Next Article in Journal
New Trends in Catalytic Reaction for High-Temperature and Low-Emission Combustion Technologies
Previous Article in Journal
Theme Issue in Memory to Professor Jiro Tsuji (1927–2022)
Previous Article in Special Issue
Diphenyl Carbonate: Recent Progress on Its Catalytic Synthesis by Transesterification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Growth of Mn-Co3O4 on Mesoporous ZSM-5 Zeolite for Boosting Lean Methane Catalytic Oxidation

1
School of Materials Science and Engineering, Northeastern University, Shenyang 110189, China
2
Key Laboratory of Dielectric and Electrolyte Functional Material Hebei Province, Northeastern University at Qinhuangdao, Qinhuangdao 066004, China
3
State Key Laboratory of Biochemical Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
4
School of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(7), 397; https://doi.org/10.3390/catal14070397
Submission received: 9 May 2024 / Revised: 18 June 2024 / Accepted: 21 June 2024 / Published: 23 June 2024
(This article belongs to the Special Issue Feature Papers in "Industrial Catalysis" Section)

Abstract

:
The low-temperature oxidation of methane gas in coal mine exhaust gas is important for reducing the greenhouse effect and protecting the environment. Unfortunately, the carbon–hydrogen bonds in methane molecules are highly stable, requiring higher reaction temperatures to achieve effective catalytic oxidation. However, metal oxide-based catalysts face the problem of easy sintering and the deactivation of active components at high temperatures, which is an important challenge that catalysts need to overcome in practical applications. In this work, a series of Mn-Co3O4 active components were grown in situ on ZSM-5 zeolite with mesoporous pore structures treated with an alkaline solution via a hydrothermal synthesis method. Due to the presence of polyethylene glycol as a structure-directing agent, manganese can be uniformly doped into the Co3O4 lattice. The large specific surface area of ZSM-5 zeolite allows the active component Mn-Co3O4 to be uniformly dispersed, effectively preventing the sintering and growth of active component particles during the catalytic reaction process. It is worth mentioning that the Mn-Co3O4/meso-ZSM-5-6.67 catalyst has a methane conversion rate of up to 90% at a space velocity of 36,000 mL·g−1·h−1 and a reaction temperature of 363 °C. This is mainly due to the mesoporous ZSM-5 carrier with a high specific surface area, which is conducive to the adsorption and mass transfer of reaction molecules. The active component has an abundance of oxygen vacancies, which is conducive to the activation of reaction molecules and enhances its catalytic activity, which is even higher than that of noble metal-based catalysts. The new ideas for the preparation of metal oxide-based low-temperature methane oxidation catalysts proposed in this work are expected to provide new solutions for low-temperature methane oxidation reactions and promote technological progress in related fields.

1. Introduction

The gas that leaks from gas mines contains approximately 0.1% to 1% low-concentration methane. It has been estimated that these leaked methane gases account for approximately 70% of the total methane gas present in the atmosphere [1]. Methane contributes 21 times more to the greenhouse effect than the same volume of carbon dioxide. Its atmospheric lifespan is 12 ± 3 years [2], which poses a significant threat to the natural environment. The challenge of methane combustion lies in the activation of strong C-H bonds in methane and the subsequent combination of this step with activated oxygen. Methane has a stable tetrahedral structure with an intramolecular C-H bond energy of up to 415.48 kJ/mol [3], requiring higher temperatures during combustion. In order to reduce energy wastage during methane combustion and to improve the safety of the treatment process, researchers have started to study the factors and mechanisms influencing the combustion of methane at low concentrations in an attempt to achieve complete catalytic combustion of methane at lower temperatures (<600 °C) [4]. Among them, noble metal-based catalysts (such as Pd and Pt [5]) exhibit good catalytic activity in low-temperature methane catalytic combustion, but their large-scale preparation and application are limited due to their high cost and easy sintering deactivation during the reaction process.
It is worth noting that transition metal oxides have shown tremendous potential in this area. Transition metal oxides usually have rich valence state changes, which help to provide multiple active sites in catalytic reactions [6], thereby improving catalytic efficiency. The spinel structure of Co3O4 is the best alternative to precious metal catalysts. The metal-oxygen bond (M-O) of Co3O4 is weaker than that of other transition metal oxides, and the surface oxygen species activity is higher, allowing the Co3O4 catalyst to catalyze methane efficiently at low temperatures. For Co3O4, researchers have found a more significant grain size effect compared to other transition metal oxide catalysts, meaning that as the catalyst grain size decreases, the methane conversion rate increases up to a certain limit and then decreases. Pandey et al. found that the catalytic activity of the catalyst is highest when the size of the Co3O4 particles is 2.3 nm, while the catalytic activity decreases when the size is below or above 2.3 nm. Therefore, by precisely adjusting the particle size, cobalt trioxide catalysts can achieve optimal catalytic performance and are more convenient to use than other transition metal oxide catalysts. Co3O4, as a low-cost and highly catalytic transition metal oxide, has received much attention from researchers in recent years [7].
The heterogeneous composite of Co3O4 with other oxides enhances the catalytic oxidation performance due to the highly dispersed active components caused by the interaction between oxide particles. Li et al. [8] proposed a combination of Co3O4 and SiAl@Al2O3, which was applied for the first time in the catalytic combustion of methane, improving the thermal environment of the catalyst bed and achieving high efficiency in the utilization of lean methane. G. Di et al. [9] prepared Co3O4/CeO2 composite oxides with different cobalt loadings. The addition of 70% CeO2 to the catalyst significantly improved its activity and long-term stability compared with pure Co3O4. The microstructure of the catalyst is also an important factor affecting its catalytic performance. Co3O4 nanorods containing a large number of exposed (110) crystal planes exhibit excellent catalytic activity against spherical particles mainly surrounded by (111) crystal planes and can catalyze methane at lower temperatures. This confirms that controlling the morphology of nanostructured cobalt oxides is beneficial for exposing more catalytic active sites [10]. A number of researchers have demonstrated that flower-shaped Co3O4 can fully expose specific crystal planes [11], thereby increasing the concentration of surface active oxygen species and promoting the catalytic effect on CH4.
Although Co3O4-based catalysts exhibit good catalytic activity in methane catalytic combustion, there are still some challenges that need to be overcome. For instance, there is a need to enhance its catalytic activity, stability, and anti-sintering performance. Additionally, there is a requirement to optimize the preparation process in order to facilitate large-scale production. Researchers are conducting in-depth research on these issues, including improvements to the formulation of catalysts, optimization of preparation conditions, and exploration of new synthesis methods. Guo et al. [12] adopted low-temperature phosphating technology for the first time to prepare a phosphorus-atom-doped Co3O4 nanowire array (P-Co3O4/NF) supported by foam nickel. In comparison to Co3O4, the electrocatalytic performance of P-Co3O4/NF has been significantly enhanced. Pu et al. [13] prepared Co3O4 catalysts under different pH environments using a simple precipitation method. At a pH of 9.0, the catalyst exhibited a high Oads/Olatt ratio, which enabled it to exhibit high catalytic efficiency at low temperatures. Furthermore, the catalyst prepared using Co(C2H3O2)2 as a precursor exhibited superior catalytic activity.
It is important to highlight that catalyst carriers play a pivotal role in methane catalytic combustion. (1) The carrier can enhance the catalytic activity of the catalyst by modifying its dispersion state and the degree of exposure of active components. For instance, the surface of the carrier is replete with active sites that can form robust interactions with the active components of the catalyst, thereby augmenting catalytic activity. (2) The carrier can enhance the mechanical strength and thermal stability of the catalyst, preventing it from sintering or agglomeration at high temperatures. By fixing the active components of the catalyst on the surface of the carrier, the migration and aggregation of catalyst particles can be prevented, thereby improving the heat resistance and stability of the catalyst. (3) The interaction between the carrier and the active components of the catalyst can influence the reaction pathway and product distribution, thereby facilitating the efficient conversion of methane to carbon dioxide and water. (4) The high dispersion of active component particles on the surface of the carrier results in a larger contact area between the catalytic active component particles and the reaction gas, thereby improving the utilization rate of the unit catalyst.
ZSM-5 zeolite exhibits a considerable specific surface area and a distinctive pore structure [14], in addition to remarkable thermal stability and acidic properties [15]. It also exhibits catalytic activity. Following the loading of the active catalyst, the catalytic activity of the composite catalyst will also be significantly enhanced. In a study by Fei et al. [16], core–shell structured Co3O4-ZSM-5 composite catalysts were synthesized using different hydrothermal methods, and their catalytic performance on dichloromethane was subsequently evaluated. The results demonstrated that the catalysts prepared via the microwave hydrothermal method exhibited a higher proportion of Co3+/CO2+, superior oxygen mobility, and superior catalytic performance. Furthermore, the synthesis of the Co3O4/ZSM-5 core–shell catalyst was enhanced by the introduction of additional elements, including Cr, Ce, Nb, and Mn [17]. The incorporation of Cr2O3, generated through the addition of Cr, was found to interact with Co3O4, resulting in a high Co3+/CO2+ ratio and enhanced catalytic activity. In a separate study, Li et al. [18] prepared a series of ZSM-5/SBA-15 composite-supported Co catalysts and evaluated their suitability for Fischer–Tropsch synthesis. At a ZSM-5 loading of 20%, Co3O4 exhibited high dispersion and the most effective catalytic activity on CH4.
To sum up, the Co3O4-catalysed methane catalytic combustion reaction is significantly promoted by the catalyst’s high specific surface area, high Co3+/Co2+ ratio, and high surface active oxygen concentration. Nevertheless, the current Co3O4 particles are susceptible to deactivation due to sintering in catalytic reactions. In this work, a series of Mn-doped Co3O4 particles were synthesized and grown in situ on ZSM-5 zeolite with a mesoporous pore structure, which had been treated with an alkaline solution in order to address the issue of the facile sintering of active components. The large specific surface area of ZSM-5 zeolite facilitates the uniform dispersion of the active component Mn-Co3O4. The Mn-Co3O4/ZSM-5-6.67 catalyst exhibited a methane conversion rate of 90% at a space velocity of 36,000 mL·g−1·h−1 and a reaction temperature of 363 °C. The principal reason for this is that the mesoporous ZSM-5 carrier with a high specific surface area is conducive to an enhanced methane conversion rate. The adsorption and mass transfer of reaction molecules results in the formation of abundant oxygen vacancies in the active components, which is beneficial for the activation of reaction molecules and enhances their catalytic activity.

2. Results and Discussion

The preparation process of the Mn-Co3O4/meso-ZSM-5-6.67 is illustrated in Scheme 1. Firstly, dissolve Co(NO3)2 and Mn(CH3COOH)2 in deionized water to obtain a solution, and then add ZSM-5, which has been treated with an alkaline solution. Secondly, dissolve NH4HCO3 in deionized water to obtain a solution, and then mix the above two solutions evenly. In the following hydrothermal process, the Co element combines with oxygen species to enable CoCO3 to grow in situ on the surface of meso-ZSM-5 zeolite. Subsequently, the final catalyst Mn-Co3O4/meso-ZSM-5-6.67 was obtained through calcination. Figure 1 illustrates the phase composition between the hydrothermal reaction and calcination stages during the preparation of the meso-ZSM-5 supported Mn-Co3O4. The products obtained from the hydrothermal reaction at 170 °C exhibited distinct characteristic diffraction peaks at 24.83°, 32.35°, 38.37°, 42.41°, 46.30°, and 53.28°, corresponding to the (012), (104), (110), (113), (202) and (116) crystal planes, respectively. This is consistent with the diffraction peak of standard CoCO3 (PDF # 11-0692), but there is a certain shift to the left, which is caused by manganese doping. The ionic radius of Mn3+ (0.065 nm) is greater than that of Co3+ (0.061 nm), which results in some Mn3+ occupying the position of Co3+ in CoCO3. This leads to an increase in the lattice spacing. According to the Bragg formula 2dsinθ = nλ, it can be seen that as the interplanar spacing increases, the diffraction angle 2θ decreases. Consequently, the diffraction peak shifts to the left.
The calcined products exhibit distinct characteristic diffraction peaks at 19.32°, 31.12°, 36.76°, 44.96°, 55.43°, 59.17°, and 64.98°, corresponding to (111), (220), (331), (400), (422), (511), and (440) crystal planes, respectively. These peaks are consistent with the diffraction peaks of standard Co3O4 (PDF # 42-1467).
The XRD patterns of Mn-Co3O4/meso-ZSM-5 catalysts with varying amounts of ZSM-5 addition are shown in Figure 2. Pure ZSM-5 exhibits distinct characteristic peaks at 7.95°, 8.80°, 23.10°, 23.37°, and 24.00°, corresponding to (101), (200), (332), (051), and (303) crystal planes, respectively (Figure 2a). It can be observed that the diffraction peak of the ZSM-5 sample can be considered to be in good agreement with the diffraction peak of the standard ZSM-5 (PDF # 42-0305), indicating that the purity of the ZSM-5 sample used in the experiment is relatively high.
The quantity of carrier employed will influence the concentration of active components within the catalyst, which in turn affects the diffraction peak intensity of the active components. Upon the addition of a small quantity of ZSM-5 to the cobalt manganese solution system, the diffraction peaks of Co3O4 become more pronounced (Figure 2b). Distinct diffraction peaks can be observed at 36.76°, 59.17°, 44.92°, and 64.98°, corresponding to the (331), (511), (400) and (440) crystal planes of Co3O4, respectively. Figure 2e illustrates the diffraction angle 2θ. The diffraction peak in the range of 36°~38° corresponds to a diffraction angle of approximately 36.63° for the (331) crystal face of Co3O4. It can be observed that as the quantity of ZSM-5 incorporated into the catalyst increases (Figure 2b–d), the diffraction peak intensity of the (331) crystal plane gradually diminishes. This is attributed to a reduction in the relative content of the active component Mn-Co3O4 in the catalyst.
In order to gain insight into the internal structure and chemical bonds of the catalyst, we conducted infrared spectroscopy experiments on the sample. Figure 3a illustrates the FTIR spectra of Mn-Co3O4/meso-ZSM-5 catalysts with varying amounts of ZSM-5 incorporation. The peak at 1638 cm−1 in each group of samples is indicative of a bending vibration peak of water, thereby suggesting that each group of samples contains adsorbed water [19]. A group of peaks at 1233 cm−1 and 1079 cm−1 can be observed, which are related to the asymmetric stretching vibration of Si-O-Si bonds. These peaks can be attributed to the internal chemical bonds in SiO4 or AlO4 in ZSM-5 [20]. The external asymmetric stretching vibration near 1233 cm−1 is considered to be the characteristic vibration of the double pentagonal ring in ZSM-5 [21]. The absorption peak near 1079 cm−1 is attributed to the asymmetric tensile vibration within the Si-O-T bond rod. This change can be attributed to the slightly lower relative atomic mass of aluminum compared to silicon [22]. The absorption peaks at approximately 544 cm−1 and 447 cm−1 are characteristic absorption peaks of the ZSM-5 crystal structure [21]. The absorption peak at 447 cm−1 is attributed to the T-O bending vibration of the tetrahedra within the silicate and AlO4. Furthermore, Co3O4 exhibits a significant absorption peak at 665 cm−1, which is attributed to the O-Co-O bond bridging vibration [23]. Figure 3b,c illustrate the position and Raman spectra of the spectral samples collected from the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. As shown in Figure 3c, there are five distinct peaks at 150 cm−1 to 800 cm−1, whose peak positions are consistent with those of pure Co3O4 [24]. The 185.7 cm−1 peak corresponds to the F 2 g 1 mode of tetrahedral CoO4 sites, the 742.0 cm−1 peak corresponds to the Eg mode, the 541.1 cm−1 peak corresponds to the F 2 g 2 mode, the 600.2 cm−1 peak corresponds to the F 2 g 3 mode, and the 676.3 cm−1 peak corresponds to the A1g mode of the octahedral CoO6 site in the Co3O4 phase of the crystal [25].
The activity of catalysts is related to their specific surface area and pore structure. The Brunauer–Emmett–Teller (BET) method can be employed to determine the specific surface area and pore structure of the catalyst, thereby facilitating a more comprehensive understanding of its performance and mechanism. Figure 4a presents a comparison of adsorption and desorption in meso-ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalyst. It can be observed that both the pure meso-ZSM-5 zeolite and the Mn-Co3O4/meso-ZSM-5 catalyst exhibit typical Type IV isotherm characteristics and H3 hysteresis loops [26], indicating that both the pure meso-ZSM-5 zeolite and the Mn-Co3O4/meso-ZSM-5-6.67 catalyst have mesoporous structures. Figure 4b presents a comparison of the pore size distribution in meso-ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalyst. The specific surface area of pure ZSM-5 is 343.63 cm2/g, while that of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst is 295.77 cm2/g. The objective of this study is to determine the pore size distribution of mesoporous materials using the BJH model. From Figure 4b, it can be observed that for pure meso-ZSM-5, mesopores emerge as a result of alkaline etching [27], with the majority of mesoporous pore sizes concentrated around 3.5 nm. In the case of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst, the small-sized mesoporous pores are concentrated around 3.7 nm, while the large-sized mesopores are concentrated around 8.6 nm. This is due to the mesopores formed by the accumulation of Co3O4 particles. The objective of this study is to determine the pore size distribution of microporous materials using the Horvath Kawazoe (HK) model. From Figure 4c, it can be seen that the micropores of pure meso-ZSM-5 are concentrated at 0.57 nm, while those of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst are concentrated at 0.42 nm. The aforementioned micropores are intrinsic to the ZSM-5 material. It can be observed that the mesopore size of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst is larger than that of pure meso-ZSM-5. The presence of mesopores is beneficial for the mass transfer of gas molecules.
In order to ascertain the microstructure of the catalyst, we proceeded to conduct scanning electron microscopy (SEM) testing on the samples. Figure 5a,c present the SEM images of the ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalysts, respectively. From the figures, it can be observed that the microstructure of ZSM-5 exhibits a tetragonal shape. Upon loading ZSM-5 with Mn-Co3O4, the particle morphology tends to become spherical, and the average particle size decreases. From Figure 5b, d, we can see that the average particle size of ZSM-5 particles is 2.19 μ m , while the average particle size of Mn-Co3O4/meso-ZSM-5-6.67 particles is 1.55 μ m . The reduction in particle size can be attributed to the acidity of the cobalt nitrate aqueous solution, which has an etching effect on ZSM-5 under high-temperature hydrothermal conditions.
To gain further insight into the distribution of active components within ZSM-5 zeolite and Mn-Co3O4/meso-ZSM-5-6.67 catalyst, transmission electron microscopy (TEM) analysis was conducted on ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalyst. Figure 5e shows the transmission electron microscopy (TEM) image of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. It can be observed that the Mn-Co3O4/meso-ZSM-5-6.67 catalyst comprises elliptical grains with rough grain boundaries, indicative of the loading of Mn-Co3O4 particles on the surface of ZSM-5. It was observed that a cavity structure appeared in the TEM image at this time, which can be attributed to the etching effect of the alkaline solution during the alkaline treatment. This process resulted in the removal of the Al element from the ZSM-5 zeolite, thereby forming a cavity. Figure 5f presents an enlarged TEM image of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. In Figure 5g, portions of the crystal planes of Co3O4 are indicated, including the (311) crystal plane with a crystal plane spacing of d = 0.244 nm and the (400) crystal plane with a crystal plane spacing of d = 0.202 nm. Figure 5h represents the HAADF-STEM mapping image of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. Additionally, an elemental surface scanning analysis was conducted on the particles in the Mn-Co3O4/meso-ZSM-5-6.67 catalyst (Figure 5i–n), which clearly demonstrated the presence of Co and Mn elements. This further substantiates the successful loading of Mn-Co3O4 on the surface of ZSM-5 zeolite.
The results of the methane catalytic performance testing of the Mn-Co3O4/meso-ZSM-5 catalyst under different concentrations of meso-ZSM-5 zeolite addition are presented in Figure 6. The optimal catalytic activity of the catalyst is observed at a concentration of 6.67 g/L of ZSM-5 zeolite. At this point in time, the T50 value of the catalyst is 305 °C, while the T90 value of the catalyst is 365 °C. It is possible to achieve a conversion rate of 100% under 400 °C (GHSV = 36,000 mL·g−1·h−1). At a concentration of 20.00 g/L of ZSM-5 zeolite, the T50 value and the T90 value of the catalyst are 362 °C and 448 °C, respectively. At a concentration of 53.33 g/L, the temperature at which 50% conversion occurs is 428 °C, while the temperature at which 90% conversion occurs is 583 °C. It can be observed that the catalytic performance of pure ZSM-5 is the lowest due to the absence of the loaded active component Mn-Co3O4. As the proportion of the active component Mn-Co3O4 increases, the catalytic performance of the catalyst also improves.
In methane catalytic oxidation reactions, the higher the catalytic activity of the catalyst, the lower the corresponding activation energy. Figure 6b–d illustrate the Arrhenius fitting curves of the three catalysts. It can be observed that the Mn-Co3O4/meso-ZSM-5-6.67 catalyst exhibits the highest catalytic activity and the lowest activation energy, with a value of only 8.71 KJ/mol. As the quantity of ZSM-5 zeolite employed increases, the proportion of active components diminishes, the catalytic activity of the catalyst also declines, and the corresponding activation energy also rises. The activation energies of Mn-Co3O4/meso-ZSM-5-6.67 and Mn-Co3O4/meso-ZSM-5-6.67 were found to be 14.18 KJ/mol and 35.54 KJ/mol, respectively.
To further investigate the effect of ZSM-5 zeolite with different Si/Al ratios on the methane conversion rate, two ZSM-5-supported Mn-Co3O4 catalysts with Si/Al ratios of 70 and 170 were prepared, respectively, using the same preparation method as in this experiment. From Figure 6e,f, it can be seen that as the Si/Al ratio in ZSM-5 zeolite increases, the methane conversion rate gradually decreases. When the Si/Al ratio of ZSM-5 zeolite changed from the original 30 to 70 and 170, the T50 value of the catalyst also changed from the original 305 °C to 357 °C and 380 °C, and the T90 value changed from the original 363 °C to 434 °C and 493 °C. Correspondingly, the activation energy of the catalytic reaction gradually increased to 12.31 KJ/mol and 23.57 KJ/mol, both higher than the original value of 8.71 KJ/mol. As the Si/Al ratio in ZSM-5 zeolite continues to increase, the acidity of the zeolite support decreases. The highly acidic zeolite support can effectively suppress the removal of aluminum elements during catalytic reactions, making the catalyst highly stable and active. Conversely, zeolite supports with lower acidity lead to a decrease in catalyst performance due to their inability to effectively inhibit aluminum removal.
Table 1 demonstrates that the prepared catalyst exhibits superior catalytic performance to most existing noble metal-based catalysts and metal oxide catalysts [11,28,29,30,31,32,33,34,35]. It is capable of catalyzing methane at lower temperatures. It can be seen in Table 1 that the T90 values of most precious metal-based catalysts exceed 390 °C, whereas the T90 values of the same metal oxide catalyst are predominantly above 400 °C. In order to more effectively illustrate the remarkable catalytic efficacy of the catalyst synthesized in this study, the specific catalytic reaction conditions are presented in Table 1.
The catalytic oxidation mechanism of methane on Mn-Co3O4/meso-ZSM-5 catalyst was analyzed using in situ infrared spectroscopy. Figure 7a depicts the in situ infrared spectrum of methane with a volume fraction of 1% adsorbed on the Mn-Co3O4/meso-ZSM-5-6.67 catalyst at 250 °C. From the graph, it can be observed that there are distinct absorption bands at 3014 cm−1, 2349 cm−1, and 1304 cm−1. The vibration peaks observed at 3014 cm−1 and 1304 cm−1 are attributed to the antisymmetric stretching vibration of the C-H bond of gaseous methane molecules. This is a fundamental requirement for the occurrence of catalytic combustion reactions [36]. The C=O band is located at 1600–1800 cm−1 and belongs to the formic acid and carbonate groups [37], which are the main intermediate substances in the catalytic combustion process of methane. The asymmetric stretching vibration peak at 2349 cm−1 is attributed to gaseous carbon dioxide [38], indicating that carbon dioxide is formed in the absence of oxygen molecules. Furthermore, the intensity of the vibration peak is observed to gradually increase with the increase of adsorption time.
The in situ infrared spectra of methane oxidation on Mn-Co3O4/meso-ZSM-5-6.67 catalysts at different temperatures are shown in Figure 7b. The stretching vibration peaks of the C-H bond at 3014 cm−1 and 1304 cm−1 are attributed to gaseous CH4 [39]. As the temperature increases, the methane signal gradually decreases, indicating that the catalytic oxidation rate of methane is continuously increasing. The absorption band centered at 1504 cm−1 represents the asymmetric stretching of C H 3 [37], while the vibration peak of -CH2- appears at 2849 cm−1. Concomitantly, the intensity of the C H 3 and -CH2- bands increases with rising temperature [40]. The absorption band at 2384–2292 cm−1 is attributed to the C=O vibration mode of methane combustion product CO2 [40]. This signal is observed to increase with temperature. Furthermore, the absorption peak observed at 3743 cm−1 is attributed to the -OH group generated and adsorbed by the methane oxidation reaction. [38], yet it is absent in the absence of gaseous O2 (Figure 7a), indicating that hydroxyl groups are not readily formed in the absence of gaseous O2.
The in situ infrared results allow us to propose a reaction mechanism for methane oxidation on the Mn−Co3O4/meso−ZSM-5-6.67 catalyst. Scheme 2 depicts the reaction mechanism on it. As illustrated in Scheme 2, the mechanism of the reaction between reactants and catalyst lattice oxygen ions can be observed. The initial stage of the reaction involves the reactant methane (CH4) combining with the lattice oxygen on the catalyst surface. This process results in the formation of an intermediate product, which is subsequently oxidized to yield the final products H2O and CO2. Concurrently, the catalyst generates oxygen vacancies that are subsequently reduced. The second step is to replenish the oxygen vacancies on the surface of the catalyst and re-oxidize them in order to achieve regeneration.
In order to further investigate the changes in elemental valence states of pure ZSM-5 and Mn−Co3O4/meso−ZSM-5 catalysts and to verify the active site role of Co3+ in the catalytic process, XPS testing was conducted on the samples. The characteristic peaks of Co 2p and Mn 2p are marked in Figure 8 [41,42,43,44]. It can be observed that the ZSM-5 zeolite does not contain any Co or Mn elements prior to loading, yet these elements become apparent in Mn−Co3O4/meso−ZSM-5-6.67 catalyst, thereby indicating that the Co and Mn elements have been successfully loaded onto the ZSM-5 zeolite.
Furthermore, we conducted a detailed analysis and comparison of the surface states of the catalyst before and after the reaction. Figure 9a presents a comparative analysis of the Al 2p spectra of pure ZSM-5 zeolite and Mn-Co3O4/meso-ZSM-5-6.67 catalyst. A clear Al peak can be observed in pure ZSM-5 zeolite, with a peak position of 74.6 eV. However, the Al peak in Mn-Co3O4/meso-ZSM-5-6.67 catalyst did not show a clear peak position, indicating a significant decrease in the surface Al element content of the catalyst. This is due to the fact that cobalt trioxide particles are loaded on the surface of ZSM-5 zeolite, which makes it challenging to detect the Al element. Figure 9b presents a comparison of the Si 2p spectra of the sample before and after loading the catalyst. The peak position of the Si 2p spectrum in pure ZSM-5 zeolite is approximately 103.2 eV, which is comparable to the peak position of 102.8 eV in Mn-Co3O4/meso-ZSM-5-6.67 catalyst. This indicates that the valence state of the Si element has not been altered. The observed decrease in the peak position of the Si element is attributed to the shielding effect of the cobalt trioxide loaded on the surface of the zeolite, which effectively reduces the sensitivity of the Si element to the electron beam.
Figure 9c presents a detailed analysis of the Co 2p spectrum of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. Two distinct peaks are observed in the spectrum, with binding energies of 796.1 eV and 780.4 eV [45]. These peaks are attributed to the spin orbitals of Co 2p1/2 and Co 2p3/2, respectively. The peak positions of Co2+ vibrational satellites on the Co 2p1/2 and Co 2p3/2 spin orbits are 803.8 eV and 787.1 eV, respectively. The binding energy of Co3+ is located at 795.9 eV (Co 2p1/2) and 779.9 eV (Co 2p3/2), while the binding energy of Co2+ is located at 797.9 eV (Co 2p1/2) and 781.6 eV (Co 2p3/2). The ratio of Co3+/CO2+ is 0.368. Figure 9d presents the analysis of the Co 2p orbitals of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst following the stability assessment testing. The two distinct peaks observed in the spectrum, situated at approximately 796.5 eV and 780.7 eV [46], respectively, can be attributed to the spin orbitals of Co 2p1/2 and Co 2p3/2. The peak positions of the Co2+ vibration satellites on these two spin orbits are 804.0 eV and 787.3 eV, respectively. The binding energies of Co3+ are 795.4 eV (Co 2p1/2) and 780.2 eV (Co 2p3/2), while those of Co2+ are 796.9 eV (Co 2p1/2) and 781.4 eV (Co 2p3/2), respectively. The ratio of Co3+/CO2+ is 0.948. The binding energy peaks at 796.9 eV~797.9 eV and 781.4 eV~781.6 eV are indicative of tetrahedral coordination of Co2+, while the binding energy peaks at 795.4 eV~795.9 eV and 779.4 eV~780.2 eV are indicative of octahedral coordination of Co3+. The ratio of Co3+/CO2+ increased from 0.368 to 0.948, accompanied by a proportional increase in the relative amount of Co3+. This is due to the fact that under an oxidizing atmosphere, Co2+ is gradually oxidized to Co3+, and the catalytic reaction is accompanied by a transformation of the valence state of this Co element.
Figure 9e presents a detailed analysis of the Mn 2p spectrum of the Mn-Co3O4/meso-ZSM-5-6.67 catalyst. The two characteristic peaks observed at approximately 653.3 eV and 642.5 eV [47] in the spectrum are attributed to the spin orbitals of Mn 2p1/2 and Mn 2p3/2, respectively. The binding energy of Mn3+ is located at 654.6 eV (Mn 2p1/2) and 643.9 eV (Mn 2p3/2), while that of Mn2+ is located at 652.7 eV (Mn 2p1/2) and 641.6 eV (Mn 2p3/2). The ratio of Mn3+/Mn2+ is 0.891. Figure 9f presents the analysis of the Mn 2p orbitals of the Mn-Co3O4/mesoZSM-5-6.67 catalyst following the stability assessment testing. The two characteristic peaks observed at approximately 651.2 eV and 641.2 eV in the spectrum are attributed to the spin orbitals of Mn 2p1/2 and Mn 2p3/2, respectively. The binding energy of Mn3+ is located at 653.6 eV (Mn 2p1/2) and 642.5 eV (Mn 2p3/2), while the binding energies of Mn2+ are located at 651.9 eV (Mn 2p1/2) and 640.9 eV (Mn 2p3/2). The ratio of Mn3+/Mn2+ was found to be 1.041. The ratio of Mn3+/Mn2+ before and after the reaction increased from 0.891 to 1.041, indicating that Mn2+ was converted to Mn3+ during the reaction process. This is beneficial for the conversion of Co2+ to Co3+, as it can expose more Co3+ sites on the catalyst surface, thereby further verifying the active site role of Co3+ in the catalytic process.
Figure 9g illustrates that the peak position of O 1s in pure ZSM-5 is approximately 532.5 eV [48]. Figure 9h,i demonstrate that the peak position of O 1s in the Mn-Co3O4/meso-ZSM-5-6.67 catalyst and catalyst following the stability assessment testing are approximately 532.1 eV and 531.2 eV [49], respectively. The three peaks exhibit a similar profile, indicating that the valence state of oxygen has not undergone any significant changes. Two distinct oxygen species were identified in the O 1s spectra following Mn-Co3O4/meso-ZSM-5-6.67 catalyst and catalyst following the stability assessment testing: surface adsorbed oxygen (Oads) and surface lattice oxygen (Olatt). The ratio of Oads/Olatt in Mn-Co3O4/meso-ZSM-5-6.67 catalyst is 2.623, while that of the sample after stability testing is 1.083. A greater ratio of Oads/Olatt indicates a greater quantity of oxygen adsorbed on the surface of the sample, which in turn leads to the formation of more oxygen vacancies on the surface. This results in a higher concentration of active oxygen species, which in turn enhances the catalyst activity.

3. Materials and Methods

3.1. Experimental Supplies

The drug information utilized in this experiment is as follows: Sodium hydroxide was procured from Tianjin Oubokai Chemical Co., Ltd. (Tianjin, China) with analytical grade purity. Tetrapropyl ammonium hydroxide was procured from Shanghai McLean Biochemical Technology Co., Ltd. (Shanghai, China) with a concentration of 25%. ZSM-5 zeolite was procured from the Catalyst Plant of Nankai University, with the ratio of Si/Al is 25–30. Cobalt nitrate hexahydrate was procured from Aladdin Reagent (Shanghai) Co., Ltd. (Shanghai, China) with analytical purity. Manganese acetate tetrahydrate was procured from Aladdin Reagent (Shanghai) Co., Ltd. with analytical purity. Ammonium bicarbonate was procured from Aladdin Reagent (Shanghai) Co., Ltd. with analytical purity.

3.2. ZSM-5 Zeolite Pretreatment

The pretreatment procedure is as follows: Weigh 0.80 g of sodium hydroxide and 4.07 g of tetrapropyl ammonium hydroxide using an electronic balance. Dissolve them in 100 mL of deionized water and stir the mixed alkaline solution at 65 °C for 30 min. The ZSM-5 zeolite was ground in a mortar and placed in a beaker. The aforementioned mixed alkaline solution was added, and the mixture was stirred evenly. It was then left to stand, centrifuged, and separated to obtain the solid product. The sample was dried at 110 °C for 12 h and finally calcinated at 550 °C in an air atmosphere for 5 h to remove tetrapropyl ammonium hydroxide from the sample. This process yielded modified ZSM-5 zeolite with a micro-mesoporous structure.

3.3. Catalysts Preparation

Firstly, weigh 4.35 g of cobalt nitrate and 0.73 g of manganese acetate, then dissolve them in 60 mL of deionized water. Once dissolved, the solution should be stirred well and divided into three equal parts. Subsequently, weigh 0.2 g, 0.6 g, and 1.6 g of ZSM-5 zeolite treated with an alkaline solution, add them to the aforementioned cobalt manganese mixed solution, and stir thoroughly to obtain the precursor mixture. Subsequently, 1.42 g of ammonium bicarbonate should be weighed and dissolved in 30 mL of deionized water. The solution of ammonium bicarbonate should be stirred thoroughly and divided into three equal parts. Subsequently, the precursor mixture is to be added to the ammonium bicarbonate solution in a continuous and slow manner, with stirring throughout the process. Once all the components have been added, the solution should be left to stand at room temperature and stirred continuously for a period of two hours in order to ensure thorough mixing. Subsequently, the suspension is placed in a high-pressure reactor and subjected to a 170 °C drying oven for 12 h, during which time it undergoes a crystallization process. Subsequently, the product is subjected to centrifugation, washing, and filtration with anhydrous ethanol and deionized water. Subsequently, the solid product is subjected to a vacuum drying process at a temperature of 70 °C for a period of six hours. Subsequently, the dried sample was subjected to calcination in a muffle furnace at 350 °C for a period of four hours, resulting in the formation of a black product. The preparation process is illustrated in Scheme 1. The catalyst should be named according to the mass concentration of ZSM-5 zeolite. For example, the addition of 0.2 g of ZSM-5 zeolite to 30 mL of deionized water results in the formation of a catalyst designated as Mn-Co3O4/meso-ZSM-5-6.67. Similarly, the catalyst with 0.6 g of ZSM-5 zeolite added is designated as Mn-Co3O4/meso-ZSM-5-20.00, while the catalyst with 1.6 g of ZSM-5 zeolite added is designated as Mn-Co3O4/meso-ZSM-5-53.33.

3.4. Catalysts Characterization

X-ray diffraction (XRD) was conducted on a DX-2500 X-ray diffractometer (Dandong Fangyuan Instrument Co., Ltd., Dandong, China) with a working voltage of 40 kV, a working current of 40 mA, and the scanning angle range was 5° to 90°. The Fourier transform infrared spectrometer (FTIR-8400s, Shimadzu Co., Ltd., Kyoto, Japan) was employed for FTIR testing, with a scanning range of 3200 cm−1 to 400 cm−1. Electron microscopy measurements were conducted using scanning electron microscopy (SEM) from Phenom Prox (Phenom World Company, Eindhoven, The Netherlands), with an accelerated voltage of 10 kV and maintained in a vacuum state. A gap-specific surface area analyzer (SSA-4000) was employed for specific surface area testing (Beijing Builder Electronic Technology Co., Ltd., Beijing, China). X-ray photoelectron spectroscopy (XPS) was utilized for XPS testing, with excitation rays of Al Kα. The working voltage was 15 kV, the working current was 15 mA, and the power was 225 W. The calibration standard for the electron binding energy of other elements is a polluted carbon C 1s (284.8eV). Transmission electron microscopy (TEM) testing was conducted using a Tecnai G2 F20 transmission electron microscope (Hillsboro, OR, USA). Raman spectroscopy analysis was performed using the LabRAM HR-800 Raman spectrometer (RS) from Horiba Jobin Yvon, Palaiseau, France. In situ infrared testing was conducted using an FTIR-8400s Fourier transform infrared spectrometer (Shimadzu Co., Ltd.), with a scanning range of 4000 cm−1 to 1000 cm−1 during the experiment. Brunauer-Emmett-Teller test (BET) of catalyst particles was conducted using a Kubo X1000 Surface Area and Pore Size Analyzer (Beijing Builder Electronic Technology Co., Ltd., Beijing, China).

3.5. Catalysts Performance Evaluation

The Mn-Co3O4/meso-ZSM-5 composite catalysts synthesized in this experiment were evaluated for the catalytic oxidation performance of lean methane in a fixed bed reactor. The reactor is equipped with a quartz reaction tube with a fixed sand core in the middle. During the experiment, A gas mixture composed of 1 vol% CH4 and 99% equilibrium air was passed through a mass flow meter with a controlled flow rate of 60 mL/min, an air velocity (GHSV) of 36,000 mL·g−1·h−1, and a reaction temperature of 200 °C to 700 °C. The gas composition of the inlet and outlet of the reaction tube was analyzed online by a gas chromatograph equipped with a TCD detector. (Best Instrument Technology Co., Ltd., Beijing, China)
In this experiment, T50 and T90, namely the temperature at which CH4 is converted to 50% and 90%, were selected as the criteria for evaluating the catalytic performance of the catalyst. The formula for calculating the conversion rate of low-concentration CH4 is X = (Cin − Cout)/Cin × 100%.
In the formula, X is the conversion rate of low-concentration methane gas, Cin is the initial concentration of methane before catalytic combustion, and Cout is the residual concentration of methane after catalytic combustion.

4. Conclusions

This study first synthesized a series of Mn-CoCO3/meso-ZSM-5 catalyst precursors via the hydrothermal method. Subsequently, the Mn-Co3O4/meso-ZSM-5 catalysts were prepared through the calcination method. The introduction of manganese ions into the lattice of Co3O4 results in lattice distortion, which in turn causes the formation of oxygen vacancies and the continuous conversion of Co2+ to Co3+ during the catalytic reaction. This enhances the catalytic activity of the catalyst for methane combustion. Among the catalysts, the Mn-Co3O4/meso-ZSM-5-6.67 catalyst exhibits the most promising catalytic activity for methane combustion (T50 = 305 °C, T90 = 363 °C). Its catalytic activity is superior to that of similar cobalt manganese catalyst systems and even better than that of precious metal-based catalysts. The primary reason for this is attributed to the mesoporous ZSM-5 carrier with a high specific surface area, which is conducive to the adsorption and mass transfer of reaction molecules. The active component is characterized by a high density of oxygen vacancies, which facilitate the activation of reaction molecules and, as a consequence, enhance their catalytic activity. The novel approaches to the synthesis of metal oxide-based methane low-temperature oxidation catalysts proposed in this study are anticipated to offer novel solutions to methane low-temperature oxidation reactions and facilitate technological advancement in related fields.

Author Contributions

Y.Z.: Writing—Original Draft, Methodology, Investigation, Data Curation; R.W.: Writing—Original Draft, Data Curation, Formal analysis; L.Y.: Writing—Reviewing and Editing; J.G.: Writing—Reviewing and Editing; F.H.: Writing—Reviewing and Editing; T.Z.: Writing—Reviewing and Editing; F.L.: Writing—Reviewing and Editing; H.W.: Conceptualization, Funding acquisition, Supervision, Resources; J.Q.: Conceptualization, Funding acquisition, Formal analysis, Writing—Reviewing and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China under Grant (No. 51972306, 21604007).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Stolaroff, J.K.; Bhattacharyya, S. Review of methane mitigation technologies with application to rapid release of methane from the arctic. Environ. Sci. Technol. 2012, 46, 6455–6469. [Google Scholar] [CrossRef]
  2. Ercolino, G.; Stelmachowski, P. Reactivity of mixed iron-cobalt spinels in the lean methane combustion. Top. Catal. 2017, 60, 1370–1379. [Google Scholar] [CrossRef]
  3. Lu, R.; Lin, J. A interpretation of stepwise bond dissociation energies of CH4. Comput. Theor. Chem. 2014, 1037, 10–13. [Google Scholar] [CrossRef]
  4. Mudiyanselage, K.; Senanayake, S.D. Importance of the metal-oxide interface in catalysis: In situ studies of the water-gas shift reaction by ambient-pressure X-ray photoelectron spectroscopy. Angew. Chem. Int. Ed. 2013, 52, 5101–5105. [Google Scholar] [CrossRef]
  5. Kumar, R.S.; Hayes, R.E. Kinetic investigation of the promoting effect of cobalt on Pd-Pt/SnO2 catalyzed wet methane combustion. Appl. Catal. A 2023, 666, 119416. [Google Scholar] [CrossRef]
  6. Fung, V.; Janik, M. Toward understanding and controlling organic reactions on metal oxide catalysts. J. Phys. Chem. C 2023, 127, 13451–13465. [Google Scholar] [CrossRef]
  7. Zhao, N.; Chen, Y. Preparation of high performance Co3O4/Al2O3 catalysts by doping Al into ZIF-67: Effect of Al sources on Fischer-Tropsch synthesis. Appl. Surf. Sci. 2021, 570, 151127. [Google Scholar] [CrossRef]
  8. Li, D.; Xu, R. Encapsulated Co3O4/(SiAl@Al2O3) thermal storage functional catalysts for catalytic combustion of lean methane. Appl. Therm. Eng. 2020, 181, 116012. [Google Scholar] [CrossRef]
  9. Liotta, L.F.; Di Carlo, G. Co3O4/CeO2 composite oxides for methane emissions abatement: Relationship between Co3O4-CeO2 interaction and catalytic activity. Appl. Catal. B Environ. 2006, 66, 217–227. [Google Scholar] [CrossRef]
  10. Lv, Y.; Li, Y. Synthesis of Co3O4 nanotubes and their catalytic applications in CO oxidation. Catal. Commun. 2013, 42, 116–120. [Google Scholar] [CrossRef]
  11. Lv, C.; Du, D. The flower-like Co3O4 hierarchical microspheres for methane catalytic oxidation. Inorganics 2022, 10, 49. [Google Scholar] [CrossRef]
  12. Guo, S. Heteroatom doping improves the electrocatalytic performance of cobalt tetroxide. Acta Phys. Chim. Sin. 2020, 36, 2001012. [Google Scholar] [CrossRef]
  13. Zheng, Y.; Yu, Y. Combustion of lean methane over Co3O4 catalysts prepared with different cobalt precursors. RSC Adv. 2020, 10, 4490–4498. [Google Scholar] [CrossRef]
  14. Lang, Q.; Lu, P. Zeolites for the environment. Green Carbon 2024, 2, 12–32. [Google Scholar] [CrossRef]
  15. Velinova, R.; Grahovski, B. Reaction kinetics and mechanism of the catalytic oxidation of propane over Co-ZSM-5 zeolites. React. Kinet. Mech. Catal. 2022, 135, 83–103. [Google Scholar] [CrossRef]
  16. Fei, X.; Cao, S. Comparative study of Co3O4-ZSM-5 catalysts synthesized by different hydrothermal methods for the catalytic oxidation of dichloromethane. Chin. Chem. Lett. 2021, 32, 1224–1228. [Google Scholar] [CrossRef]
  17. Fei, X.; Ouyang, W. Effect of Cr doping in promoting the catalytic oxidation of dichloromethane (CH2Cl2) over Cr-Co@Z catalysts. J. Hazard. Mater. 2021, 413, 125327. [Google Scholar] [CrossRef]
  18. Li, Z.; Wu, L. Characterizations and product distribution of Co-based Fischer-Tropsch catalysts: A comparison of the incorporation manner. Fuel Cells 2018, 220, 257–263. [Google Scholar] [CrossRef]
  19. Zakaria, Z.Y.; Linnekoski, J. Catalyst screening for conversion of glycerol to light olefins. Chem. Eng. J. 2012, 207, 803–813. [Google Scholar] [CrossRef]
  20. Jiang, Y.-j.; Juan, J.C. Preparation and catalytic application of novel water tolerant solid acid catalysts of zirconium sulfate/HZSM-5. Chem. Res. Chin. Univ. 2007, 23, 349–354. [Google Scholar] [CrossRef]
  21. Shirazi, L.; Jamshidi, E. The effect of Si/Al ratio of ZSM-5 zeolite on its morphology, acidity and crystal size. Cryst. Res. Technol. 2008, 43, 1300–1306. [Google Scholar] [CrossRef]
  22. Ali, M.A.; Brisdon, B. Synthesis, characterization and catalytic activity of ZSM-5 zeolites having variable silicon-to-aluminum ratios. Appl. Catal. A Gen. 2003, 252, 149–162. [Google Scholar] [CrossRef]
  23. Estepa, L.; Daudon, M. Contribution of Fourier transform infrared spectroscopy to the identification of urinary stones and kidney crystal deposits. Biospectroscopy 1997, 3, 347–369. [Google Scholar] [CrossRef]
  24. Wang, R.M.; Liu, C.M. Porous nanotubes of Co3O4: Synthesis, characterization, and magnetic properties. Appl. Phys. Lett. 2004, 85, 2080–2082. [Google Scholar] [CrossRef]
  25. Mo, S.; Zhang, Q. Integrated cobalt oxide based nanoarray catalysts with hierarchical architectures: In situ Raman spectroscopy investigation on the carbon monoxide reaction mechanism. ChemCatChem 2018, 10, 3012–3026. [Google Scholar] [CrossRef]
  26. Raman, G.; Das, J. Highly efficient mesoporous ZSM-5 for trace olefin removal from aromatic stream. Inorg. Chim. Acta 2023, 544, 121201. [Google Scholar] [CrossRef]
  27. Yoo, W.C.; Zhang, X. Synthesis of mesoporous ZSM-5 zeolites through desilication and re-assembly processes. Microporous Mesoporous Mater. 2012, 149, 147–157. [Google Scholar] [CrossRef]
  28. Zhang, Z.; Sun, L. Anti-sintering Pd@silicalite-1 for methane combustion: Effects of the moisture and SO2. Appl. Surf. Sci. 2019, 494, 1044–1054. [Google Scholar] [CrossRef]
  29. Yang, W.; Polo-Garzon, F. Boosting the activity of Pd single atoms by tuning their local environment on ceria for methane combustion. Angew. Chem. Int. Ed. 2023, 62, e202217323. [Google Scholar] [CrossRef]
  30. Ma, Y.; Li, S. Construction of a Pd(PdO)/Co3O4@SiO2 core-shell structure for efficient low-temperature methane combustion. Nanoscale 2021, 13, 5026–5032. [Google Scholar] [CrossRef]
  31. Kumar, R.S.; Hayes, R.E. Pd and Pd-Pt catalysts supported on SnO2 and c-Al2O3: Kinetic studies of wet lean methane combustion. Chem. Eng. Sci 2023, 269, 118488. [Google Scholar] [CrossRef]
  32. Duan, H.; You, R. Pentacoordinated Al3+-stabilized active Pd structures on Al2O3-coated palladium catalysts for methane combustion. Angew. Chem. Int. Ed. 2019, 58, 12043–12048. [Google Scholar] [CrossRef]
  33. Wu, H.; Pantaleo, G. Co3O4 particles grown over nanocrystalline CeO2: Influence of precipitation agents and calcination temperature on the catalytic activity for methane oxidation. Catal. Sci. Technol. 2015, 5, 1888–1901. [Google Scholar] [CrossRef]
  34. Jiang, Y.; Li, W. A rod-like Co3O4 with high efficiency and large specific surface area for lean methane catalytic oxidation. Mol. Catal. 2022, 522, 112229. [Google Scholar] [CrossRef]
  35. Teng, F.; Chen, M. High combustion activity of CH4 and catalluminescence properties of CO oxidation over porous Co3O4 nanorods. Appl. Catal. B Environ. 2011, 110, 133–140. [Google Scholar] [CrossRef]
  36. Wu, M.; Li, W. Highly active and stable palladium catalysts supported on surface-modified ceria nanowires for lean methane combustion. ChemCatChem 2021, 13, 664–673. [Google Scholar] [CrossRef]
  37. Wang, Y.; Arandiyan, H. High performance Au-Pd supported on 3D hybrid strontium-substituted lanthanum manganite perovskite catalyst for methane combustion. ACS Catal. 2016, 6, 6935–6947. [Google Scholar] [CrossRef]
  38. Xiong, J.; Yang, J. Pd-Promoted Co2NiO4 with lattice Co-O-Ni and interfacial Pd-O activation for highly efficient methane oxidation. Appl. Catal. B Environ. 2021, 292, 120201. [Google Scholar] [CrossRef]
  39. Jodlowski, P.J.; Jedrzejczyk, R.J. In situ spectroscopic studies of methane catalytic combustion over Co, Ce, and Pd mixed oxides deposited on a steel surface. J. Catal. 2017, 350, 1–12. [Google Scholar] [CrossRef]
  40. Chen, S.; Li, S. Elucidation of active sites for CH4 catalytic oxidation over Pd/CeO2 via tailoring metal-support interactions. ACS Catal. 2021, 11, 5666–5677. [Google Scholar] [CrossRef]
  41. Yao, Q.; Xiao, F. Regeneration of spent lithium manganate into cation-doped and oxygen-deficient MnO2 cathodes toward ultralong lifespan and wide-temperature-tolerant aqueous Zn-ion batteries. Battery Energy 2023, 2, 20220065. [Google Scholar] [CrossRef]
  42. Ding, C.; Wang, Y. Mn-doping ensuring cobalt silicate hollow spheres with boosted electrochemical property for hybrid supercapacitors. Battery Energy 2023, 2, 20230042. [Google Scholar] [CrossRef]
  43. Liu, Z.; Xu, H. Strain-rich high-entropy perovskite oxide of (La0.8Sr0.2)(Mn0.2Fe0.2Cr0.2Co0.2Ni0.2)O3 for durable and effective catalysis of oxygen redox reactions in lithium-oxygen battery. Battery Energy 2024, 3, 20230053. [Google Scholar] [CrossRef]
  44. Mitra, S.; Sudakar, C. High rate capability and cyclic stability of Ni-rich layered oxide LiNi0.83Co0.12Mn0.05−xAlxO2 cathodes: Nanofiber versus nanoparticle morphology. Battery Energy 2024, 3, 20230066. [Google Scholar] [CrossRef]
  45. You, N.; Cao, S. Constructing P-CoMoO4@ NiCoP heterostructure nanoarrays on Ni foam as efficient bifunctional electrocatalysts for overall water splitting. Nano Mater. Sci. 2023, 5, 278–286. [Google Scholar] [CrossRef]
  46. Zeng, X.; Zhang, H. Coupling of ultrasmall and small CoxP nanoparticles confined in porous SiO2 matrix for a robust oxygen evolution reaction. Nano Mater. Sci. 2022, 4, 393–399. [Google Scholar] [CrossRef]
  47. An, C.; Wang, S. Construction and ultrasonic inspection of the high-capacity Li-ion battery based on the MnO2 decorated by Au nanoparticles anode. Microstructures 2024, 4, 2024003. [Google Scholar] [CrossRef]
  48. Yan, Y.; Ma, Z. Surface microstructure-controlled ZrO2 for highly sensitive room-temperature NO2 sensors. Nano Mater. Sci. 2021, 3, 268–275. [Google Scholar] [CrossRef]
  49. Gao, Y.; Song, T. Electronic interaction and oxgen vacancy engineering of g-C3N4/α-Bi2O3 Z-scheme heterojunction for enhanced photocatalytic aerobic oxidative homo-/hetero-coupling of amines to imines in aqueous phase. Green Carbon 2023, 1, 105–117. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram for preparation of catalysts.
Scheme 1. Schematic diagram for preparation of catalysts.
Catalysts 14 00397 sch001
Figure 1. XRD pattern: (a) products after hydrothermal reaction. (b) Products after calcination. The black triangle represents the characteristic peak of Co3O4 and the black circle represents the characteristic peak of CoCO3.
Figure 1. XRD pattern: (a) products after hydrothermal reaction. (b) Products after calcination. The black triangle represents the characteristic peak of Co3O4 and the black circle represents the characteristic peak of CoCO3.
Catalysts 14 00397 g001
Figure 2. XRD patterns: (a) meso-ZSM-5. (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/mesoZSM-5-20.00. (d) Mn-Co3O4/meso-ZSM-5-53.33. (e) Enlarged view of diffraction angle at 36–38° in XRD pattern. The inverted triangle represents the characteristic peak of Co3O4 and the black circle represents the characteristic peak of ZSM-5.
Figure 2. XRD patterns: (a) meso-ZSM-5. (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/mesoZSM-5-20.00. (d) Mn-Co3O4/meso-ZSM-5-53.33. (e) Enlarged view of diffraction angle at 36–38° in XRD pattern. The inverted triangle represents the characteristic peak of Co3O4 and the black circle represents the characteristic peak of ZSM-5.
Catalysts 14 00397 g002
Figure 3. (a) FTIR spectra: (I) Mn-Co3O4/meso-ZSM-5-6.67. (II) Mn-Co3O4/meso-ZSM-5-20.00. (III) Mn-Co3O4/meso-ZSM-5-53.33. (b) Sample selection region of Raman spectrum. (c) Raman spectrum of Mn-Co3O4/meso-ZSM-5-6.67.
Figure 3. (a) FTIR spectra: (I) Mn-Co3O4/meso-ZSM-5-6.67. (II) Mn-Co3O4/meso-ZSM-5-20.00. (III) Mn-Co3O4/meso-ZSM-5-53.33. (b) Sample selection region of Raman spectrum. (c) Raman spectrum of Mn-Co3O4/meso-ZSM-5-6.67.
Catalysts 14 00397 g003
Figure 4. (a) N2 adsorption-desorption isotherms, (b) mesoporous distribution, and (c) microporous distribution of meso-ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalysts: (I) ZSM-5. (II) Mn-Co3O4/meso-ZSM-5-6.67. The dashed line perpendicular to the horizontal axis represents the numerical value of the pore size in the samples.
Figure 4. (a) N2 adsorption-desorption isotherms, (b) mesoporous distribution, and (c) microporous distribution of meso-ZSM-5 and Mn-Co3O4/meso-ZSM-5-6.67 catalysts: (I) ZSM-5. (II) Mn-Co3O4/meso-ZSM-5-6.67. The dashed line perpendicular to the horizontal axis represents the numerical value of the pore size in the samples.
Catalysts 14 00397 g004
Figure 5. SEM images: (a) meso-ZSM-5. (c) Mn-Co3O4/meso-ZSM-5-6.67. Aperture distribution histogram: (b) meso-ZSM-5. (d) Mn-Co3O4/meso-ZSM-5-6.67. TEM and HRTEM images: (eg) Mn-Co3O4/meso-ZSM-5-6.67. The right part of the dashed line represents the lattice stripes on the (311) crystal plane of Co3O4. (hn) HAADF-STEM mapping images of Mn-Co3O4/meso-ZSM-5-6.67.
Figure 5. SEM images: (a) meso-ZSM-5. (c) Mn-Co3O4/meso-ZSM-5-6.67. Aperture distribution histogram: (b) meso-ZSM-5. (d) Mn-Co3O4/meso-ZSM-5-6.67. TEM and HRTEM images: (eg) Mn-Co3O4/meso-ZSM-5-6.67. The right part of the dashed line represents the lattice stripes on the (311) crystal plane of Co3O4. (hn) HAADF-STEM mapping images of Mn-Co3O4/meso-ZSM-5-6.67.
Catalysts 14 00397 g005
Figure 6. (a) Catalytic conversion of different samples: (I) Mn-Co3O4/meso-ZSM-5-6.67. (II) Mn-Co3O4/meso-ZSM-5-20.00. (III) Mn-Co3O4/meso-ZSM-5-53.33. (IV) meso-ZSM-5. Arrhenius curves: (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/meso-ZSM-5-20.00. (d) Mn-Co3O4/meso-ZSM-5-53.33. (e) Catalytic conversion of Mn-Co3O4/meso-ZSM-5-6.67 (Si/Al ratio is 70). (f) Catalytic conversion of Mn-Co3O4/meso-ZSM-5-6.67 (Si/Al ratio is 170).
Figure 6. (a) Catalytic conversion of different samples: (I) Mn-Co3O4/meso-ZSM-5-6.67. (II) Mn-Co3O4/meso-ZSM-5-20.00. (III) Mn-Co3O4/meso-ZSM-5-53.33. (IV) meso-ZSM-5. Arrhenius curves: (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/meso-ZSM-5-20.00. (d) Mn-Co3O4/meso-ZSM-5-53.33. (e) Catalytic conversion of Mn-Co3O4/meso-ZSM-5-6.67 (Si/Al ratio is 70). (f) Catalytic conversion of Mn-Co3O4/meso-ZSM-5-6.67 (Si/Al ratio is 170).
Catalysts 14 00397 g006
Figure 7. In situ infrared spectra: (a) Mn-Co3O4/meso-ZSM-5 catalysts were subjected to different times at the same temperature. (b) Mn-Co3O4/meso-ZSM-5 catalysts were subjected to different temperatures for different periods of time.
Figure 7. In situ infrared spectra: (a) Mn-Co3O4/meso-ZSM-5 catalysts were subjected to different times at the same temperature. (b) Mn-Co3O4/meso-ZSM-5 catalysts were subjected to different temperatures for different periods of time.
Catalysts 14 00397 g007
Scheme 2. The reaction mechanism of methane catalytic oxidation proposed on Mn-Co3O4/meso-ZSM-5 catalyst.
Scheme 2. The reaction mechanism of methane catalytic oxidation proposed on Mn-Co3O4/meso-ZSM-5 catalyst.
Catalysts 14 00397 sch002
Figure 8. XPS full scan spectra: (a) meso-ZSM-5. (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction.
Figure 8. XPS full scan spectra: (a) meso-ZSM-5. (b) Mn-Co3O4/meso-ZSM-5-6.67. (c) Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction.
Catalysts 14 00397 g008
Figure 9. XPS diagram: (a) Al 2p level I: meso-ZSM-5. II: Mn-Co3O4/meso-ZSM-5-6.67. (b) Si 2p level I: meso-ZSM-5. II: Mn-Co3O4/meso-ZSM-5-6.67. (c) Co 2p level of Mn-Co3O4/meso-ZSM-5-6.67. (d) Co 2p level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction. (e) Mn 2p level of Mn-Co3O4/meso-ZSM-5-6.67. (f) Mn 2p level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction. (g) O 1s level of meso-ZSM-5. (h) O 1s level of level of Mn-Co3O4/meso-ZSM-5-6.67. (i) O 1s level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction.
Figure 9. XPS diagram: (a) Al 2p level I: meso-ZSM-5. II: Mn-Co3O4/meso-ZSM-5-6.67. (b) Si 2p level I: meso-ZSM-5. II: Mn-Co3O4/meso-ZSM-5-6.67. (c) Co 2p level of Mn-Co3O4/meso-ZSM-5-6.67. (d) Co 2p level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction. (e) Mn 2p level of Mn-Co3O4/meso-ZSM-5-6.67. (f) Mn 2p level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction. (g) O 1s level of meso-ZSM-5. (h) O 1s level of level of Mn-Co3O4/meso-ZSM-5-6.67. (i) O 1s level of Mn-Co3O4/meso-ZSM-5-6.67 after catalytic reaction.
Catalysts 14 00397 g009
Table 1. Overview of activity of catalysts in the catalytic combustion of methane.
Table 1. Overview of activity of catalysts in the catalytic combustion of methane.
CatalystsReaction ConditionsT50/°CT90/°CReference
Mn-Co3O4/meso-ZSM-5-6.671%CH4, 99%air, 36,000 mL·g−1·h−1305363this work
Pd@S-11%CH4, 20%O2, 24,000 mL·g−1·h−1308360[28]
CM-Mg0.31%CH4, 20%O2,
balance N2, 36,000 mL·g−1·h−1
318445[29]
Pd (PdO)/Co3O4@SiO21%CH4, 21%O2, 30,000 mL·g−1·h−1357445[30]
Pd-Pt/SnO21%CH4, 19.5%O2, 21,000 mL·g−1·h−1413450[31]
Al2O3/C-Pd/SiO21%CH4, 10% O2/Ar, 36,000 mL·g−1·h−1325390[32]
Co3O4/CeO20.3%CH4, 0.6%O2, 36,000 mL·g−1·h−1343450[33]
Co3O4 (Flower-like)1%CH4, 99%air, 36,000 mL·g−1·h−1380430[11]
Co3O4 (Rod-like)1%CH4, 30,000 mL·g−1·h−1320385[34]
Co3O4-NW1%CH4, 20%O2, 30,000 mL·g−1·h−1308366[35]
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, Y.; Wei, R.; Yang, L.; Ge, J.; Hu, F.; Zhang, T.; Lu, F.; Wang, H.; Qi, J. In Situ Growth of Mn-Co3O4 on Mesoporous ZSM-5 Zeolite for Boosting Lean Methane Catalytic Oxidation. Catalysts 2024, 14, 397. https://doi.org/10.3390/catal14070397

AMA Style

Zhang Y, Wei R, Yang L, Ge J, Hu F, Zhang T, Lu F, Wang H, Qi J. In Situ Growth of Mn-Co3O4 on Mesoporous ZSM-5 Zeolite for Boosting Lean Methane Catalytic Oxidation. Catalysts. 2024; 14(7):397. https://doi.org/10.3390/catal14070397

Chicago/Turabian Style

Zhang, Yuxuan, Ruibo Wei, Lin Yang, Jinming Ge, Feiyang Hu, Tingting Zhang, Fangyin Lu, Haiwang Wang, and Jian Qi. 2024. "In Situ Growth of Mn-Co3O4 on Mesoporous ZSM-5 Zeolite for Boosting Lean Methane Catalytic Oxidation" Catalysts 14, no. 7: 397. https://doi.org/10.3390/catal14070397

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop