Next Article in Journal
Influence of Different Types of Surfactants on the Flotation of Natural Quartz by Dodecylamine
Previous Article in Journal
Quantum Magnetism of the Iron Core in Ferritin Proteins—A Re-Evaluation of the Giant-Spin Model
Previous Article in Special Issue
Phosphorus Modification of Iron: Mechanistic Insights into Ammonia Synthesis on Fe2P Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Low-Noble-Metal Ru@CoMn2O4 Spinel Catalyst for the Efficient Oxidation of Propane

1
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(10), 2255; https://doi.org/10.3390/molecules29102255
Submission received: 12 March 2024 / Revised: 7 May 2024 / Accepted: 8 May 2024 / Published: 11 May 2024
(This article belongs to the Special Issue New Metal Catalysts for Sustainable Chemistry)

Abstract

:
Noble metals have become a research hotspot for the oxidation of light alkanes due to their low ignition temperature and easy activation of C-H; however, sintering and a high price limit their industrial applications. The preparation of effective and low-noble-metal catalysts still presents profound challenges. Herein, we describe how a Ru@CoMn2O4 spinel catalyst was synthesized via Ru in situ doping to promote the activity of propane oxidation. Ru@CoMn2O4 exhibited much higher catalytic activity than CoMn2O4, achieving 90% propane conversion at 217 °C. H2-TPR, O2-TPD, and XPS were used to evaluate the catalyst adsorption/lattice oxygen activity and the adsorption and catalytic oxidation capacity of propane. It could be concluded that Ru promoted synergistic interactions between cobalt and manganese, leading to electron transfer from the highly electronegative Ru to Co2+ and Mn3+. Compared with CoMn2O4, 0.1% Ru@CoMn2O4, with a higher quantity of lattice oxygen and oxygen mobility, possessed a stronger capability of reducibility, which was the main reason for the significant increase in the activity of Ru@CoMn2O4. In addition, intermediates of the reaction between adsorbed propane and lattice oxygen on the catalyst were monitored by in situ DRIFTS. This work highlights a new strategy for the design of a low-noble-metal catalyst for the efficient oxidation of propane.

1. Introduction

The exhaust emission of petrochemical and coal chemical industries, especially oil storage and transportation processes, usually contains light alkanes. Propane, as a kind of light alkane, is the most difficult class of volatile organic compounds (VOCs) to catalytically degrade. The total bond energy of the C-H bond of propane is as high as 4016 kJ/mol [1,2,3], much more than that of aromatic and oxygenic hydrocarbons in the field of printing, dyeing, and spraying industries which are easy to catalytically oxidize [4,5,6]. Consequently, the degradation of propane requires high-activity catalysts and high combustion temperatures [7,8,9,10]. Noble-metal catalysts have exhibited excellent performance in the low-temperature catalytic combustion process, wherein the catalysts with Pt and Pd as active components for light alkanes’ combustion have been widely reported [11,12,13,14]. However, the high price of Pt and Pd limit their wide utilization. In contrast, Ru can greatly reduce the production cost due to its low price (approximately half the price), and it has been proved to have high catalytic activity in many chemical reactions [15], such as the ammonia synthesis reaction [16], the olefin complex decomposition reaction [17], organic compound wet oxidation [18], the catalytic combustion of VOCs [19,20], etc. However, noble-metal catalysts easily sinter at high temperature. Improving the utilization rate of noble metals and enhancing stability are key to the development of Ru catalysts.
Transition-metal oxides used as carriers can effectively improve the activity and stability of catalysts. Relying on the structural formula of AB2O4, spinel oxides can accelerate the electron transfer between transition metals [21] and optimize the electronic structure of the active site [22], thus improving the redox performance. Mn- and Co-based spinel oxides are excellent catalysts for VOC combustion. Benjamin et al. successfully prepared Co-Mn-oxide spinel catalysts for CO oxidation at ambient temperature [23]. It has been reported that high-valence Mn in the cubic phase, including Mn3+ and Mn4+, facilitated the adsorption and dissociation of oxygen and electron migration, thus showing better oxygen reduction performance. Zhang [24] investigated toluene oxidation over a nanocrystalline Co3O4 catalyst and found that its unique structural properties gave it a relatively low Co-O bonding energy and a large number of lattice defects. With the advantages of rich valence states and active sites, and controllable composition and structure, spinel oxides are potential noble-metal carriers. In addition, the noble metal embedded in spinel oxides can modulate the electronic structure of the catalyst and promote the orbital hybridization between the noble metal and the carrier, thus affecting its catalytic performance and stability. Ru@NiCo2O4 has been successfully used in the oxygen evolution reaction, demonstrating that Ru can effectively modulate the metal valence and activity of a spinel [25]. Xiong et al. [26] referenced Pd-Co2NiO4 spinel catalysts for methane oxidation, and the deposition of PdOx species on the Co2NiO4 surface moderated the metal–carrier interaction with good stability. However, the Ru-loaded Co-Mn spinel for the catalytic combustion of light alkanes has not been reported before.
In this study, a Co-Mn spinel was used as a carrier to prepare low-noble-metal Ru spinel catalysts with a kind of typical light alkane (propane) to enhance the stability of the noble-metal catalyst. Meanwhile, the addition of the noble metal Ru is expected to increase the electronic fluidity and form rich lattice oxygen and adsorbed oxygen to improve the activity of the catalytic oxidation of propane. The aim is to develop catalysts with high activity, low loading, and good stability for the degradation of propane which meet the practical needs of industrial applications. Meanwhile, XRD, XPS, In situ DRIFT analysis, and other catalyst characterization tools were used to establish the conformational relationship and study the reaction mechanism.

2. Results and Discussion

2.1. Catalytic Performance

The influence of Ru loading methods and Ru concentration on the catalytic combustion of propane were examined. It can be seen from Figure 1a that the introduction of Ru obviously improved the activity of propane catalytic combustion. The T90 (temperature under 90% propane conversion) of 0.1% Ru@CoMn2O4 and 0.1% Ru/CoMn2O4 was decreased by 45 °C and 13 °C, respectively, compared with that of CoMn2O4. Furthermore, the oxidation activity of the in situ doped 0.1% Ru@CoMn2O4 catalysts was significantly higher than that of the 0.1% Ru/CoMn2O4 catalyst obtained by the impregnation method. This may be because the in situ doped Ru was not only deposited on the surface of the catalyst but also dispersed inside the spinel, which resulted in higher activity and better stability. In a word, the Ru@CoMn2O4 catalyst displayed excellent catalytic activity, even better than some other noble-metal catalysts reported in recent years [27,28,29,30,31,32,33], indicating that the Ru@CoMn2O4 catalyst has potential for practical industrial applications in alkane removal. The catalyst activities in some of the relevant literature are shown in Table 1. It is worth noting that T90 of the Pt-Co3O4 NPs/PFs catalyst in ref. [30] was only 184 °C. This was due to its low GSHW (30,000 mL g·cat−1·h−1) and high Pt load (0.2 wt %). In addition, as shown in Figure 1b, the T90 of 0.1% Ru@CoMn2O4 was only 6 °C higher than that of 1% Ru@CoMn2O4. Considering the preparation of low noble-metal content, the subsequent physicochemical properties of the catalysts were investigated by using 0.1% Ru@CoMn2O4. The products of the propane oxidation were detected online with the DX4000 infrared analyzer (Gasmet, Vantaa, Finland). CO and other products were barely detected, at less than 1%. The CO2 selectivity of the catalysts is shown in Figure 1c. The CO2 selectivity of 0.1%Ru@CoMn2O4 approached 100%. It can be concluded that the 0.1%Ru@CoMn2O4 catalyst had excellent selectivity for propane oxidation. The stability of 0.1%Ru@CoMn2O4 is shown in Figure 1d. After 72 h of continuous reaction, C3H8 conversion was nearly constant, and no harmful by-products were detected, indicating that the 0.1%Ru@CoMn2O4 catalyst has high stability and potential for industrial applications in alkane removal.

2.2. Physical Characteristics

XRD was used to determine the crystal structure of the spinel catalysts. The wide-angle XRD spectra of the samples are shown in Figure 2. It can be seen that the peak positions of 0.1%Ru@CoMn2O4, 0.1%Ru/CoMn2O4, and CoMn2O4 were basically the same as those of the standard card (JCPDS No. 77-0471) of CoMn2O4. These results suggest that the acetates of cobalt and manganese were decomposed at high temperatures into metal oxides and formed spinel catalysts by complete phase transitions, rather than simple mixtures of metal oxides. Specifically, the diffraction peaks situated at 18.2°, 29.2°, 31.2°, 33.0°, 36.3°, 44.8°, 59.0°, 60.9°, and 65.2° corresponded to the (101), (112), (200), (103), (211), (220), (321), (224), and (400) crystal planes of the CoMn2O4 spinel structure [34]. No diffraction peak of the Ru species was detected in 0.1%Ru@CoMn2O4 and 0.1%Ru/CoMn2O4, most likely due to the relatively low loading and uniform dispersal of Ru. In addition, the doping of Ru reduced the intensity of the diffraction peak, suggesting that the crystallinity decreased after Ru doping.
SEM was conducted to examine the surface morphology of the catalysts, and the results are shown in Figure 3a–c. It can be seen that all of the catalysts were rod-like structures with lengths of about 500–600 nm, indicating that the doping of Ru did not significantly affect the morphology of the CoMn2O4 spinel. EDS mapping was conducted to further investigate the surface element distribution of 0.1%Ru@CoMn2O4. As shown in Figure 3d, Ru was uniformly dispersed on the CoMn2O4 spinel. In addition, the contents of O, Mn, Co, and Ru in 0.1%Ru@CoMn2O4 are illustrated in Figure 3e. The atomic ratio of Co/Mn was close to 1:2, which further verified the formation of the CoMn2O4 spinel. The content of Ru was 0.06%, slightly lower than the theoretical content.
Figure 4 shows the nitrogen adsorption–desorption isotherms and the pore size distribution of the samples measured at 77 K. As can be seen from Figure 4, all of the samples exhibited typical type IV isotherms (IUPAC classification) with a separate H3 hysteresis loop, implying the formation of stacked pore structures. The abundant mesopores of the samples provided good diffusion conditions for the catalytic oxidation of propane [35]. The specific surface area, average pore size, and pore volume were calculated according to the classical BET and BJH models, and the structural parameters of the catalysts are summarized in Table 2. It can be seen that the surface area was obviously decreased after Ru doping, indicating that the introduction of the Ru species had a negative impact on the destruction of the textural structure. The surface area of 0.1%Ru@CoMn2O4 was smaller than that of 0.1%Ru/CoMn2O4, making it clear that the in situ introduction of Ru had a greater effect than the impregnation method. According to the activity evaluation results mentioned above, the physical properties were not the main factors affecting the activity of the catalyst.

2.3. Surface Chemical Properties

The redox capacity of the catalysts had a great impact on the catalytic activity for propane oxidation. The redox properties of the samples were investigated by H2-TPR, and the results are shown in Figure 5. Three reduction peaks can be observed in the H2-TPR curve of the CoMn2O4 catalyst. It contains four reduction processes. The peaks at 248 °C and 413 °C were attributed to the reduction of Mn4+→Mn3+ and of Co3+→Co2+, respectively. The peak at 565 °C contains two overlapped processes, including the reduction of Mn3+→Mn2+ and of Co2+→Co0 [36]. In contrast, the reduction peaks of 0.1% Ru@CoMn2O4 and 0.1%Ru/CoMn2O4 were obviously shifted in the low-temperature direction. Although the peak of Ru was not detected, the results sufficiently show that the doping of Ru significantly improved the low-temperature redox capacity of the catalyst. This may be owing to the stronger synergistic effect between the highly electronegative Ru and CoMn2O4; hence, the electrons were likely to migrate rapidly between them through the bridge oxygen structure. In addition, the reduction temperature of 0.1% Ru@CoMn2O4, especially for the shifting of Mn4+ to Mn3+, was lower than that of 0.1%Ru/CoMn2O4. Therefore, it can be concluded that the good low-temperature redox performance of 0.1% Ru@CoMn2O4 contributed to the acceleration of the catalytic oxidation cycle and improved propane oxidation activity to some extent.
O2-TPD is an effective technique to study the desorption behavior of O2 on catalysts and is commonly used to evaluate the mobility and activation capacity of different oxygen species on catalysts. Typically, the capacity of oxygen desorption on the catalyst is O2(ad) > O2(ad) > O(ad) > O2−(lat) [37]. The adsorbed oxygen species (O2, O2, and O) are more readily desorbed than the lattice oxygen species [38]. Normally, oxygen species desorbed below 300 °C are physically/chemically adsorbed oxygen species on the catalyst surface, while oxygen species desorbed at higher temperatures (>300 °C) can be attributed to lattice oxygen. Based on the intensity of the oxygen desorption peaks (Figure 6), it can be seen that the content of lattice oxygen was 0.1%Ru@CoMn2O4 > 0.1%Ru/CoMn2O4 > CoMn2O4, and the temperature of the lattice oxygen desorption was shifted toward lower temperatures with the introduction of Ru, which suggests that Ru significantly enhanced the amount and mobility of lattice oxygen of the catalyst. It is possible that the highly electronegative Ru could promote the electron transfer from the cobalt–manganese spinel to Ru to form more lattice oxygen; at the same time, the doping of Ru caused the weakening of the bonding energies of Mn-O and Co-O, which increased the mobility of the lattice oxygens of CoMn2O4 and improved the catalytic oxidative performance of the lattice oxygen on propane [39,40,41,42]. In addition, the reduction temperature of 0.1% Ru@CoMn2O4, especially for the shifting of Mn4+ to Mn3+, was lower than that of 0.1%Ru/CoMn2O4. According to the results of H2-TPR and O2-TPD, it is concluded that the catalyst prepared by the in situ introduction of Ru had a better redox ability than that prepared by the impregnation method.
The surface state of the catalyst was analyzed by using the XPS technique to investigate the effect of cation substitution on the catalyst. Figure 7a demonstrates the high-resolution spectrum of Mn 2p3/2; the peaks at 640.8–641.0 eV, 642.0–642.4 eV, and 644.1–644.6 eV were attributed to Mn2+, Mn3+, and Mn4+, respectively [2,29,40]. The content of Mnn+ was calculated based on the fitted peak areas, and the results are summarized in Table 3. It can be observed that all of the samples were dominated by Mn3+. In addition, it is worth noting that the content of Mn4+ was significantly higher in the 0.1% Ru@CoMn2O4 catalyst compared with CoMn2O4. In the spinel structure, Mn preferentially occupied the octahedral coordination site as Mn3+. Partial Mn3+ cations were oxidized into Mn4+ in the octahedral sites due to the Jahn–Teller effect, which seriously distorted the localized lattice structure. As a result, the Mn-O bond energy became weaker, and the lattice oxygen easily participated in the reaction with propane [43,44]. It has been reported that the ratio of Mn4+/Mn is considered to have an important effect on the oxidation of VOCs, as the high-valence manganese species contribute to the oxidation of the catalysts [45]. Therefore, it was speculated that the doping of Ru, with the partial electron transfer from the lower electronegative Mn (X = 1.55) to the higher electronegative Ru (X = 2.20), elevated the valence of Mn at the octahedral sites, which contributed to the mobility of the lattice oxygen and facilitated the enhancement of the catalytic efficiency for propane.
Figure 7b shows the high-resolution spectrum of Co 2p3/2. The peaks at 780.4 eV, 782.2 eV, and 786.5 eV were attributed to Co3+, Co2+, and the strong satellite peak associated with Co2+, respectively [46]. Based on the spinel structure, it is hypothesized that Co was predominantly present on the tetrahedral structure and Mn on the octahedral sites. The metal ions at the center of the octahedral sites were directly opposite to O-2p, with a higher orbital overlap; hence, they possessed a higher oxygen binding capacity compared with the tetrahedral sites. Therefore, the Mn ions in the octahedral sites may be the main active sites for the propane oxidation reaction, while the redox cycle between Co2+ and Co3+ promoted the activation of propane oxidation.
The O 1s spectrum could be divided into two types of oxygen: lattice oxygen (Olat) at 530.0 eV and chemisorbed oxygen (Oad) at 531.5 eV [47]. The results of the split-peak fitting of the two oxygen species (Figure 7c and Table 3) showed that the Olat/Oad ratios of CoMn2O4, 0.1%Ru/CoMn2O4, and 0.1%Ru@CoMn2O4 were 1.17, 1.27, and 1.36, respectively. Consistently with the previous conclusions of O2-TPD, it was regarded that the doping of Ru increased the amount of lattice oxygen. In addition, 0.1%Ru@CoMn2O4 possessed more lattice oxygen, not only loaded onto the spinel surface but also embedded in the crystal lattice.

2.4. Reaction Mechanism

The intermediate species of the reaction between the adsorbed propane and lattice oxygen on the catalyst were monitored by in situ DRIFTS. Figure 8a shows the in situ DRIFT spectrum of propane adsorbed on the 0.1% Ru@CoMn2O4 catalyst at 50 °C. The peaks at 2967, 1402, and 1212 cm−1 were attributed to the C-H stretching vibration, C-H bending vibration, and C-C stretching vibration, respectively [27,48], indicating that propane was adsorbed on the catalyst surface. The peak at 1633 cm−1 was attributed to C=C stretching or adsorbed water [3,27,31,49,50]. The peak at 1054 cm−1 was attributed to the C-O stretching vibration of alcohols [27,29], which showed that propane could be partially oxidized to alcohols by lattice oxygen. The broad spectrum at 3200–3500 cm−1 was attributed to hydroxyl groups which might be from alcohol, carboxyl, or adsorbed water [31,32,51]. After 30 min of propane adsorption, C3H8 was cut off, and only N2 was left. The intermediate products were detected by increasing the temperature. As can be seen from Figure 8b, with the increase in the temperature, the peaks assigned to the C-O stretching vibration (1073 cm−1), C-C stretching vibration (1208 cm−1), C-H bending vibration (1402 cm−1), C-H stretching vibration (2872 cm−1, 2899 cm−1), and the C=C stretching (1622 cm−1) of propylene gradually weakened, while the vibrations of C=O (1249 cm−1), bidentate/uncoordinated CO32− (1515 cm−1), and acetate (1575 cm−1) appeared accordingly [27,29,31], implying that propane was further oxidized to carbonate and carboxylate species by lattice oxygen.
It has been reported that the oxidation of propane follows the Mars–van Krevelen (MvK) mechanism [27,52,53]; the process on 0.1%Ru@CoMn2O4 is hypothesized, as shown in Scheme 1. The highly electronegative Ru led to the transfer of electrons from Mn to Ru, which increased the valence state of Mn on the octahedral sites. Propane was adsorbed on the high-valence Mn cation (Mn3+, Mn4+) centers of the octahedral sites to form chemisorbed species, which reacted with the lattice oxygen associated with the high-valence Mn cations to obtain incomplete oxidation products (e.g., alcohols and carboxylic acids). At the same time, the high-valence Mn cations captured electrons and were reduced to low-valent Mn cations (e.g., Mn2+). Considering the redox couples of Mn3+/Mn2+ (Eθ = 1.54 V) and Co3+/Co2+ (Eθ = 1.92 V), Mn2+ tended to be oxidized by Co3+ [54], and lattice oxygen was transferred from the adjacent Co3+ center to the Mn2+ center to compensate for the loss of oxygen. Molecular oxygen could be adsorbed onto Co2+, where it was transformed into lattice oxygen. In this way, oxidation–reduction repeated indefinitely, allowing the reaction to continue until propane was completely oxidized to CO2 and H2O.

3. Experiment

3.1. Catalyst Preparation

Ru@CoMn2O4 was synthesized by following the sol–gel method. In a typical procedure, 4.90 g of manganese acetate (Aladdin, Shanghai, China, AR; 99.0%), 2.49 g of cobalt acetate (Aladdin, Shanghai, China, AR; 99.5%), and 0.0047 g of ruthenium trichloride (Macklin, Shanghai, China; 36.0–40.0%) were added to 150 mL of a mixed alcohol solution (ethanol:ethylene glycol = 3:1) (Macklin, Shanghai, China, AR; 99.5%) and stirred at 40 °C until completely dissolved to form solution A. A total of 4.5 g of Oxalic acid (Macklin, Shanghai, China, AR; 99.0%) was added to 50 mL of the above mixed alcohol solution and stirred until dissolved to form solution B. Solution B was added to solution A, and the mixture was stirred vigorously at 40 °C for 0.5 h. The precipitate was collected by centrifugation and dried at 80 °C for 24 h. The product was calcined at 300 °C for 3 h in an air atmosphere at a temperature rate of 5 °C/min and then kept at 240 °C for 1 h and at 80 °C for 1 h. Finally, the composite catalyst was obtained at room temperature, denoted as 0.1% Ru@CoMn2O4.
As a control sample, CoMn2O4 was prepared, referring to the method of 0.1% Ru@CoMn2O4, except for the absence of Ru. The other control sample (0.1% Ru/CoMn2O4) was prepared by using the impregnation method, impregnating the same mass fraction of Ru onto CoMn2O4.

3.2. Catalyst Activity Test

Catalytic activity was evaluated in a fixed-bed quartz reactor (id = 6 mm) at atmospheric pressure. The reaction gas consisted of 0.2% C3H8, 10% O2, and N2 as balance. A total of 0.1 g of the catalyst (40–60 mesh) was used in the reaction, and gas hourly space velocity (GHSV) was 60,000 mL g−1h−1. The temperature of the activity evaluation experiment ranged from 200 to 300 °C at the heating rate of 5 °C/min, and each temperature point was kept for 0.5 h. Stability tests were carried out at 250 °C under the same reaction conditions as the activity evaluation. The quantitative analysis of the reactants (C3H8) and products (CO2 and CO) was performed online by using the DX4000 infrared analyzer (Gasmet, Vantaa, Finland). C3H8 conversion was calculated by using the following equation:
X C 3 H 8 = C i n C o u t C i n × 100 %
where Cin and Cout represent the C3H8 concentrations assigned to the inlet and outlet, respectively.

3.3. Material Characterization

The structures were determined by using a Bruker D8 Advance X-ray diffractometer (XRD) (Bruker, Ettlingen, Germany) with Cu Ka radiation (40 mA, 40 kV). N2 adsorption–desorption isotherms were obtained by using a Quantachrome Auto Sorb iQ-MP (Quantachrome, Boynton Beach, FL, USA). The BJH model was used to evaluate the pore size distribution, and the BET multi-molecular adsorption model was used to evaluate the surface area. The surface morphology of the samples was examined by JSM-7001F field emission SEM (JEOL Ltd., Tokyo, Japan), while the elemental distribution was measured by EDS. H2-TPR and O2-TPD were conducted on a Quantachrome ChemStar (Quantachrome, Boynton Beach, FL, USA)with a thermal conductivity detector (TCD) to detect the signals. XPS data were collected by using AXIS ULTRA DLD photoelectron spectroscopy (Shimadzu, Kyoto, Japan) with Al Kα X-ray radiation (1486.6 eV).
In situ DRIFTS was performed with Tensor 27 (Bruker, Ettlingen, Germany). First, degassing was carried out at 300 °C for 1 h by using N2 as the desorption gas. After cooling, 2000 ppm C3H8/N2 was introduced for adsorption (25 mL·min−1) till saturation to detect the process of propane adsorption. Then, C3H8 was cut off, and the temperature was increased at the heating rate of 5 °C/min to detect the changes in the surface species of the catalyst in N2 atmosphere. The temperature ranged from 50 to 250 °C, and each temperature point was kept for 0.5 h till stabilization. Infrared spectral data of the samples were acquired by accumulating 64 scans at a resolution of 8 nm−1. All of the data were subjected to background subtraction.

4. Conclusions

In conclusion, Ru doping has brought a significant improvement in the propane catalytic performance of CoMn2O4 spinel catalysts. Specifically, 0.1%Ru@CoMn2O4 displayed excellent catalytic activity and stability. The T90 of 0.1%Ru@CoMn2O4 was 217 °C, lower than that of CoMn2O4. H2-TPR, O2-TPD, and XPS results show that the electronic structure of CoMn2O4 was obviously modulated by the charge redistribution with the introduction of Ru. More high-valence Mn4+ species were generated on the surface, and lattice oxygen species were easier to mobilize due to the charge transfer and synergy between Mn and Co, facilitating the adsorption and activation capacity for propane. In addition, in situ DRIFTS revealed another path with olefin species as the intermediates on 0.1%Ru@CoMn2O4 surface, which is conducive to a clearer understanding of the propane oxidation path. This work uncovers the electronic regulation effect of the Ru species doped in CoMn2O4 on C3H8 oxidation and provides guidelines for light-alkane removal applications. Overall, it provides a design strategy for novel catalysts with dual-active-component catalysts and extends the application of these catalysts in the important field of VOC elimination.

Author Contributions

Conceptualization, Y.C. and Z.Z.; methodology, Y.C. and Z.Z.; software, Y.C. and S.M.; validation, Y.C.; data curation, Y.C.; writing—original draft preparation, Y.C.; writing—review and editing, Z.Z., Y.H. and W.S.; supervision, W.S. and Z.H.; project administration, Z.H.; funding acquisition, Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the DNL Cooperation Fund, CAS (DNL 202004), and the Strategic Priority Research Program of Chinese Academy of Sciences (No. XDA29020501).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shan, C.P.; Wang, Y.C.; Li, J.B.; Zhao, Q.; Han, R.; Liu, C.X.; Liu, Q.L. Recent Advances of VOCs Catalytic Oxidation over Spinel Oxides: Catalyst Design and Reaction Mechanism. Environ. Sci. Technol. 2023, 476, 146550. [Google Scholar] [CrossRef]
  2. He, J.; Zheng, F.; Zhou, Y.; Li, X.; Wang, Y.; Xiao, J.; Li, Y.; Chen, D.; Lu, J. Catalytic oxidation of VOCs over 3D@2D Pd/CoMn2O4 nanosheets supported on hollow Al2O3 microspheres. J. Colloid Interface Sci. 2022, 613, 155–167. [Google Scholar] [CrossRef] [PubMed]
  3. Wu, Q.Q.; Yan, J.R.; Jiang, M.X.; Dai, Q.G.; Wu, J.Y.; Ha, M.N.; Ke, Q.P.; Wang, X.Y.; Zhan, W.C. Phosphate-assisted synthesis of ultrathin and thermally stable alumina nanosheets as robust Pd support for catalytic combustion of propane. Appl. Catal. B Environ. 2021, 286, 119949. [Google Scholar] [CrossRef]
  4. Moreno-Roman, E.J.; Gonzalez-Cobos, J.; Guilhaume, N.; Gil, S. Toluene and 2-propanol mixture oxidation over Mn2O3 catalysts: Study of inhibition/promotion effects by in-situ DRIFTS. Chem. Eng. J. 2023, 470, 144114. [Google Scholar] [CrossRef]
  5. Zhang, H.H.; Dai, L.Y.; Feng, Y.; Xu, Y.H.; Liu, Y.X.; Guo, G.S.; Dai, H.X.; Wang, C.C.; Wang, C.; Hsi, H.C.; et al. A Resource utilization method for volatile organic compounds emission from the semiconductor industry: Selective catalytic oxidation of isopropanol to acetone over Au/alpha-Fe2O3 nanosheets. Appl. Catal. B Environ. 2020, 275, 119011. [Google Scholar] [CrossRef]
  6. Song, S.Q.; Wu, X.; Lu, C.H.; Wen, M.C.; Le, Z.G.; Jiang, S.J. Solid strong base K-Pt/NaY zeolite nano-catalytic system for completed elimination of formaldehyde at room temperature. Appl. Surf. Sci. 2018, 442, 195–203. [Google Scholar] [CrossRef]
  7. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z.P. Recent Advances in the Catalytic Oxidation of Volatile Organic Compounds: A Review Based on Pollutant Sorts and Sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef] [PubMed]
  8. Jian, Y.F.; Tian, M.J.; He, C.; Xiong, J.C.; Jiang, Z.Y.; Jin, H.; Zheng, L.R.; Albilali, R.; Shi, J.W. Efficient propane low-temperature destruction by Co3O4 crystal facets engineering: Unveiling the decisive role of lattice and oxygen defects and surface acid-base pairs. Appl. Catal. B Environ. 2021, 283, 119657. [Google Scholar] [CrossRef]
  9. Zhang, S.; Liu, S.J.; Zhu, X.C.; Yang, Y.; Hu, W.S.; Zhao, H.T.; Qu, R.Y.; Zheng, C.H.; Gao, X. Low temperature catalytic oxidation of propane over cobalt-cerium spinel oxides catalysts. Appl. Surf. Sci. 2019, 479, 1132–1140. [Google Scholar] [CrossRef]
  10. Wang, Z.; Huang, Z.P.; Brosnahan, J.T.; Zhang, S.; Guo, Y.L.; Guo, Y.; Wang, L.; Wang, Y.S.; Zhan, W.C. Ru/CeO2 Catalyst with Optimized CeO2 Support Morphology and Surface Facets for Propane Combustion. Environ. Sci. Technol. 2019, 53, 5349–5358. [Google Scholar] [CrossRef]
  11. Fu, Q.J.; Wang, S.; Wang, T.; Xing, D.F.; Yue, X.; Wang, M.Z.; Wang, S.D. Insights into the promotion mechanism of ceria-zirconia solid solution to ethane combustion over Pt-based catalysts. J. Catal. 2022, 405, 129–139. [Google Scholar] [CrossRef]
  12. Ding, Y.Q.; Wu, Q.Q.; Lin, B.; Guo, Y.L.; Guo, Y.; Wang, Y.S.; Wang, L.; Zhan, W.C. Superior catalytic activity of a Pd catalyst in methane combustion by fine-tuning the phase of ceria-zirconia support. Appl. Catal. B Environ. 2020, 266, 118631. [Google Scholar] [CrossRef]
  13. Ding, Y.; Wang, S.; Zhang, L.; Lv, L.R.; Xu, D.K.; Liu, W.; Wang, S.D. Investigation of supported palladium catalysts for combustion of methane: The activation effect caused by SO2. Chem. Eng. J. 2020, 382, 122969. [Google Scholar] [CrossRef]
  14. Jiang, F.; Zeng, L.; Li, S.R.; Liu, G.; Wang, S.P.; Gong, J.L. Propane Dehydrogenation over Pt/TiO2-Al2O3 Catalysts. ACS Catal. 2015, 5, 438–447. [Google Scholar] [CrossRef]
  15. Okal, J.; Zawadzki, M. Combustion of propane over novel zinc aluminate-supported ruthenium catalysts. Appl. Catal. B Environ. 2011, 105, 182–190. [Google Scholar] [CrossRef]
  16. Honkala, K.; Hellman, A.; Remediakis, I.N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C.H.; Norskov, J.K. Ammonia synthesis from first-principles calculations. Science 2005, 307, 555–558. [Google Scholar] [CrossRef] [PubMed]
  17. Grubbs, R.H. Olefin metathesis. Tetrahedron 2004, 60, 7117–7140. [Google Scholar] [CrossRef]
  18. Hosokawa, S.; Hayashi, Y.; Imamura, S.; Wada, K.; Inoue, M. Effect of the Preparation Conditions of Ru/CeO2 Catalysts for the Liquid Phase Oxidation of Benzyl Alcohol. Catal. Lett. 2009, 129, 394–399. [Google Scholar] [CrossRef]
  19. Aouad, S.; Abi-Aad, E.; Aboukais, A. Simultaneous oxidation of carbon black and volatile organic compounds over Ru/CeO2 catalysts. Appl. Catal. B Environ. 2009, 88, 249–256. [Google Scholar] [CrossRef]
  20. Kamiuchi, N.; Mitsui, T.; Muroyama, H.; Matsui, T.; Kikuchi, R.; Eguchi, K. Catalytic combustion of ethyl acetate and nano-structural changes of ruthenium catalysts supported on tin oxide. Appl. Catal. B Environ. 2010, 97, 120–126. [Google Scholar] [CrossRef]
  21. Liu, S.; Ni, D.; Li, H.-F.; Hui, K.N.; Ouyang, C.-Y.; Jun, S.C. Effect of cation substitution on the pseudocapacitive performance of spinel cobaltite MCo2O4 (M = Mn, Ni, Cu, and Co). J. Mater. Chem. A 2018, 6, 11044. [Google Scholar] [CrossRef]
  22. Liu, Y.; Xiao, C.; Huang, P.; Cheng, M.; Xie, Y. Regulating the Charge and Spin Ordering of Two-Dimensional Ultrathin Solids for Electrocatalytic Water Splitting. Chem 2018, 4, 1263–1283. [Google Scholar] [CrossRef]
  23. Faure, B.; Alphonse, P. Co–Mn-oxide spinel catalysts for CO and propane oxidation at mild temperature. Appl. Catal. B Environ. 2016, 180, 715–725. [Google Scholar] [CrossRef]
  24. Zhang, W.; Díez-Ramírez, J.; Anguita, P.; Descorme, C.; Valverde, J.L.; Giroir-Fendler, A. Nanocrystalline Co3O4 catalysts for toluene and propane oxidation: Effect of the precipitation agent. Appl. Catal. B Environ. 2020, 273, 118894. [Google Scholar] [CrossRef]
  25. Peng, C.; Liu, H.; Chen, J.; Zhang, Y.; Zhu, L.; Wu, Q.; Zou, W.; Wang, J.; Fu, Z.; Lu, Y. Modulating the potential-determining step in oxygen evolution reaction by regulating the cobalt valence in NiCo2O4 via Ru substitution. Appl. Surf. Sci. 2021, 544, 148897. [Google Scholar] [CrossRef]
  26. Xiong, H.; Wiebenga, M.H.; Carrillo, C.; Gaudet, J.R.; Pham, H.N.; Kunwar, D.; Oh, S.H.; Qi, G.; Kim, C.H.; Datye, A.K. Design considerations for low-temperature hydrocarbon oxidation reactions on Pd based catalysts. Appl. Catal. B Environ. 2018, 236, 436–444. [Google Scholar] [CrossRef]
  27. Ma, S.; Hou, Y.; Li, Y.; Ding, X.; Yang, Y.; Tian, J.; Cui, Y.; Huang, Z. Regulation of A-site cations in AMn2Ox spinel catalysts on the deep oxidation of light alkanes VOCs. Fuel 2023, 334, 126785. [Google Scholar] [CrossRef]
  28. Feng, C.; Chen, C.; Wang, J.; Xiong, G.; Wang, Z.; Pan, Y.; Fei, Z.; Lu, Y.; Liu, Y.; Zhang, R.; et al. Total oxidation of propane in Ag-doped MnCeOx catalysts: The role of Ag species. Fuel 2023, 332, 126208. [Google Scholar] [CrossRef]
  29. Feng, C.; Wang, Y.; Chen, C.; Fu, X.; Pan, Y.; Xin, H.; Wang, Z.; Lu, Y.; Li, X.; Zhang, R.; et al. Fabrication of highly dispersed Pd-Mn3O4 catalyst for efficient catalytic propane total oxidation. J. Colloid Interface Sci. 2023, 650, 1415–1423. [Google Scholar] [CrossRef]
  30. Xiong, G.; Feng, C.; Chen, H.-C.; Li, J.; Jiang, F.; Tao, S.; Wang, Y.; Li, Y.; Pan, Y. Atomically Dispersed Pt-Doped Co3O4 Spinel Nanoparticles Embedded in Polyhedron Frames for Robust Propane Oxidation at Low Temperature. Small Methods 2023, 7, 2300121. [Google Scholar] [CrossRef]
  31. Feng, C.; Chen, C.; Xiong, G.; Yang, D.; Wang, Z.; Pan, Y.; Fei, Z.; Lu, Y.; Liu, Y.; Zhang, R.; et al. Cr-doping regulates Mn3O4 spinel structure for efficient total oxidation of propane: Structural effects and reaction mechanism determination. Appl. Catal. B Environ. 2023, 328, 122528. [Google Scholar] [CrossRef]
  32. Shan, C.; Zhang, Y.; Zhao, Q.; Li, J.; Wang, Y.; Han, R.; Liu, C.; Liu, Q. New insight into opposite oxidation behavior in acetone and propane catalytic oxidation over CoMn based spinel oxides. Chem. Eng. J. 2023, 476, 146550. [Google Scholar] [CrossRef]
  33. Liu, Y.; Hu, H.; Zheng, J.; Xie, F.; Gu, H.; Rostamnia, S.; Pan, F.; Liu, X.; Zhang, L. Interfacial engineering enables surface lattice oxygen activation of SmMn2O5 for catalytic propane combustion. Appl. Catal. B Environ. Energy 2023, 330, 122649. [Google Scholar] [CrossRef]
  34. Zhang, L.; Shi, L.; Huang, L.; Zhang, J.-P.; Gao, R.; Zhang, D. Rational Design of High-Performance DeNOx Catalysts Based on MnxCo3–xO4 Nanocages Derived from Metal–Organic Frameworks. ACS Catal. 2014, 4, 1753–1763. [Google Scholar] [CrossRef]
  35. Nie, L.H.; Meng, A.Y.; Yu, J.G.; Jaroniec, M. Hierarchically Macro-Mesoporous Pt/gamma-Al2O3 Composite Microspheres for Efficient Formaldehyde Oxidation at Room Temperature. Sci. Rep. 2013, 3, 3215. [Google Scholar] [CrossRef]
  36. Dong, C.; Qu, Z.P.; Qin, Y.; Fu, Q.; Sun, H.C.; Duan, X.X. Revealing the Highly Catalytic Performance of Spinel CoMn2O4 for Toluene Oxidation: Involvement and Replenishment of Oxygen Species Using In Situ Designed-TP Techniques. ACS Catal. 2019, 9, 6698–6710. [Google Scholar] [CrossRef]
  37. Tian, H.; He, J.; Liu, L.; Wang, D.; Hao, Z.; Ma, C. Highly active manganese oxide catalysts for low-temperature oxidation of formaldehyde. Microporous Mesoporous Mater. 2012, 151, 397–402. [Google Scholar] [CrossRef]
  38. Mo, J.; Zhang, Y.; Xu, Q.; Lamson, J.J.; Zhao, R. Photocatalytic purification of volatile organic compounds in indoor air: A literature review. Atmos. Environ. 2009, 43, 2229–2246. [Google Scholar] [CrossRef]
  39. Zhang, X.; Dai, L.Y.; Liu, Y.X.; Deng, J.G.; Jing, L.; Wang, Z.W.; Pei, W.B.; Yu, X.H.; Wang, J.; Dai, H.X. Effect of support nature on catalytic activity of the bimetallic RuCo nanoparticles for the oxidative removal of 1,2-dichloroethane. Appl. Catal. B Environ. 2021, 285, 119804. [Google Scholar] [CrossRef]
  40. Wang, J.; Dai, L.Y.; Deng, J.G.; Liu, Y.X.; Jing, L.; Hao, X.Q.; Pei, W.B.; Yu, X.H.; Rastegarpanah, A.; Dai, H.X. An investigation on catalytic performance and reaction mechanism of RuMn/meso-TiO2 derived from RuMn intermetallic compounds for methyl ethyl ketone oxidation. Appl. Catal. B Environ. 2021, 296, 120361. [Google Scholar] [CrossRef]
  41. Pei, Y.; He, W.; Wang, M.; Wang, J.; Sun, T.; Hu, L.; Zhu, J.; Tang, Y.; Wang, J. RuCo alloy trifunctional electrocatalysts with ratio-dependent activity for Zn-air batteries and self-powered water splitting. Chem. Commun. 2021, 57, 11872. [Google Scholar] [CrossRef] [PubMed]
  42. Li, L.; Bu, L.; Huang, B.; Wang, P.; Shen, C.; Bai, S.; Chan, T.-S.; Shao, Q.; Hu, Z.; Huang, X. Compensating Electronic Effect Enables Fast Site-to-Site Electron Transfer over Ultrathin RuMn Nanosheet Branches toward Highly Electroactive and Stable Water Splitting. Adv. Mater. 2021, 33, 2105308. [Google Scholar] [CrossRef]
  43. Hu, H.L.; He, H.C.; Xie, R.K.; Cheng, C.; Yan, T.R.; Chen, C.; Sun, D.; Chan, T.S.; Wu, J.P.; Zhang, L. Achieving reversible Mn2+/Mn4+ double redox couple through anionic substitution in a P2-type layered oxide cathode. Nano Energy 2022, 99, 107390. [Google Scholar] [CrossRef]
  44. He, M.; Ji, J.; Liu, B.; Huang, H. Reduced TiO2 with tunable oxygen vacancies for catalytic oxidation of formaldehyde at room temperature. Appl. Surf. Sci. 2019, 473, 934–942. [Google Scholar] [CrossRef]
  45. Zeng, L.; Song, W.; Li, M.; Zeng, D.; Xie, C. Catalytic oxidation of formaldehyde on surface of H-TiO2/H-C-TiO2 without light illumination at room temperature. Appl. Catal. B Environ. 2014, 147, 490–498. [Google Scholar] [CrossRef]
  46. Fan, Z.; Zhang, Z.; Fang, W.; Yao, X.; Zou, G.; Shangguan, W. Low-temperature catalytic oxidation of formaldehyde over Co3O4 catalysts prepared using various precipitants. Chin. J. Catal. 2016, 37, 947–954. [Google Scholar] [CrossRef]
  47. Chen, H.; He, J.; Zhang, C.; He, H. Self-assembly of novel mesoporous manganese oxide nanostructures and their application in oxidative decomposition of formaldehyde. J. Phys. Chem. C 2007, 111, 18033–18038. [Google Scholar] [CrossRef]
  48. Zhao, J.; Tu, C.; Sun, W.; Xia, H.; Zhang, H.; Dai, Q.; Wang, X. The catalytic combustion of CH2Cl2 over SO42--TixSn1-x modified with Ru. Catal. Sci. Technol. 2020, 10, 742–756. [Google Scholar] [CrossRef]
  49. Li, G.; Li, N.; Sun, Y.; Qu, Y.; Jiang, Z.; Zhao, Z.; Zhang, Z.; Cheng, J.; Hao, Z. Efficient defect engineering in Co-Mn binary oxides for low-temperature propane oxidation. Appl. Catal. B Environ. 2021, 282, 119512. [Google Scholar] [CrossRef]
  50. Zhu, W.; Chen, X.; Li, C.; Liu, Z.; Liang, C. Manipulating morphology and surface engineering of spinel cobalt oxides to attain high catalytic performance for propane oxidation. J. Catal. 2021, 396, 179–191. [Google Scholar] [CrossRef]
  51. Al-Abadleh, H.A.; Grassian, V.H. FT-IR study of water adsorption on aluminum oxide surfaces. Langmuir 2003, 19, 341–347. [Google Scholar] [CrossRef]
  52. Dong, T.; Liu, W.M.; Ma, M.D.; Peng, H.G.; Yang, S.Y.; Tao, J.X.; He, C.; Wang, L.; Wu, P.; An, T.C. Hierarchical zeolite enveloping Pd-CeO2 nanowires: An efficient adsorption/catalysis bifunctional catalyst for low temperature propane total degradation. Chem. Eng. J. 2020, 393, 124717. [Google Scholar] [CrossRef]
  53. Chai, G.T.; Zhang, W.D.; Liotta, L.F.; Li, M.Q.; Guo, Y.L.; Giroir-Fendler, A. Total oxidation of propane over Co3O4-based catalysts: Elucidating the influence of Zr dopant. Appl. Catal. B Environ. 2021, 298, 120606. [Google Scholar] [CrossRef]
  54. Liu, P.; Liao, Y.X.; Li, J.J.; Chen, L.W.; Fu, M.L.; Wu, P.Q.; Zhu, R.L.; Liang, X.L.; Wu, T.L.; Ye, D.Q. Insight into the effect of manganese substitution on mesoporous hollow spinel cobalt oxides for catalytic oxidation of toluene. J. Colloid Interface Sci. 2021, 594, 713–726. [Google Scholar] [CrossRef]
Figure 1. Effects of (a) different Ru loading methods and (b) loading content on catalytic activity of propane; (c) CO2 selectivity of catalysts (reaction conditions: 0.2 vol% C3H8, 10 vol% O2, N2 as balance gas, and GHSV = 60,000 mL g·cat−1 h−1); (d) stability of 0.1%Ru@CoMn2O4 catalyst (reaction conditions: 0.2 vol% C3H8, 10 vol% O2, N2 as balance gas, GHSV = 60,000 mL g·cat−1 h−1, and 250 °C).
Figure 1. Effects of (a) different Ru loading methods and (b) loading content on catalytic activity of propane; (c) CO2 selectivity of catalysts (reaction conditions: 0.2 vol% C3H8, 10 vol% O2, N2 as balance gas, and GHSV = 60,000 mL g·cat−1 h−1); (d) stability of 0.1%Ru@CoMn2O4 catalyst (reaction conditions: 0.2 vol% C3H8, 10 vol% O2, N2 as balance gas, GHSV = 60,000 mL g·cat−1 h−1, and 250 °C).
Molecules 29 02255 g001
Figure 2. The XRD patterns of the catalysts.
Figure 2. The XRD patterns of the catalysts.
Molecules 29 02255 g002
Figure 3. SEM images of (a) CoMn2O4, (b) 0.1%Ru@CoMn2O4, and (c) 0.1%Ru/CoMn2O4. (d) EDS energy spectrum data and (e) elemental line sweeps of 0.1%Ru@CoMn2O4.
Figure 3. SEM images of (a) CoMn2O4, (b) 0.1%Ru@CoMn2O4, and (c) 0.1%Ru/CoMn2O4. (d) EDS energy spectrum data and (e) elemental line sweeps of 0.1%Ru@CoMn2O4.
Molecules 29 02255 g003
Figure 4. (a) N2 adsorption–desorption curves; (b) pore size distributions of the samples.
Figure 4. (a) N2 adsorption–desorption curves; (b) pore size distributions of the samples.
Molecules 29 02255 g004
Figure 5. H2-TPR spectra of samples.
Figure 5. H2-TPR spectra of samples.
Molecules 29 02255 g005
Figure 6. O2-TPD spectra of samples.
Figure 6. O2-TPD spectra of samples.
Molecules 29 02255 g006
Figure 7. (a) Mn 2p, (b) Co 2p, and (c) O 1s XPS spectra of catalysts.
Figure 7. (a) Mn 2p, (b) Co 2p, and (c) O 1s XPS spectra of catalysts.
Molecules 29 02255 g007aMolecules 29 02255 g007b
Figure 8. In situ DRIFT spectra of (a) C3H8 adsorption over 0.1%Ru@CoMn2O4 at 50 °C and (b) adsorbed species changes at different temperatures (reaction conditions: (a) 0.2 vol% C3H8/N2; (b) N2 atmosphere).
Figure 8. In situ DRIFT spectra of (a) C3H8 adsorption over 0.1%Ru@CoMn2O4 at 50 °C and (b) adsorbed species changes at different temperatures (reaction conditions: (a) 0.2 vol% C3H8/N2; (b) N2 atmosphere).
Molecules 29 02255 g008
Scheme 1. Proposed mechanism for propane oxidation over 0.1%Ru@CoMn2O4.
Scheme 1. Proposed mechanism for propane oxidation over 0.1%Ru@CoMn2O4.
Molecules 29 02255 sch001
Table 1. Comparison of catalytic results for 0.1%Ru@CoMn2O4 catalyst with those reported in the literature.
Table 1. Comparison of catalytic results for 0.1%Ru@CoMn2O4 catalyst with those reported in the literature.
CatalystReaction ConditionsT90
(°C)
Ref.
NiMn2O40·05% C3H8, 5% O2, GHSV = 30,000 mL g·cat−1·h−1232[27]
Ag/MnCeOx−70·25%C3H8, 21%O2, GHSV = 30,000 mL g·cat−1·h−1242[28]
Pd-Mn3O40·25%C3H8, and balance air, at 30,000 mL g·cat−1·h−1240[29]
Pt-Co3O4 NPs/PFs0·25%C3H8, and balance air, at 30,000 mL g·cat−1·h−1184[30]
MnxCr3−xO40·25%C3H8, and balance air, at 30,000 mL g·cat−1·h−1264[31]
ZnCoMn spinel800 ppm C3H8, and balance air, at 30,000 mL g·cat−1·h−1360[32]
Co3O4/SmMn2O50·2%C3H8, 2%O2, 5% H2O, at 30,000 mL g·cat−1·h−1247[33]
0.1%Ru@CoMn2O40·2% C3H8, 10% O2, GHSV = 60,000 mL g·cat−1·h−1217This work
Table 2. The porous structural parameters and T90 of the samples.
Table 2. The porous structural parameters and T90 of the samples.
CatalystSBET
(m2/g)
Average Pore Size (nm)Pore Volume
(cm3/g)
CoMn2O4132.5714.130.47
0.1%Ru/CoMn2O4106.2711.540.31
0.1%Ru@CoMn2O482.9717.220.36
Table 3. XPS results of surface Co, Mn, and O.
Table 3. XPS results of surface Co, Mn, and O.
CatalystMn2+/MnMn3+/MnMn4+/MnCo2+/CoCo3+/CoOlat/OtotalOad/Ototal
CoMn2O40.310.540.150.470.530.540.46
0.1%Ru/CoMn2O40.270.490.240.540.460.560.44
0.1%Ru@CoMn2O40.150.580.270.480.520.590.41
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

Cui, Y.; Zeng, Z.; Hou, Y.; Ma, S.; Shen, W.; Huang, Z. A Low-Noble-Metal Ru@CoMn2O4 Spinel Catalyst for the Efficient Oxidation of Propane. Molecules 2024, 29, 2255. https://doi.org/10.3390/molecules29102255

AMA Style

Cui Y, Zeng Z, Hou Y, Ma S, Shen W, Huang Z. A Low-Noble-Metal Ru@CoMn2O4 Spinel Catalyst for the Efficient Oxidation of Propane. Molecules. 2024; 29(10):2255. https://doi.org/10.3390/molecules29102255

Chicago/Turabian Style

Cui, Yan, Zequan Zeng, Yaqin Hou, Shuang Ma, Wenzhong Shen, and Zhanggen Huang. 2024. "A Low-Noble-Metal Ru@CoMn2O4 Spinel Catalyst for the Efficient Oxidation of Propane" Molecules 29, no. 10: 2255. https://doi.org/10.3390/molecules29102255

Article Metrics

Back to TopTop