Next Article in Journal
Protein Adsorption on a Multimodal Cation Exchanger: Effect of pH, Salt Type and Concentration, and Elution Conditions
Previous Article in Journal
Hydrophobic Coating of Paperboard Using Oak Wood-Derived Lignin Nanoparticles and Chitosan Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave-Enhanced Catalytic Performance of Benzene Oxidation on MOF-Derived Mn/Ce-Co Oxides

1
School of Chemical and Environmental Engineering, Wuhan Polytechnic University, Wuhan 430023, China
2
Pilot Plant of Eco-Environment Chemical Industry and Carbon-Neutral Transformative Technologies, Wuhan 430023, China
3
Hubei Provincial Engineering Research Center of Soil and Groundwater Pollution Prevention and Control, Wuhan 430023, China
4
School of Environmental Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(16), 3388; https://doi.org/10.3390/molecules30163388
Submission received: 18 July 2025 / Revised: 10 August 2025 / Accepted: 13 August 2025 / Published: 15 August 2025
(This article belongs to the Section Organometallic Chemistry)

Abstract

Microwave-assisted processing has shown tremendous promise in accelerating chemical reactions and reducing energy consumption through targeted dielectric heating. This study develops MOF-derived Mn-Co and Ce-Co oxide catalysts for energy-efficient benzene oxidation via microwave catalysis. The MnCo spinel oxides (particularly MnCo11-400) exhibit superior microwave absorption and catalytic activity due to enhanced oxygen mobility and tailored dielectric properties. Microwave irradiation enables rapid benzene mineralization over the MnCo11-400 catalyst, achieving 78% conversion at 30 W and complete conversion at 50 W, demonstrating exceptional energy efficiency at low power inputs. Microwaves significantly lower the reaction temperature compared to conventional thermal catalysis (ΔT = 100–250 °C). Stability tests confirm robustness over repeated power cycling (80% conversion retained after 3 × 1 h on/off cycles). Furthermore, an adsorption–microwave oxidation synergistic strategy is demonstrated: pre-adsorbed low-concentration benzene (1.15 mmol) at ambient temperature undergoes complete mineralization within 20 min under 30 W microwave irradiation. The intermittent microwave operation achieves equivalent benzene removal to continuous thermal processing while significantly reducing energy demand. This work establishes MOF-derived spinel oxides as high-performance microwave catalysts for low-temperature VOC abatement.

1. Introduction

In recent years, global air pollution has become increasingly severe, and the emission of volatile organic compounds (VOCs) has seriously affected the ambient air quality and human health. Studies have shown that [1,2] VOCs are the main precursors of secondary organic aerosols and fine particulate matter, which are important contributors to atmospheric photochemical smog and acid deposition. Under ultraviolet irradiation, VOCs will react with NOx to generate secondary pollutants like ozone, resulting in photochemical pollution [3]. Hexane, heptane, and octane can affect the human central nervous system [4]. Polycyclic aromatic hydrocarbons and many chlorine-containing organic compounds such as benzene, toluene, xylene, styrene, dichlorotoluene, etc., are carcinogenic, teratogenic, and mutagenic [5]. With the increasing environmental issues, the end treatment of VOCs is imperative.
Adsorption [6], condensation [7], catalytic oxidation [8], photocatalytic degradation [9], and biological methods [10] have been applied to control VOC emissions. Among the techniques, the catalytic oxidation method is considered to be one of the most suitable and broad application prospects for treating VOCs with a large concentration variation range due to the low light-off temperature, high conversion and mineralization rate, less secondary pollution, and wide application range.
As one of the energy resources with a high thermal effect and strong dipole polarization ability, microwaves have been widely used in the treatment of solid, liquid, and gas waste in recent years. Microwave energy has been considered an excellent alternative heating source to promote chemical reactions because of its ability to provide fast, selective, and highly efficient volumetric heating to materials [11,12,13,14]. The thermal effect is the main factor of reaction promotion via microwave irradiation [15,16,17,18]. The temperature of microscale “hot spots” formed from microwaves can reach as high as 1200 °C, and the temperature difference can be higher than 200 °C [16]. J. Dobosz et al. [19] used α-Fe2O3 as the active component of the catalyst, and propane removal efficiency reached 99.9% through the microwave-catalytic method. A microwave single-mode cavity was introduced into the catalytic reactions by Nigar et al. [20], resulting in 99% degradation efficiency of n-hexane. Our previous work [21,22,23] has proven that the microwave-assisted catalytic removal technique is more effective for small-scale VOC sources with low concentrations from the standpoint of energy efficiency. Thus, it will be an innovative reform to apply the microwave technique in VOCs removal, offering significant improvements over conventional heating, including shorter reaction times, greater energy efficiency, and higher yields of products.
The development of microwave catalysts is the key to realizing the microwave-assisted catalytic process. Transition metal oxides have high activity in catalyzing VOC reactions due to their abundant oxygen vacancies, electron transfer, and valence state changes, and they are considered to be potential alternative materials to precious metal catalysts [24]. In addition, previous studies [25,26] have shown that transition metal oxide materials have high dielectric loss, and their microwave heating performance is also quite excellent. On the other hand, metal organic frameworks (MOFs) are composed of metal nodes and organic connectors and have attracted widespread attention due to their high specific surface area, adjustable porosity, diverse structures, and adjustable functional sites, which make MOFs suitable candidates for catalytic materials or precursors [27,28,29,30]. MOF-derived porous metal oxides have highly porous structures and high surface areas, which are conducive to the adsorption of VOCs and reactivity of oxygen species [31,32]. There are few reports on the effective regulation of microwave heating characteristics and catalytic performance by adjusting the metal elements in MOFs.
In this work, to improve the efficiency of the complete catalytic benzene oxidation, we investigated the effect of the composition of bimetallic MOF-derived Ce-Co and Mn-Co oxides and their catalytic properties. The temperature rise characteristics and benzene oxidation activity of the catalysts under microwave irradiation were studied in detail. The effectiveness of the adsorption–oxidation process for complete removal of gaseous benzene at low concentrations, which consists of adsorption and catalytic oxidation via rapid heating, a characteristic of microwave heating, was also investigated.

2. Results and Discussion

2.1. Structural and Morphology Characterization

The phase composition of the catalyst was characterized by X-ray diffraction (XRD) patterns in Figure 1. According to the ICDD standard card, the CeCo-300 catalysts were mainly composed of an amorphous phase that could not be identified via XRD measurement (Figure S1). The increase in temperature enhances the crystallinity of the Ce-Co oxide samples. When calcined at 400 °C, the phase composition mainly evolved into CeO2(PDF#75-0076)-Co3O4(PDF#42-1467) for CeCo12-400, CeO2-CoO(PDF#75-0419) for CeCo11-400, and CeO2-CoO for CeCo21-400 [33]. For MnCo-400 catalysts, the phase composition differs from CeCo-400, with spinel oxide MnCo2O4(PDF#84-0482) forming in MnCo12-400 and MnCo11-400, while CoMn2O4(PDF#77-0471) is generated in MnCo21-400 [34,35]. Co-Mn spinel oxide is an excellent oxidation catalyst with higher redox properties, thermal and catalytic stability, and resistance to coke formation than the single oxides [21,22].
Figure 2 and Figure S1 show the XPS spectra of the Ce-Co and Mn-Co oxides. For the CeCo-400, the Ce 3d spectra (Figure 2a) comprise ten peaks corresponding to five pairs of spin-orbit doublets. The peaks located at around 897.8, 881.9, 902.4, and 885.6 eV can be assigned to the v, u, v′ and u′ shake-down final states of Ce3+, while the other peaks centered at around 882.8, 901.1, 888.6, 907.2, 898.5, and 916.6 eV can be classified as the v, u, v″, u″, and v″′, u″′ final states of Ce4+ [36]. The curves are fitted with Co2+ and Co3+ components via Co 2p spectra (Figure 2b), characterized by the binding energy located at about 780.8, 796.1 eV and 779.3, 794.4 eV, respectively. The peak 786.0 eV corresponds to the satellite peaks of Co2+ species [36]. Peaks with binding energies at 529.0, 531.2, and 532.3 eV in the O 1s spectra (Figure 2c) can be attributed to surface lattice oxygen (Olatt), oxygen vacancy (Ovac), and surface adsorption oxygen species (Oads), respectively [37]. For the MnCo-400, the peaks in Mn 2p spectra (Figure 2d) around 652.6 eV and 640.7 eV are assigned to Mn 2p1/2 and Mn 2p3/2 of Mn3+ species, respectively, while the peaks around 653.5 eV and 642.1 are assigned to those of Mn4+ species [23]. Similar to CeCo-400 samples, the Co2+ and Co3+ are also indicated in Co 2p (Figure 2e), Olatt, Ovac, and Oads are indicated in O 1s spectra (Figure 2f). Moreover, the results calculated based on XPS spectra are summarized in Table 1. The molar ratios of Mn3+/Mn4+ followed the order of MnCo11-400 (1.20) > MnCo12-400 (1.17) > MnCo21-400 (0.91). A high ratio of Mn3+ in catalysts weakens Mn-O bonds on the surface, facilitating oxygen release for oxidation reactions and enhancing the dissociation and activation of surrounding oxygen atoms. The proportion of Mn3+ is also in agreement with the concentration of surface oxygen vacancies (Ovac%), which increase to maintain electrostatic balance as the Mn3+ content rises. Consequently, the catalytic properties are governed by the degree of charge-transfer imbalance, the relative concentration of Mn3+/Mn4+ redox couples, and the oxygen vacancy density. The Ce3+/Ce4+ and Ovac% of Ce-Co oxides are lower than that of Mn-Co oxides, which may be one of the reasons for the lower catalytic performance.
Figure 3 shows the surface morphologies of the Ce-Co, Mn-Co oxides calcined at 400 °C. For the Ce-Co oxides calcined at 300 °C (Figure S3), they crystallized into regular-shaped truncated rhombohedral prisms with a different size. The mean length along the long edge of CeCo12-300 is about 1 μm, exceeding those of CeCo11-300 (≈0.5 μm) and CeCo21-300 (≈0.2 μm). Exposure to 400 °C during calcination led to a significant transformation in the morphology of CeCo-400 and MnCo-400, resulting in the partial collapse of the initial rhombic structure and the subsequent development of porous aggregated nanoparticles. The calcination-induced transformation from initial rhombic structures to porous aggregated nanoparticles is beneficial to the enhanced catalytic performance for benzene oxidation, which was in agreement with the experimental results (Figure 4a and Figure S3). On the one hand, the hierarchical pore network facilitates efficient reactant diffusion and accessibility to active sites [38]. Nanoparticle interfacial boundaries create defect-rich zones with a higher oxygen vacancy concentration, promoting oxygen spillover at the interfaces of the catalyst for accelerated redox cycling [39]. Moreover, the aggregated architecture provides exceptional thermal stability while optimizing microwave coupling for energy-efficient oxidation [21,40]. The N2 adsorption–desorption isotherm and the pore size distribution of CeCo-400 and MnCo-400 catalysts are plotted in Figure 4. The N2 adsorption–desorption isotherms of the catalysts exhibit a combination of Type I and IV characteristics, indicative of a hierarchical microporous/mesoporous structure. And the presence of mesopores is confirmed through the observation of hysteresis loops in the adsorption/desorption branches and the absence of an adsorption plateau at high relative pressures. Mesopores primarily originate from interparticle voids [41,42]. The specific BET surface area (SBET) and pore data are presented in Table 2. The SBET of CeCo-400 is slightly higher than that of MnCo-400, while larger mesopores were observed in MnCo-400 (20~22 nm) with a larger pore volume.

2.2. Catalytic Performance of Benzene Oxidation

Figure 5 and Figure S4 show the catalytic performance of Ce-Co, Mn-Co oxides in the temperature range of 150–375 °C under the traditional heating method. For all reactions, benzene was completely decomposed into H2O and CO2 without other byproducts of incomplete oxidation via a time-based in situ FTIR with a gas cell. Compared with Ce-Co oxides calcined at 300 and 400 °C, the activity of CeCo-400 is superior to CeCo-300 in general, while the CeCo21-300 is slightly higher than CeCo21-400 in the test range, which is mainly attributed to the enhanced crystallinity of 400 °C calcined samples. In contrast, CeCo12-400 exhibits the highest activity, suggesting that a higher Co ratio for Ce-Co oxides is beneficial to benzene catalytic activity. Compared with CeCo-400 and MnCo-400, the catalytic activity of MnCo-400 is higher than that of the corresponding CeCo-400 samples, which may be attributed to the formation of the spinel structures. The superior activity of spinel structures may be related to their unique atomic-scale cooperativity, where the redox synergy between Mn3+/Mn4+ and Co2+/Co3+ at octahedral sites facilitates electron transfer during benzene C-H bond scission [38,43], and three-dimensional oxygen diffusion channels between tetrahedral and octahedral sublattices enable rapid lattice oxygen mobility [44,45]. MnCo11-400 exhibited the highest low-temperature activity and achieved the complete benzene oxidation at 250 °C. For control, tests without a catalyst were conducted under identical conditions. Benzene degradation is negligible (~3.5%) even at 375 °C, confirming that thermal effects alone cannot explain the high conversions achieved with catalysts. What is more, the on-stream reaction stability of CeCo-400 and MnCo-400 was measured at 225 °C for 20 h, and there was no significant decline in the benzene conversion of all the catalysts, indicating high stability of the catalysts (Figure 5b).

2.3. Microwave-Enhanced Catalytic Performance of Benzene Oxidation

Microwave irradiation is effective for the rapid temperature rise of the catalysts. Figure 6a shows temperature–time profiles of MnCo11-400, for example, while other samples show a similar heating behavior. The heating behavior at each power showed a constant and regular temperature rising. The temperature of MnCo11-400 reached a steady state after nearly 10 min microwave irradiation, and the steady-state temperature increased with the power increasing. Figure 6b shows the temperature–power profiles of the CeCo-400 and MnCo-400 samples. The catalyst temperature can be easily controlled according to the microwave output for catalytic reactions under different conditions. MnCo-400 shows a higher temperature than CeCo-400 via microwave irradiation of the same power, suggesting that the microwave heating behavior can be enhanced by adjusting the elemental composition and atomic ratio. The temperature disparities observed between MnCo-400 and CeCo-400 under identical microwave power originate from divergent dielectric loss properties (Figure 6b). MnCo2O4 exhibits a high loss tangent, enabling efficient microwave-to-thermal energy conversion via small polaron hopping (Co2+↔Co3+) and Maxwell–Wagner interfacial polarization at grain boundaries [46,47]. In contrast, the lower tan δ values of constituent phases CoO and CeO2 reflect weaker electromagnetic coupling, primarily limited to ionic polarization and oxygen vacancy defect oscillations [48,49].
Figure 7 shows the catalyst activities under microwave heating. Since the catalysts can be heated well through microwaves, especially Mn-Co oxides, microwave-assisted benzene oxidation shows high activity. What is more, our previous works also demonstrated that microwave irradiation can facilitate lattice oxygen activity, which is important for benzene oxidation on Mn-based oxides via the MvK mechanism [21,22]. At each output power, benzene was completely transformed into CO2 and H2O without the generation of other by-products, and the benzene conversion increased with the output power rise for all the catalysts. Energy is efficiently converted into heat under microwave irradiation, providing the necessary energy for the reaction and enabling complete benzene conversion at relatively low power levels (20–100 W). The Mn-Co oxides exhibited higher activity than Ce-Co oxides, as shown in Figure 6a. Among them, MnCo11-400 shows the optimal catalytic performance with the benzene conversion of 78% at 30 W, and 100% at 50 W, which is attributed to the high activity and heating performance. Figure 6b demonstrates the stability of MnCo11-400 over three microwave power on-off cycles, with each test lasting about 1 h. The conversion remained stable at approximately 80% under 30 W microwave irradiation, indicating high stability and insensitivity to microwave intermittent operations of the catalysts.
Figure 8 compares the catalytic activity of benzene oxidation on Ce-Co, Mn-Co oxides under microwave and conventional heating modes. Figure 8b plots the reaction temperature difference (ΔT = TE − TM) at benzene conversions of 10, 30, 50, 70, and 90% for all catalysts. The electric heating temperatures (TE) were recorded via embedded K-type thermocouples in the catalyst bed, while microwave heating temperatures (TM) were measured via infrared pyrometry calibrated against thermocouples under reaction gas flow. The temperature required for the reactions on the catalysts is quite lower under microwave heating for the same conversion, where the ∆T is over 100 °C between them, and the highest ∆T is even over 250 °C for CeCo11-400. The selective heating nature of microwaves enables efficient energy absorption at defect sites within metal oxide catalysts, creating localized hotspots [11,12,13]. This induces a non-equilibrium thermal state at which the catalyst surface temperature substantially exceeds the bulk temperature. While traditional thermal catalysis necessitates uniform bulk heating to reach reaction temperatures at catalytic sites, microwave catalysis achieves targeted energy deposition at active sites. This facilitates high-efficiency reactions under markedly reduced thermal conditions, drastically cutting energy consumption.
Taking advantage of rapid microwave heating and the high adsorption performance of the catalysts, a combination of adsorption and microwave-assisted oxidation processes was investigated for benzene removal with low concentrations, which was energy coasting using the traditional thermal oxidation method. As shown in Figure 9a, 100 ppm benzene was pre-adsorbed on MnCo11-400 for 60 min and then subjected to microwave irradiation at 30 W with a benzene-free O2/N2 stream. The temperature of the catalyst increased rapidly after microwave irradiation, resulting in the formation of CO2 with a small amount of benzene spilling. The total amount of benzene decomposition was 1.15 mmol, consistent with the amount of CO2 generation (6.8 mmol) for 20 min. Thus, in the combined process, low-concentration benzene was pre-adsorbed onto the catalyst at room temperature and then desorbed from the catalyst surface in a short time and completely oxidized to CO2 via microwave irradiation. For comparison, conventional electric heating at 200 °C for 30 min achieves benzene decomposition of 1.35 mmol, with CO2 yield stoichiometrically consistent at 8.1 mmol (Figure 9b). Under equivalent reaction conditions, microwave–catalytic treatment via three intermittent 20-min cycles (total 60 min) attains comparable benzene decomposition to 75 min of continuous electric heating. It demonstrates that low-concentration benzene streams can be processed more efficiently through microwave irradiation, establishing an energy-saving approach for enhanced decomposition efficiency.

3. Materials and Methods

3.1. Catalyst Preparation

CeCo-MOF and MnCo-MOF precursors were prepared using the hydrothermal method. Specifically, for the mole ratio of 1:1, 1.5 mmol of Ce(NO3)3·6H2O (or Mn(NO3)2·6H2O) and 1.5 mmol of (CH3COO)2Co·4H2O were dissolved with 40 mL DMF, and then 0.67g of 1,4-dicarboxybenzene was dissolved in 40 mL DMF to form uniform solutions under stirring for 30 min at first, respectively. All the reagents are purchased from Aladdin in Shanghai, China. The above solutions were mixed evenly and transferred to a 100 mL Teflon-lined autoclave and reacted at 80 °C for 24 h. Then, followed by filtering and washing with DMF and ethanol three times, respectively, the obtained powders were dried at 60 °C for 24 h. Ce-Co, Mn-Co oxides were further obtained with the CeCo-MOF and MnCo-MOF calcination at 400 °C for 3 h, denoted as CeCo11-400 and MnCo11-400. For catalysts with other metal ratios (1:2 and 2:1), the synthesis follows the same method, and the naming convention also applies, with the total metal ion amount fixed at 3 mmol.

3.2. Characterization

The crystal phases of the samples were detected via the X-ray diffraction (XRD, Bruker D8 Advance, Berlin, Germany) with Cu Kα radiation (λ = 0.1541 nm) operated at 40 kV and 40 mA. A particle size analysis was performed using Scherrer’s equation (1) applied to the dominant diffraction peaks of each crystalline phase. The surface morphology of the samples was observed via Verios 5 XHR with scanning electron microscopy (SEM, Thermo Fisher Scientific Inc., Waltham, MA, USA). The surface elemental states and absorbed species were characterized through X-ray photoelectron spectroscopy (XPS, Thermo Kalpha, Waltham, MA, USA). N2 adsorption isotherm was measured using Micromeritics ASAP 2460 (Micromeritics, Norcross, GA, USA) at 77 K after degassing at 473 K for 2 h. The specific surface area (SBET) of the samples was estimated via the Brunauer–Emmett–Teller (BET) model.
D = K λ β cos θ
where the following applies: D, crystallite size (nm); K, shape factor; λ: Cu-Kα radiation; β: corrected FWHM (radians); and θ, Bragg angle.

3.3. Catalytic Reactions

For the case of the catalytic reaction under microwave heating, a 100 mg sample was set up in the quartz tube reactor located in the microwave generation equipment (Figure 10). The catalyst temperature was controlled by regulating the microwave power and measured using an IR thermometer. Before the catalytic reaction, the pretreatment of samples under microwave heating with a power range of 20–100 W was carried out in 20% N2 -O2 flow gas, and then the heating behavior of the samples was measured at different powers. The catalytic oxidation of benzene was carried out over the catalyst in a fixed-bed flow reactor. The reaction conditions were as follows: 400 ppm benzene–N2 balance, 20% O2; gas flow rate, 100 mL·min−1; and weight hourly space velocity (WHSV), 60,000 mL·g−1·h−1. The evolution of reactants was recorded through in situ Fourier transform infrared spectroscopy. The reaction under conventional heating was carried out using the same reactor placed in an outer heating cylindrical furnace with the same conditions as microwave heating.
The benzene conversion was calculated according to the following equations:
B e n z e n e   C o n v e r s i o n , X   = ( C 6 H 6 ) i n ( C 6 H 6 ) o u t ( C 6 H 6 ) i n × 100 %

4. Conclusions

This work demonstrates that MOF-derived Mn-Co spinel oxides exhibit exceptional microwave–catalytic performance for benzene oxidation, attributable to synergistically engineered dielectric properties and spinel-facilitated oxygen mobility enhancement. The key findings reveal the following. (1) Low-temperature efficiency: Complete benzene conversion is achieved at a low microwave power of 50 W on MnCo11-400. Microwave-specific localized hotspots significantly lower the reaction temperature of 100–250 °C compared to conventional heating, reducing energy consumption. (2) Structural stability: MnCo11-400 maintains more than 80% conversion over repeated power cycling (3 × 1 h on/off), confirming robustness under intermittent operation. (3) Process innovation: An adsorption-rapid microwave oxidation synergy strategy mineralizes pre-adsorbed low-concentration benzene (1.15 mmol) within 20 min at 30 W, achieving parity with 25 min continuous thermal treatment while cutting energy input by >20%. The tailored Mn-Co spinel structure, with a high defect density and mesoporosity (pore volume: 0.20 cm3·g−1), synergizes microwave absorption and catalytic function. This work establishes MOF-derived spinel oxides as high-efficiency microwave catalysts for sustainable VOC abatement, with direct implications for low-energy emission control technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30163388/s1, Figure S1: XRD patterns of CeCo12-300, CeCo11-300, and CeCo21-300; Figure S2: XPS survey scan spectrum of (a) CeCo-400, (b) MnCo-400; Figure S3: SEM images of (a) CeCo12-300, (b) CeCo11-300, (c) CeCo21-300; Figure S4: Catalytic activity of CoMn-300; Table S1: The crystallite size of the catalysts.

Author Contributions

Conceptualization, S.D.; methodology, P.Z. and Z.L.; formal analysis, Z.L. and C.W.; investigation, L.W.; data curation, P.Z.; writing—original draft preparation, S.D.; writing—review and editing, S.D. and S.L.; funding acquisition, S.D. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the program of the Natural Science Foundation of Hubei Province (2023AFB447), the Key Research and Development Program of Hubei Province (2024BCB082 and 2023BCB108), and the “Unveiling the List and Appointing the Leader” Project of East Lake High-Tech Development Zone (2023KJB208).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, D.; Chen, X.; Shu, C.; Huang, Y.; Zhang, P.; Xiao, B.; Zhou, D. Ca-decorated MoBOH as a promising adsorbent for CH2O, C6H6, C3H6O, and C2HCl3 removal at room temperature: A first-principle study. Appl. Surf. Sci. 2021, 563, 150233. [Google Scholar] [CrossRef]
  2. Azalim, S.; Brahmi, R.; Agunaou, M. Wash coating of cordierite honeycomb with Ce–Zr–Mn mixed oxides for VOC catalytic oxidation. Chem. Eng. J. 2013, 223, 536–546. [Google Scholar] [CrossRef]
  3. Mazzuca, G.; Ren, X.; Loughner, C.; Estes, M.; Crawford, J.; Pickering, K.; Weinheimer, A.; Dickerson, R. Ozone production and its sensitivity to NOx and VOCs: Results from the DISCOVER-AQ field experiment. Atmos. Chem. Phys. 2016, 16, 14463–14474. [Google Scholar] [CrossRef]
  4. Chang, Y. Neurotoxic effects of n-hexane on the human central nervous system: Evoked potential abnormalities in n-hexane polyneuropathy. J. Neurol. Neurosurg. Psychiatry 1987, 50, 269–274. [Google Scholar] [CrossRef]
  5. Kim, R.; Kim, K.; Lee, H. Gene expression profiles of human lung epithelial cells exposed to toluene. J. Toxicol. Environ. Health 2011, 4, 269–276. [Google Scholar] [CrossRef]
  6. Dumont, E.; Darracq, G.; Couvert, A.; Couriol, C.; Amrane, A.; Thomas, D.; Andrès, Y.; Cloirec, P. Volumetric mass transfer coefficients characterising VOC absorption in water/silicone oil mixtures. Chem. Eng. J. 2013, 221, 308–314. [Google Scholar] [CrossRef]
  7. Dunn, R.; El-Halwagi, M. Selection of optimal VOC-condensation systems. Waste Manag. 1994, 14, 103–113. [Google Scholar] [CrossRef]
  8. Liu, X.; Deng, J.; Xie, H. Catalytic removal of volatile organic compounds using ordered porous transition metal oxide and supported noble metal catalysts. Chin. J. Catal. 2016, 37, 1193–1205. [Google Scholar] [CrossRef]
  9. Tasbihi, M.; Calin, I.; Šuligoj, A. Photocatalytic degradation of gaseous toluene by using TiO2 nanoparticles immobilized on fiberglass cloth. J. Photochem. Photobiol. A 2017, 336, 89–97. [Google Scholar] [CrossRef]
  10. Guieysse, B.; Hort, C.; Platel, V.; Munoz, R.; Ondarts, M.; Revah, S. Biological treatment of indoor air for VOC removal: Potential and challenges. Biotechnol. Adv. 2008, 26, 398–410. [Google Scholar] [CrossRef]
  11. Einaga, H.; Nasu, Y.; Oda, M.; Saito, H. Catalytic performances of perovskite oxides for CO oxidation under microwave irradiation. Chem. Eng. J. 2016, 283, 97–104. [Google Scholar] [CrossRef]
  12. Varisli, D.; Korkusuz, C.; Dogu, T. Microwave-assisted ammonia decomposition reaction over iron incorporated mesoporous carbon catalysts. Appl. Catal. B Environ. 2017, 201, 370–380. [Google Scholar] [CrossRef]
  13. Bolotov, V.; Chesnokov, V.; Tanashev, Y.; Parmon, V. The oxidative dehydrogenation of ethane: Convectional vs. microwave heating of Ba-containing catalysts. Chem. Eng. Process 2018, 129, 103–108. [Google Scholar] [CrossRef]
  14. Chen, J.; Xu, W.; Zhu, J.; Wang, X.; Zhou, J. Highly effective direct decomposition of H2S by microwave catalysis on core-shell Mo2N-MoC@SiO2 microwave catalyst. Appl. Catal. B Environ. 2020, 268, 118454. [Google Scholar] [CrossRef]
  15. Zhang, X.; Hayward, D.O.; Mingos, D.M.P. Apparent equilibrium shifts and hot-spot formation for catalytic reactions induced by microwave dielectric heating. Chem. Commun. 1999, 11, 975–976. [Google Scholar] [CrossRef]
  16. Zhang, X.; Hayward, D.O.; Lee, C.; Mingos, D.M.P. Microwave assisted catalytic reduction of sulfur dioxide with methane over MoS2 catalysts. Appl. Catal. B Environ. 2001, 33, 137–148. [Google Scholar] [CrossRef]
  17. Durka, T.; Gerven, T.V.; Stankiewicz, A. Microwaves in heterogeneous gas-phase catalysis: Experimental and numerical approaches. Chem. Eng. Technol. 2009, 32, 1301–1312. [Google Scholar] [CrossRef]
  18. Gangurde, L.S.; Sturm, G.S.J.; Valero-Romero, M.J.; Mallada, R.; Santamaria, J.; Stankiewicz, A.I.; Stefanidis, G.D. Synthesis, characterization, and application of ruthenium-doped SrTiO3 perovskite catalysts for microwave-assisted methane dry reforming. Chem. Eng. Process 2018, 127, 178–190. [Google Scholar] [CrossRef]
  19. Dobosz, J.; Zawadzki, M. Total oxidation of lean propane over α-Fe2O3 using microwaves as an energy source. React. Kinet. Mech. Catal. 2015, 114, 157–172. [Google Scholar] [CrossRef]
  20. Nigar, H.; Julián, I.; Mallada, R. Microwave-assisted catalytic combustion for the efficient continuous cleaning of VOC-containing air streams. Environ. Sci. Techno. 2018, 52, 5892–5901. [Google Scholar] [CrossRef]
  21. Ding, S.; Zhu, C.; Hojo, H.; Einaga, H. Enhanced Catalytic Performance of spinel-type Cu-Mn Oxides for benzene oxidation under microwave irradiation. J. Hazard. Mater. 2022, 424, 127523. [Google Scholar] [CrossRef] [PubMed]
  22. Ding, S.; Hojo, H.; Einaga, H. Microwave-assisted removal of benzene with high efficiency on cobalt modified copper-manganese spinel oxides. J. Environ. Chem. Eng. 2022, 10, 108212. [Google Scholar] [CrossRef]
  23. Li, S.; Tu, L.; Li, X.; Chen, P.; Zhang, Y.; Ding, S. Investigation into the effect of MnO2 structure toward thermal/microwave-assisted catalytic oxidation of benzene. Appl. Surf. Sci. 2025, 707, 163649. [Google Scholar] [CrossRef]
  24. Pang, Y.; Lei, H. Degradation of p-nitrophenol through microwave-assisted heterogeneous activation of peroxymonosulfate by manganese ferrite. Chem. Eng. J. 2016, 287, 585–592. [Google Scholar] [CrossRef]
  25. Zhou, M.; Zhang, X.; Wei, J.; Zhao, S.; Wang, L.; Feng, B. Morphology controlled synthesis and novel microwave absorption properties of hollow urchinlike α-MnO2 nanostructures. J. Phys. Chem. C 2011, 115, 1398–1402. [Google Scholar] [CrossRef]
  26. Duan, Y.; Ma, H.; Li, X.; Liu, S.; Ji, Z. The microwave electromagnetic characteristics of manganese dioxide with different crystallographic structures. Physica B 2010, 405, 1826–1831. [Google Scholar]
  27. Chen, X.; Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In situ pyrolysis of Ce-MOF to prepare CeO2 catalyst with obviously improved catalytic performance for toluene combustion. Chem. Eng. J. 2018, 344, 469–479. [Google Scholar] [CrossRef]
  28. Zhang, X.; Hou, F.; Yang, Y.; Wang, Y.; Liu, N.; Chen, D.; Yang, Y. A facile synthesis for cauliflower like CeO2 catalysts from Ce-BTC precursor and their catalytic performance for CO oxidation. Appl. Surf. Sci. 2017, 423, 771–779. [Google Scholar] [CrossRef]
  29. Zhao, P.; Qin, F.; Huang, Z.; Sun, C.; Shen, W.; Xu, H. MOF-derived hollow porous Ni/CeO2 octahedron with high efficiency for N2O decomposition. Chem. Eng. J. 2018, 349, 72–81. [Google Scholar] [CrossRef]
  30. Yao, H.; Cai, S.; Yang, B.; Han, L.; Wang, P.; Li, H.; Yan, T.; Shib, L.; Zhang, D. In situ decorated MOF-derived Mn–Fe oxides on Fe mesh as novel monolithic catalysts for NOx reduction. New J. Chem. 2020, 44, 2357–2366. [Google Scholar] [CrossRef]
  31. Wang, Q.; Astruc, D. State of the art and prospects in metal–organic framework (MOF)-based and MOF-derived nanocatalysis. Chem. Rev. 2020, 120, 1438–1511. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Q.; Li, Z.; Bañares, M.A.; Weng, L.; Gu, Q.; Price, J.; Han, W.; Yeung, K. A novel approach to high-performance aliovalent-substituted catalysts-2D bimetallic MOF-derived CeCuOx microsheets. Small 2019, 15, 1903525. [Google Scholar] [CrossRef] [PubMed]
  33. Peng, R.; Sun, X.; Li, S.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Shape effect of Pt/CeO2 catalysts on the catalytic oxidation of toluene. Chem. Eng. J. 2016, 306, 1234–1246. [Google Scholar] [CrossRef]
  34. Yuan, X.; Ge, H.; Wang, X.; Dong, C.; Dong, W.; Riaz, M.S.; Huang, F. Controlled phase evolution from Co nanochains to CoO nanocubes and their application as OER catalysts. ACS Energy Lett. 2017, 2, 1208–1213. [Google Scholar] [CrossRef]
  35. Gan, G.; Yang, Z.; Li, Y.; Zhang, G. Efficient photothermal mineralization of toluene over MnCo2O4 with different exposed facets: Revealing the role of oxygen vacancy and photo-/thermo-synergistic mechanism. Appl. Catal. B Environ. 2024, 357, 124308. [Google Scholar] [CrossRef]
  36. Liotta, L.F.; Di Carlo, G.; Pantaleo, G.; Venezia, A.M.; Deganello, 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]
  37. Yang, R.; Guo, Z.; Cai, L. Investigation into the phase–activity relationship of MnO2 nanomaterials toward ozone-assisted catalytic oxidation of toluene. Small 2021, 17, 2103052. [Google Scholar] [CrossRef]
  38. Sadeghi, M.A.; Aghighi, M.; Barralet, J.; Gostick, J.T. Pore network modeling of reaction-diffusion in hierarchical porous particles: The effects of microstructure. Chem. Eng. J. 2017, 330, 1002–1011. [Google Scholar] [CrossRef]
  39. Yin, Z.H.; Zhang, W.P.; Luo, W.; Liu, H.; Wang, J.J. Oxygen spillover: Revolutionizing catalytic performance and disrupting mechanistic paradigms in sustainable chemistry. Adv. Mater. 2025, e11007. [Google Scholar] [CrossRef]
  40. Sani, Y.; Azis, R.A.S.; Ismail, I.; Yaakob, Y.; Mohammed, J. Enhanced electromagnetic microwave absorbing performance of carbon nanostructures for RAMs: A review. Appl. Surf. Sci. 2023, 18, 100455. [Google Scholar] [CrossRef]
  41. Li, L.; Liu, X.L.; Geng, H.Y.; Hu, B.; Song, G.W.; Xu, Z.S. A MOF/graphite oxide hybrid (MOF: HKUST-1) material for the adsorption of methylene blue from aqueous solution. J. Mater. Chem. A 2013, 1, 10292–10299. [Google Scholar] [CrossRef]
  42. Teixeira, R.S.; Schmidt, D.V.; dos Santos, F.S.; Cipriano, D.F.; Faria, D.N.; Freitas, J.C.; Pietre, M.K. Nanostructured faujasites with different structural and textural properties for adsorption of cobalt and nickel. Braz. J. Chem. Eng. 2024, 41, 1271–1283. [Google Scholar] [CrossRef]
  43. Song, L.; Zhang, H.; Nie, Z.; Tian, J.; Liu, Y.; Ma, C.; Ye, D. Ni Doping Promotes C–H bond activation and conversion of key intermediates for total oxidation of methane over Co3O4 catalysts. ACS Catal. 2023, 13, 15779–15793. [Google Scholar] [CrossRef]
  44. Meng, J. Understanding and Design of Interstitial Oxygen Conductors. Comput. Mater. Sci. 2025, 259, 114110. [Google Scholar] [CrossRef]
  45. Xing, Y.; Kim, I.; Kang, K.T.; Park, B.; Wang, Z.; Kim, J.C.; Oh, S.H. Atomic-scale operando observation of oxygen diffusion during topotactic phase transition of a perovskite oxide. Matter 2022, 5, 3009–3022. [Google Scholar] [CrossRef]
  46. Tan, S.; Zhang, Y.; Zhou, Z.; Guan, X.; Liao, Y.; Li, C.; Feng, G. Construction of Co2NiO4@MnCo2O4.5 Nanocomposites with multiple hetero-interfaces for enhanced electromagnetic wave absorption. Particuology 2023, 81, 86–97. [Google Scholar] [CrossRef]
  47. Wang, S.; Hou, Y.; Wang, X. Development of a stable MnCo2O4 cocatalyst for photocatalytic CO2 reduction with visible light. ACS Appl. Mater. Inter. 2015, 7, 4327–4335. [Google Scholar] [CrossRef]
  48. Baral, A.K.; Dasari, H.P.; Kim, B.K.; Lee, J.H. Effect of sintering aid (CoO) on transport properties of nanocrystalline Gd doped ceria (GDC) materials prepared by co-precipitation method. J. Alloys Compd. 2013, 575, 455–460. [Google Scholar] [CrossRef]
  49. Charles Vincent, V.; Elaiyaraja, P.; Senthil, S.; Antony Prabhu, A.; Ratchagar, V.; Saravanan, G.; Srinivasan, S. Enhanced dielectric stability and electromagnetic shielding efficiency of CdO/CeO2 nanocomposites for high-performance electronic applications. J. Mater. Sci. Mater. Electron. 2015, 36, 1287. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the (a) Ce-Co and (b) Mn-Co oxides.
Figure 1. XRD patterns of the (a) Ce-Co and (b) Mn-Co oxides.
Molecules 30 03388 g001
Figure 2. (a) Ce 3d, (b) Co 2p, and (c) O 1s XPS spectra of CeCo12-400, CeCo11-400, and CeCo21-400, (d) Mn 2p, (e) Co 2p, and (f) O 1s XPS spectra of MnCo12-400, MnCo11-400, and MnCo21-400.
Figure 2. (a) Ce 3d, (b) Co 2p, and (c) O 1s XPS spectra of CeCo12-400, CeCo11-400, and CeCo21-400, (d) Mn 2p, (e) Co 2p, and (f) O 1s XPS spectra of MnCo12-400, MnCo11-400, and MnCo21-400.
Molecules 30 03388 g002
Figure 3. SEM images of the as-prepared samples: (a) CeCo12-400, (b) CeCo11-400, (c) CeCo21-400, (d) MnCo12-400, (e) MnCo11-400, and (f) MnCo21-400.
Figure 3. SEM images of the as-prepared samples: (a) CeCo12-400, (b) CeCo11-400, (c) CeCo21-400, (d) MnCo12-400, (e) MnCo11-400, and (f) MnCo21-400.
Molecules 30 03388 g003
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution of the CeCo-400 and MnCo-400.
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution of the CeCo-400 and MnCo-400.
Molecules 30 03388 g004
Figure 5. Catalytic activity, (a) benzene conversion at different temperatures, and (b) reaction stability at 225 °C for 20 h of Ce-Co, Mn-Co oxides calcined at 400 °C.
Figure 5. Catalytic activity, (a) benzene conversion at different temperatures, and (b) reaction stability at 225 °C for 20 h of Ce-Co, Mn-Co oxides calcined at 400 °C.
Molecules 30 03388 g005
Figure 6. Microwave heating performance, (a) heating behavior of MnCo11-400, and (b) catalyst temperature at different powers.
Figure 6. Microwave heating performance, (a) heating behavior of MnCo11-400, and (b) catalyst temperature at different powers.
Molecules 30 03388 g006
Figure 7. (a) Microwave-assisted benzene oxidation performance at different powers and (b) stability of benzene oxidation reaction on MnCo11-400 at 30 W for 3 h.
Figure 7. (a) Microwave-assisted benzene oxidation performance at different powers and (b) stability of benzene oxidation reaction on MnCo11-400 at 30 W for 3 h.
Molecules 30 03388 g007
Figure 8. Catalytic performance comparison of 400 °C-calcined Ce-Co and Mn-Co oxides under microwave heating (M) versus conventional electric heating (E): (a) Conversion profiles as a function of temperature; (b) Temperature differentials (ΔT = TE − TM) at fixed benzene conversion levels.
Figure 8. Catalytic performance comparison of 400 °C-calcined Ce-Co and Mn-Co oxides under microwave heating (M) versus conventional electric heating (E): (a) Conversion profiles as a function of temperature; (b) Temperature differentials (ΔT = TE − TM) at fixed benzene conversion levels.
Molecules 30 03388 g008
Figure 9. (a) Adsorption–desorption and catalytic oxidation on MnCo11-400 (pre-adsorption of 100 ppm benzene in O2/N2 flow gas at 30 °C for 60 min; microwave power: 30 W) and (b) benzene oxidation under conventional electric heating at 200 °C. Catalyst weight: 100 mg; 100 ppm benzene -N2 balance: 20% O2; gas flow rate: 100 mL·min−1.
Figure 9. (a) Adsorption–desorption and catalytic oxidation on MnCo11-400 (pre-adsorption of 100 ppm benzene in O2/N2 flow gas at 30 °C for 60 min; microwave power: 30 W) and (b) benzene oxidation under conventional electric heating at 200 °C. Catalyst weight: 100 mg; 100 ppm benzene -N2 balance: 20% O2; gas flow rate: 100 mL·min−1.
Molecules 30 03388 g009
Figure 10. The image of reaction apparatus.
Figure 10. The image of reaction apparatus.
Molecules 30 03388 g010
Table 1. The ratio of ions based on XPS.
Table 1. The ratio of ions based on XPS.
SamplesCe3+/Ce4+Mn3+/Mn4+Oads/Olatta Ovac%
CeCo12-4000.15-0.2620.0
CeCo11-4000.22-0.0721.7
CeCo21-4000.20-0.0521.1
MnCo12-400-1.170.2028.5
MnCo11-400-1.200.8733.2
MnCo21-400-0.910.1924.4
a The concentration of oxygen vacancies (Ovac) evaluated through the following equation: Ovac% = A O v a c A O l a t t + A O v a c + A O a d s × 100%.
Table 2. Textural properties of Ce-Co and Mn-Co oxides.
Table 2. Textural properties of Ce-Co and Mn-Co oxides.
SamplesSBET/m2·g−1Pore Size Distribution/nmPore Volume/cm3·g−1
CeCo12-40072210.10
CeCo11-4009570.07
CeCo21-40010270.05
MnCo12-40081230.27
MnCo11-40065200.20
MnCo21-40062210.19
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

Li, S.; Zhao, P.; Liu, Z.; Wang, C.; Wang, L.; Ding, S. Microwave-Enhanced Catalytic Performance of Benzene Oxidation on MOF-Derived Mn/Ce-Co Oxides. Molecules 2025, 30, 3388. https://doi.org/10.3390/molecules30163388

AMA Style

Li S, Zhao P, Liu Z, Wang C, Wang L, Ding S. Microwave-Enhanced Catalytic Performance of Benzene Oxidation on MOF-Derived Mn/Ce-Co Oxides. Molecules. 2025; 30(16):3388. https://doi.org/10.3390/molecules30163388

Chicago/Turabian Style

Li, Shefeng, Pengyi Zhao, Ziyang Liu, Chang Wang, Linling Wang, and Siyu Ding. 2025. "Microwave-Enhanced Catalytic Performance of Benzene Oxidation on MOF-Derived Mn/Ce-Co Oxides" Molecules 30, no. 16: 3388. https://doi.org/10.3390/molecules30163388

APA Style

Li, S., Zhao, P., Liu, Z., Wang, C., Wang, L., & Ding, S. (2025). Microwave-Enhanced Catalytic Performance of Benzene Oxidation on MOF-Derived Mn/Ce-Co Oxides. Molecules, 30(16), 3388. https://doi.org/10.3390/molecules30163388

Article Metrics

Back to TopTop