069

Article Metrics

Back to TopTop