Next Article in Journal
Highly Efficient Au/ZnO−ZrO2 Catalysts for CO Oxidation at Low Temperature
Next Article in Special Issue
The Combined Impact of Ni-Based Catalysts and a Binary Carbonate Salts Mixture on the CO2 Gasification Performance of Olive Kernel Biomass Fuel
Previous Article in Journal
Metagenomic Type IV Aminotransferases Active toward (R)-Methylbenzylamine
Previous Article in Special Issue
Unveiling the Role of In Situ Sulfidation and H2O Excess on H2S Decomposition to Carbon-Free H2 over Cobalt/Ceria Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transition Metal (Fe2O3, Co3O4 and NiO)-Promoted CuO-Based α-MnO2 Nanowire Catalysts for Low-Temperature CO Oxidation

1
Collaborative Innovation Centre of the Atmospheric Environment and Equipment Technology, School of Environmental Science and Engineering, Nanjing University of Information Science & Technology, Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control, Nanjing 210044, China
2
College of Light Industry and Food Engineering, Nanjing Forestry University, Nanjing 210037, China
3
State Environmental Protection Key Laboratory of Atmospheric Physical Modeling and Pollution Control, China Energy Science and Technology Research Institute Co., Ltd., Nanjing 210023, China
4
Jiangsu Shuang Liang Environmental Technology Co., Ltd., Jiangyin 214400, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(3), 588; https://doi.org/10.3390/catal13030588
Submission received: 3 February 2023 / Revised: 10 March 2023 / Accepted: 10 March 2023 / Published: 15 March 2023

Abstract

:
As a toxic pollutant, carbon monoxide (CO) usually causes harmful effects on human health. Therefore, the thermally catalytic oxidation of CO has received extensive attention in recent years. The CuO-based catalysts have been widely investigated due to their availability. In this study, a series of transition metal oxides (Fe2O3, Co3O4 and NiO) promoted CuO-based catalysts supported on the α-MnO2 nanowire catalysts were prepared by the deposition precipitation method for catalytic CO oxidation reactions. The effects of the loaded transition metal type, the loading amount, and the calcination temperature on the catalytic performances were systematically investigated. Further catalyst characterization showed that the CuO/α-MnO2 catalyst modified with 3 wt% Co3O4 and calcined at 400 °C performed the highest CO catalytic activity (T90 = 75 °C) among the investigated catalysts. It was supposed that the loading of the Co3O4 dopant not only increased the content of oxygen vacancies in the catalyst but also increased the specific surface area and pore volume of the CuO/α-MnO2 nanowire catalyst, which would further enhance the catalytic activity. The CuO/α-MnO2 catalyst modified with 3 wt% NiO and calcined at 400 °C exhibited the highest surface adsorbed oxygen content and the best normalized reaction rate, but the specific surface area limited its activity. Therefore, the appropriate loading of the Co3O4 modifier could greatly enhance the activity of CuO/α-MnO2. This research could provide a reference method for constructing efficient low-temperature CO oxidation catalysts.

1. Introduction

As a toxic pollutant in the atmosphere, carbon monoxide (CO) has widely received public attention in recent years [1]. CO is not only mainly produced during the incomplete combustion of fossil fuels and motor vehicle emissions [2], but it is also a precursor of the ozone pollution [3], which can be harmful even at low concentrations. In recent years, scientists around the world have employed different methods for the removal of CO, such as photocatalysis [4], thermocatalysis oxidation [5], etc. Thermocatalysis oxidation is the most commonly used method for the oxidation of CO. The noble metal-based catalysts (such as Au [6,7], Pt [8,9,10], Pd [11], Ru [12,13], Rh [14] and Ag [15]), transition metal oxide based catalysts (such as MnO2 [16], CeO2 [17], ZrO2 [18], Co3O4 [16]), and metal hybrid catalysts (such as MnCo2O4 spinel [19], Ce-Zr solid solution [20]) can be used for the thermocatalytic oxidation of CO at low temperatures without generating the secondary contaminants [16]. Transition metal-based catalysts have good potential for application due to their low cost, high stability, and good activity [21,22]. Among the transition metal oxides, manganese dioxide (MnO2) has many advantages, such as a low price, being environmentally friendly, and abundance in nature [2,23]. As is well known, MnO2 usually exists in different crystalline phases with different structures, such as the α-, β-, and γ-types of the one-dimensional pore structure, the δ-type of the two-dimensional pore structure, and the λ-type of the three-dimensional network structure [24], which greatly depend on the different connectivity (corner-edge sharing) exhibited by the [MnO6] octahedra [25,26,27,28]. Liang et al. [28] prepared MnO2 nanorods with four different crystalline phases for the oxidation of CO. It was found that the order of oxidation activity of different crystalline phases with the same nanorod morphology was greatly different, in decreasing order from α- ≈ δ- > γ- > β-MnO2 (The temperature at which the conversion of CO on α-MnO2 reached 100% was approximately 130 °C). This indicated that the oxidation activity of CO significantly depended on the phase structure and channel structure of the MnO2. Tian et al. [29] investigated the effects of the crystalline phases of MnO2 (α-, β-, and ε-MnO2) on the performances of the oxidation of CO and toluene. It was found that the β-MnO2 performed the highest activity for CO oxidation (T90 = 75 °C) among the three crystalline phases of MnO2, and the α-MnO2 behaved with the lowest activity for CO oxidation (T90 = 118 °C). The content of oxygen vacancy in the catalyst was also determined by in situ EPR spectra and the results showed that the catalytic activity of the catalyst was proportional to the concentration of oxygen vacancies, which were regarded as the active sites for the adsorption and dissociation of oxygen molecules. It was believed that the catalytic activity of MnO2 was greatly related to the oxygen vacancy activity [29]. As is well known, the α-MnO2 is provided with the 1D (1 × 1) (2 × 2) tunnel structures, which are attributed to the tetragonal crystal system [30].
CuO has been widely used as the active site of the low-temperature catalytic oxidation of CO due to its excellent activity and abundant availability [31]. For example, Raziyeh Jokar et al. [31] used CuO/α-MnO2 as the catalyst of the preferential oxidation of CO in the hydrogen-rich gas stream and investigated the interaction between the MnO2 and CuO (T97 = 130 °C), the superior activity of the catalyst due to the beneficial synergistic interaction between CuO and α-MnO2. Meanwhile, the catalytic activity was also influenced by several factors, such as specific surface area, crystallinity, oxygen vacancies, and redox properties. Qian et al. [32] prepared a series of CuO/MnO2 catalysts with different CuO loading amounts by the incipient wetness impregnation method for the oxidation of CO. The catalyst activity was almost the same for the CuO loadings, from 1% to 40%. Sun et al. [33] prepared a CuO/Cu1.5Mn1.5O4 spinel-type composite oxide for synergistic catalysis of CO oxidation. It was found that the synergistic effect between Cu1.5Mn1.5O4 spinel and CuO can promote the oxidation of CO, and CuO-Cu1.5Mn1.5O4 had the best oxidation activity for CO (T100 = 177 °C).
Nowadays, for the modification of transition metal oxide-based catalysts, in addition to the carrier and active site, the promoter also plays an active role in the improvement of catalytic performance [34]. Gao et al. [35] doped transition metals (Fe, Co, Ni, and Cu) with a 1:10 molar ratio on α-MnO2 nanowires by a one-step hydrothermal method to oxidize CO. Among the four transition metals, Cu0.1MnOx had the best oxidation activity for CO (T100~120 °C). Krasimir et al. [36] investigated the effects of different molar ratios of chemical compositions on the γ-Al2O3-supported CuO/MnO2/Cr2O3 catalysts for the oxidation of CO, dimethyl ether (DME), and methanol. The results showed that the Cu-Mn-Cr/γ-Al2O3 catalyst, which Cu/(Mn + Cr) has a molar ratio of 2:1 and a Mn/Cr molar ratio of 0.25, can achieve the complete oxidation of CO at 200 °C.
In order to further investigate the contribution of promoters to the catalytic performance of CuO/MnO2 catalysts in the oxidation of CO, in this work, the α-MnO2 nanowire was successfully prepared by the hydrothermal method and used as the support for the CuO-based catalysts. A series of the transition metal oxides (Fe2O3, Co3O4, and NiO) promoted CuO-based α-MnO2 nanowire catalysts were prepared by the deposition precipitation method. The effects of the type, the loading amount, and the calcination temperature of three transition metal oxides on the performance of the catalytic oxidation of CO were systematically studied. The obtained catalysts were carefully characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), nitrogen physisorption, X-ray photoelectron spectroscopy (XPS), and H2 temperature-programmed reduction (H2-TPR). These catalysts were evaluated for their catalytic performances in the oxidation of CO. The results show that the Co3O4 (3 wt.%) promoted 10 wt.% CuO/α-MnO2 catalyst calcined at 400 °C performed the greatest CO reactivity (T90 = 75 °C).

2. Results and Discussion

2.1. XRD Analysis

The crystalline phase structures of the catalysts were obtained using X-ray diffraction (XRD) analysis. The results of XRD are shown in the Figure 1. Figure 1a shows the XRD patterns of the transition metal oxide (MOx = Fe2O3, Co3O4, NiO) doped catalysts and the pristine 10CuO/α-MnO2-400 catalyst. It can be observed in Figure 1a that all the catalysts show wide and clear diffraction peaks at 2θ = 12.78°, 18.11°, 25.71°, 37.52°, 41.97°, 49.86°, 56.37°, 65.11°, and 69.71°, which could be conformed to the characteristic peaks of the α-MnO2 (PDF#44-0141). It could be observed that the intensity of the diffraction peaks of MnO2 increased after the loading of the second transition metal oxides, especially over the 10CuO-3Co3O4/α-MnO2 catalyst. This phenomenon suggests that the crystallinity of the catalysts increased after loading the transition metal oxides, especially for the catalyst loading Co3O4. Meanwhile, two weak diffraction peaks could be detected at 2θ = 35.5° and 38.8°, which corresponded to the characteristic peaks of CuO, according to the standard card of PDF#05-0661. The diffraction peak intensity of CuO became weaker after the addition of the transition metal to the catalyst, indicating that the addition of the transition metal promoted the dispersion of CuO.
The XRD patterns of the 10CuO-yCo3O4/α-MnO2-400 catalysts with different contents of Co3O4 are shown in the Figure 1b. It can be observed that the diffraction peaks of CuO are very clear over the pristine 10CuO/α-MnO2-400 catalyst without Co3O4 loading. However, it can be observed that the diffraction peaks of CuO became blurry after loading Co3O4 from 1 wt.% to 10 wt.%. The possible reason accounting for this phenomenon was that the appropriate loading amount of Co3O4 could greatly enhance the dispersion of CuO on the surface of α-MnO2 nanowire catalyst. Nevertheless, the characteristic peaks of Co3O4 could be not observed in the catalysts due to the good dispersion of Co3O4.
The XRD patterns of 10CuO-3Co3O4/α-MnO2 catalysis calcined at different temperatures are shown in the Figure 1c. It can be recognized that the intensity of the diffraction peaks of α-MnO2 became weak at high calcination temperatures. The reason might be due to the formation of a copper manganese spinel (JCPDS No.70-0260) [37]. Only a diffraction peak of CuO was observed at 38.8°. Meanwhile, the intensity of the diffraction peaks of CuO became sharp with the increase of the calcination temperature (only from 200 °C to 400 °C). This indicated that the dispersion of CuO on the catalyst surface gradually deteriorated and the crystal size of CuO nanoparticles grew due to the finite surface area of the MnO2 nanowire from 200 °C to 400 °C. As the calcination temperature rose from 400 °C to 600 °C, the diffraction peak strength of copper manganese spinel increased. The diffraction peak strength of CuO decreased, indicating that the dispersion of CuO increased.

2.2. SEM Observation

The morphologies of the 10CuO-3MOx/α-MnO2-400 catalysts loaded with different transition metal oxides (MOx = Fe2O3, Co3O4, NiO) and the pristine 10CuO/α-MnO2-400 catalyst were characterized by SEM (Figure 2). It can be observed that the morphology of the α-MnO2 nanowire support greatly changed after loading the CuO active sites and the transition metal oxides. Specifically, the length to diameter ratio of the α-MnO2 nanowire significantly decreased compared to the pristine α-MnO2 nanowire. This was caused by the high loading contents of the CuO and transition metal oxides. In addition, the α-MnO2 nanowire support might experience thermal sintering and self-assemble at high calcination temperatures. The spatial distribution of the elements over the one-dimensional 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) nanowire catalyst was studied by energy dispersive X-ray spectroscopy mapping (EDS-mapping). As shown in Figure 2, the Cu and the doped Fe/Co/Ni elements were homogenously distributed over these investigated catalysts. This indicated that the great dispersion of the CuO and the doped transition metal oxides on the surface of the α-MnO2 nanowire supports could be facilely achieved by the precipitation deposition method.

2.3. BET Analysis

In order to further investigate the structural properties of the catalysts, the specific surface areas, pore volumes, and pore size distributions of the catalysts were measured by nitrogen physisorption measurements. As shown in Figure 3a, all catalysts show IV isotherms with H3-shaped hysteresis loops. These results prove the presence of mesopores with a narrow slit-shape in the catalyst [38]. It is also interesting to find that the 10CuO-3MOx/α-MnO2 catalysts still possess the similar mesoporous structure to the 10CuO/α-MnO2 nanowire catalysts after loading different transition metal oxides. This demonstrates that the 10CuO/α-MnO2 nanowire catalyst was provided with good thermal stability. The pore size distribution curves of the corresponding catalysts are shown in Figure 3b. The pore diameter of the catalysts is in the range of 2–15 nm after loading with transition metals. Moreover, the specific parameters of the structural properties of these catalysts are shown in Table 1. It can be observed that the specific surface areas of both the 10CuO-3Fe2O3/α-MnO2 and 10CuO-3NiO/α-MnO2 catalysts decreased after the loading of transition metals. On the contrary, the specific surface area of the 10CuO-3Co3O4/α-MnO2 catalysts increased substantially. This indicates that the Co3O4 on the α-MnO2 nanowire support surface was highly dispersed. Meanwhile, the average pore sizes of all the 10CuO-3MOx/α-MnO2 (MOx = Fe2O3, Co3O4, NiO) catalysts are very similar to the pristine 10CuO/α-MnO2 nanowire. Specifically, the 10CuO-3Fe2O3/α-MnO2 catalyst has a similar pore size distribution as the 10CuO/α-MnO2 catalyst. This indicates that their mesoporous structures are not sharply impaired by the loaded metal oxides. In addition, the pore capacities of all 10CuO-3MOx/α-MnO2 (MOx = Fe2O3, Co3O4, NiO) catalysts are enhanced compared to the pristine 10CuO/α-MnO2. As for the 10CuO-3Co3O4/α-MnO2 catalyst, its surface area was twice as large as the pristine 10CuO/α-MnO2 catalyst. The higher specific surface area is beneficial for the catalyst to expose more active sites, and the larger pore volume helps the reactant accelerate the reactant mass diffusion and has a better adsorption ability to the reactant [39].

2.4. XPS Analysis

The coordination, composition, and valence state of the elements over the catalyst surface were investigated by X-ray photoelectron spectroscopy (XPS). The XPS of Mn 2p, O 1s, and Cu 2p of the 10CuO/α-MnO2-400 and the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) are shown in Figure 4. It can be observed in Figure 4a that 10CuO/α-MnO2-400 had two main peaks around 654.0 eV and 642.0 eV, which could be attributed to the binding energies of the Mn 2p1/2 at BE = 654.0 eV and Mn 2p3/2 at BE = 642.0 eV, respectively. The two main peaks were divided into four peaks after the peak fitting. The fitted peaks of Mn 2p3/2 at 642.0 eV and 643.5 eV indicate the existence of Mn3+ and Mn4+ in the 10CuO-3MOx/α-MnO2-400 catalyst [40,41,42,43,44]. The ratios of the Mn3+/(Mn3+ + Mn4+) over different catalysts followed the below order: 10CuO-3Co3O4/α-MnO2 (0.606) > 10CuO-3NiO/α-MnO2 (0.541) > 10CuO/α-MnO2 (0.414) > 10CuO-3Fe2O3/α-MnO2 (0.406). The redox electron pair in Cu-Mn oxide was the -Cu2+-O2−-Mn4+-→-Cu+-□-Mn3+- + 1/2O2 (□ indicates the oxygen vacancy) [45]. The content of Mn3+ and oxygen vacancies are proportional, or indirectly proportional, to the oxidation capacity of the catalyst [25,29,46,47]. The Mn3+ may cause the Jahn-Teller effect, which could prolong the Mn-O bond in [MnO6] [48,49], thereby prolonging the distance between the oxygen pairs and causing the stretching of the Mn-O bond length [48]. As a result, the Mn-O bond was easier to break, and the mobility of oxygen became higher. Therefore, the released surface oxygen atoms are more likely to participate in the reaction and thus promote the catalytic performance. The XPS of O 1s of 10CuO-3MOx/α-MnO2-400 was measured to elucidate the nature of the oxygen species over the 10CuO-3MOx/α-MnO2-400 catalysts. As shown in Figure 4b, all the samples show double peaks of different oxygen species. Specifically, the BEs at around 529.8 eV and 531.4 eV could be ascribable to the surface lattice oxygen (Olatt) and surface adsorbed oxygen (Oads) species [50,51], respectively. The ratios of the Oads/Olatt are also summarized in Table 2. It can be observed that the binding energy of surface lattice oxygen (Olatt) shifted to a higher binding energy with the addition of transition metals. The surface oxidation states of the Cu species were also investigated to show the redox properties of the as-prepared catalysts. As shown in Figure 4c, all the catalysts displayed two main peaks of Cu 2p1/2 (953.6 eV) and Cu 2p3/2 (933.7 eV) [43]. The Cu 2p3/2, orbitals with binding energy in the range of 930.0–935.0 eV, could be divided into two peaks. Specifically, the binding energy peak at 933.4 eV was attributed to Cu+, and the peak at 934.4 eV was attributed to Cu2+ [43]. Furthermore, it is worth noting that the Cu 2p3/2 peak is accompanied by the vibrating satellite peaks in the range of 940.0–944.0 eV [52]. Combined with the results of the Mn 2p spectrum, the catalysts formed redox pairs of Cu+/Cu2+ and Mn3+/Mn4+, which would promote the charge transference to generate more oxygen defects [37,53].
To determine the valence state of the transition metals loaded on the catalyst, the XPS spectra of Fe 2p, Co 2p, and Ni 2p are determined. The XPS profile of Fe 2p over the 10CuO-3Fe2O3/α-MnO2-400 catalyst is shown in Figure 5a. The binding energies at 710.4 eV and 725.1 eV are ascribed to Fe 2p3/2 and Fe 2p1/2, respectively [54,55]. The peak of Fe 2p3/2 can be divided into two peaks (710.3 eV and 712.5 eV) [56]. In addition, a satellite peak was observed at about 718.3 eV. This indicates that the iron species existed in the form of Fe3+ on the surface of the 10CuO-3Fe2O3/α-MnO2-400 [54,56]. The XPS of Co 2p over the 10CuO-3Co3O4/α-MnO2-400 catalyst is shown in Figure 5b. The binding energies at 780.0 eV are ascribed to the Co 2p3/2 [57]. Meanwhile, there was no significant satellite shake-up intensity at 786 eV, indicating the dominance of Co3+ on the surface of Co3O4 [57]. The XPS of Ni 2p over the 10CuO-3NiO/α-MnO2-400 catalyst is shown in Figure 5c. The binding energies at 855.1eV are ascribed to the Ni 2p3/2 [58]. The peak of Ni 2p3/2 of metallic Ni was basically at 852.6 eV, and the binding energy of NiO is about 1 eV higher than that of Ni0 [58]. The peak of NiO 2p3/2 is lower than that of Ni 2p3/2 in this work (854.8 eV). Therefore, the oxidation state of the indicated Ni element is mainly in the form of Ni2+. The higher binding energy compared to pure NiO binding energy indicates that NiO did not exist in the free form. The strong interaction was formed between the Ni and the support. The results indicate that Fe2O3, Co3O4, and NiO were successfully loaded on the catalyst. The XPS electronic binding energies of the surface elements of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) are summarized in Table 3.

2.5. H2-TPR

The H2-TPR profiles of 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO), 10CuO-yCo3O4/α-MnO2-400 (y = 0, 1, 3, 5, 7, 10), and 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) are shown in Figure 6. It can be observed that both the 10CuO-3MOx/α-MnO2-400 catalyst and the 10CuO/α-MnO2-400 catalyst show similar hydrogen consumption peaks in strength and shape in Figure 6a. Specifically, there were two sets of peaks in the range of 297–342 °C and 462–472 °C, which might be attributed to the hydrogen consumption derived from the gradual reduction of the α-MnO2 nanowire (MnO2→Mn2O3→Mn3O4), according to the pioneer report [31]. Meanwhile, it is worth noting that the loading of the transition metal oxides on the 10CuO/α-MnO2 support changed the reduction behavior of the 10CuO/α-MnO2 catalyst. Specifically, the reduction of 10CuO/α-MnO2 nanowires shifted to higher temperatures after loading the transition metal oxides. In general, the reducibility of the 10CuO/α-MnO2 catalyst decreases with the addition of transition metal oxides. The H2-TPR profiles of the 10CuO-yCo3O4/α-MnO2-400 catalysts with different Co3O4 loading amounts are shown in Figure 6b. It is of great interest to find that the positions of the reduction peaks gradually shifted to the high-temperature region with the Co3O4 loading amount increasing from 1 wt% to 10 wt%. This illustrated that the reducibility of the catalysts also gradually decreases. Therefore, the reduction temperatures of the 10CuO-yCo3O4/α-MnO2-400 catalyst were significantly higher than that of the 10CuO/α-MnO2 catalyst, except for the 10CuO-7Co3O4/α-MnO2-400 catalyst. The H2-TPR profiles of the 10CuO-3MOx/α-MnO2-T catalyst with different calcination temperatures is shown in Figure 6c. It can be observed that the reduction temperature of the 10CuO-3Co3O4/α-MnO2-T catalyst gradually shifted to a higher reduction temperature with the increase of the calcination temperature from 200 °C to 600 °C. The positions of the two reduction peaks became closer. This suggests that the interaction between the CuO and the α-MnO2 nanowire support became stronger at higher temperatures.

2.6. Catalytic Performance of the CO Oxidation

2.6.1. The Effect of the Transition Metal Oxides (MOx) on the Activities

The catalytic CO oxidation activities of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts were studied to investigate the effect of the dual loading of transition metals and CuO on the catalytic activity. As shown in Figure 7a, it can be observed that the CO conversion gradually increases with the increase of the reaction temperature until it reached 100%. As can be observed, the temperature of the 90% CO conversion over the 10CuO-3Co3O4/α-MnO2-400 catalyst was 75 °C. The activity of the 10CuO-3NiO/α-MnO2-400 catalyst basically shows a similar catalytic activity to that of the pristine 10CuO/α-MnO2-400 catalyst without modification (T90 = 77 °C). However, the CO oxidation activity of the 10CuO-3Fe2O3/α-MnO2-400 catalyst (T90 = 80 °C) was even worse than that of the pristine 10CuO/α-MnO2-400 catalyst. Therefore, only the catalytic activity of the 10CuO-3Co3O4/α-MnO2-400 was significantly improved compared with the pristine 10CuO/α-MnO2-400 catalyst. The reason for this might be that the loading of Cu and Co could generate more oxygen vacancies and active sites to the α-MnO2 nanowire. From the order of catalyst activity, it can be observed that the catalytic activity of the catalyst increases with the increase in of Mn3+ content. Meanwhile, the loading of Co3O4 increased the specific surface area and pore volume of the catalyst, providing more active sites for the reaction. The ratio of Oads/Olatt was not exactly the same as the catalytic activity of the catalyst. The reasons for this phenomenon were stated in the discussion. The CO oxidation activity was significantly improved over the 10CuO-3Co3O4/α-MnO2-400 catalyst. This result was well consistent with the Mn 2p XPS analysis. The results of the normalized reaction rates of the four catalysts are shown in Figure 7b. It can be observed that the normalized reaction rates gradually increase with the increase of the reaction temperature. The order of reaction rates of the catalysts per surface area was 10CuO-3NiO/MnO2 > 10CuO/MnO2 > 10CuO-3Co3O4/MnO2 > 10CuO-3Fe2O3/MnO2. The normalized reaction rates excluded the effect of specific surface area on catalytic activity and expressed the intrinsic catalytic activity of the catalysts. The order of the Oads/Olatt ratios was consistent with the order of the normalized reaction rates.

2.6.2. The Effect of the Co3O4 Loading Amount on the Activities

The effect of Co3O4 loading on the catalytic activity of the CO oxidation was investigated, and the profiles of the CO conversion are shown in Figure 8a. It could be noticed that the Co3O4 loading amount in the range of 0 wt.%–10 wt.% had little effect on the CO oxidation activity of the 10CuO/α-MnO2 catalyst. The order of CO catalytic activity of the catalysts is 10CuO-3Co3O4/α-MnO2-400(T90 = 75 °C) > 10CuO-10Co3O4/α-MnO2-400 (T90 = 77 °C) ≈ 10CuO/α-MnO2-400 > 10CuO-1Co3O4/α-MnO2-400 (T90 = 79 °C) ≈ 10CuO-5Co3O4/α-MnO2-400 > 10CuO-7Co3O4/α-MnO2-400 (T90 = 84 °C). The 10CuO-3Co3O4/α-MnO2-400 catalyst performed the highest activity in the low temperature region. It is shown that a certain increase in the loading of Co3O4 was beneficial to the catalytic activity of the catalyst.

2.6.3. The Effect of the Calcination Temperature on the Activities

The catalytic CO oxidation of the 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) catalyst was conducted to study the influence of the calcination temperature on catalytic activity, and the CO conversion profiles are shown in Figure 8b. The order of CO catalytic activity of catalysts is 10CuO-3Co3O4/α-MnO2-400 (T90 = 75 °C) > 10CuO-3Co3O4/α-MnO2-500 (T90 = 86 °C) > 10CuO-3Co3O4/α-MnO2-200 (T90 = 89 °C) > 10CuO-3Co3O4/α-MnO2-600 (T90 = 118 °C). It can be observed that catalytic activity has significantly decreased with the increase of the calcination temperature from 400 °C to 600 °C, especially over the 10CuO-3Co3O4/α-MnO2-600 catalyst. Specifically, the 10CuO-3Co3O4/α-MnO2-600 catalyst had low activity of CO oxidation in the low temperature region. The possible reason is that the formation of a copper manganese spinel in the catalyst after calcination at high temperatures led to the significant decrease in the CO adsorption and oxidation capacities of the catalysts. This result is consistent with the XRD analysis. The 10CuO-3Co3O4/α-MnO2-200 catalyst was also prepared for comparison. The results show that the catalytic activity of the 10CuO-3Co3O4/α-MnO2-200 catalyst was significantly lower than that of the 10CuO-3Co3O4/α-MnO2-400 catalyst. The possible reason is that the precursor of the Co3O4 could not be completely decomposed at 200 °C.

2.6.4. Stability Tests

The 12 h stability measurement was conducted over the 10CuO-3Co3O4/α-MnO2-400 catalyst under the specific conditions (80 °C, CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm), and the result is shown in Figure 9a. It can be observed that the initial activity of the 10CuO-3Co3O4/α-MnO2-400 catalyst can achieve 100% CO conversion in the first 2 h. Then, the CO conversion gradually decreased from 100% to 80% in the subsequent 2 h test. This indicates that the catalyst began to deactivate. However, the CO conversion could remain stable in the subsequent 8 h. This suggests that the 10CuO-3Co3O4/α-MnO2-400 catalyst was provided with relatively good stability to some degree.
The XRD pattern of the 10CuO-3Co3O4/α-MnO2-400 catalyst after the 12 h stability test is shown in Figure 9b. It can be observed that the diffraction peaks of the spent 10CuO-3Co3O4/α-MnO2-400 catalyst after the 12 h stability test were a bit narrower and sharper than the fresh catalyst before the stability test. The possible reason is that the 10CuO-3Co3O4/α-MnO2-400 catalyst underwent a bit of thermal agglomeration of the CuO active sites and the α-MnO2 nanowire support during the catalytic process due to the hot spots of the catalyst bed, which could partly reduce the stability of the catalyst.

3. Discussion

The oxidation of CO on Cu-doped MnO2 follows the Mars–van Krevelen (MvK) mechanism [35]. The reaction is divided into two parts: CO is first adsorbed on the catalyst surface and then reacts with surface-active oxygen on the catalyst surface to produce CO2, which then generates oxygen vacancies on the catalyst surface. O2 replenishes the depleted surface-active oxygen. After these two processes, the reaction completes a cycle [59]. There are redox electron pairs in the Cu-doped MnO2 catalyst: -Cu2+-O2−-Mn4+-→-Cu+-□-Mn3+- + 1/2O2 (□ indicates the oxygen vacancy) [45]. The content of Mn3+ on the MnO2 catalyst is higher, presumably with more oxygen vacancies on the MnO2 [60]. The order of the oxygen vacancy content of the catalysts is 10CuO-3Co3O4/α-MnO2 > 10CuO-3NiO/α-MnO2 > 10CuO/α-MnO2 > 10CuO-3Fe2O3/α-MnO2. The oxygen vacancy content is consistent with the Mn3+ content and catalyst activity. O2 is commonly activated near the oxygen vacancy, producing surface active oxygen species (Osur) [61]. It is well known that the higher the surface oxygen vacancy, the more easily O2 is activated to reactive oxygen species [50]. However, the XPS spectra of O 1s showed that the order of the Oads/Olatt ratios was not consistent with the oxygen vacancy content and catalyst activity. To evaluate the intrinsic activity of the catalysts, the surface area normalized reaction rates are determined. The results of the surface area normalized reaction rates show that the loading of NiO has the greatest effect on the intrinsic activity of the catalyst. The 10CuO-3NiO/α-MnO2 catalyst did not exhibit superior catalytic activity because the specific surface area of the catalyst after NiO loading was reduced, and the effect of specific surface area on the activity of the CO oxidation reaction could not be ignored. The 10CuO-3NiO/α-MnO2 catalyst has the most surface adsorbed oxygen and reaction rates per unit surface area, but the small specific surface area results in the catalytic activity being similar to that of the pristine 10CuO/α-MnO2-400 catalyst. The order of the intrinsic activity of the catalyst is consistent with the order of the Oads/Olatt ratio (10CuO-3NiO/α-MnO2 > 10CuO/α-MnO2 > 10CuO-3Co3O4/α-MnO2 > 10CuO-3Fe2O3/α-MnO2). This suggests that the surface-adsorbed oxygen is the reactive oxygen species involved in the oxidation of CO. CO2 was produced by CO reacting with surface-adsorbed oxygen species [62,63]. After loading different transition metals, the catalysts form different types of oxygen vacancies, which have different electron densities and affect the production of reactive oxygen species [39]. The oxygen vacancies of 10CuO-3Co3O4/α-MnO2 did not activate O2 as well, and the 10CuO-3Co3O4/α-MnO2 catalyst did not form more surface adsorbed oxygen. This might be the cause of the 10CuO-3Co3O4/α-MnO2 catalyst providing the highest oxygen vacancy but poor surface-adsorbed oxygen. Although the normalized reaction rate of the 10CuO-3Co3O4/α-MnO2 catalyst is not the highest, its high specific surface area allows for a greater number of oxygen vacancies. The large number of oxygen vacancies of the 10CuO-3Co3O4/α-MnO2 catalyst counteracted the low activity of the oxygen vacancies and performed the high catalytic activity of CO oxidation. Therefore, the 10CuO-3Co3O4/α-MnO2 catalyst exhibited the highest activity owing to its maximum specific surface area. The activity of the 10CuO-3Co3O4/α-MnO2 catalyst was slightly higher than that of the 10CuO-3NiO/α-MnO2 catalyst.

4. Materials and Method

4.1. Synthesis of α-MnO2 Nanowire Supports

The α-MnO2 nanowire support was facilely prepared via the typical hydrothermal method. Typically, 2 mmol KMnO4 (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China, AR, >99.5%) and 3 mmol MnSO4·H2O (Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China, AR, 99%) were absolutely dissolved in 40 mL deionized water and vigorously stirred for 5 min, respectively. Then, these two solutions were mixed together by adding the KMnO4 solution into the MnSO4 solution. After stirring for 30 min, the obtained suspension was transferred to the 100 mL Teflon-lined stainless-steel autoclave. The autoclave was kept at 160 °C for 12 h. The obtained precipitate after the hydrothermal reaction was separated by the centrifuge and washed with ethanol (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China, AR) and deionized water for three times. Then, the final powder was dried in the 100 °C oven overnight after the centrifugation. The obtained α-MnO2 was used as a support for subsequent experiments.

4.2. The Preparation of the Transition Metal Oxide (Fe2O3, Co3O4, NiO)-Doped CuO-Based Catalysts Supported on the α-MnO2 Nanowire

The transition metal oxides promoted CuO-based α-MnO2 nanowire supported catalysts were prepared by the deposition precipitation method as reported in our previous work [64]. The weight percentages of the CuO and transition metal oxides were controlled at x% and y% (x% = mCuO/(mCuO + mMOx + msupport) × 100%; y% = mMOx/(mCuO + mMOx + msupport) × 100%) by using the Cu(NO3)2·6H2O (Shanghai Xinbao Fine Chemical Industry Factory, Shanghai, China, AR, >99.5%), Fe(NO3)3·9H2O (Shanghai Macklin Bio-Chem Co., Ltd., Shanghai, China, AR, 98.5%), Co(NO3)2·6H2O (Shanghai Macklin Bio-Chem Co., Ltd., Shanghai, China, AR, 99%), Ni(NO3)2·6H2O (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China, AR, 99%) as the precursors. For the specific procedure, the α-MnO2 nanowire was firstly dispersed in a Cu(NO3)2·6H2O and Fe(NO3)3·9H2O/Co(NO3)2·6H2O/Ni(NO3)2·6H2O aqueous solution Then, the Na2CO3 (0.01M, Shanghai Ling Feng Chemical Reagent Co., Ltd., Shanghai, China, AR, >99.8%) solution was added by droplet into the above solution to adjust the pH to 8–9 under vigorously stirring. The resultant solution was stirred for 30 min and then kept still for 1 h. The solid powder was separated by filtration and washed by the deionized water. The obtained catalyst was dried at 120 °C in an oven overnight and then calcinated at 400 °C for 5 h with a heating rate of 2 °C/min. The α-MnO2 nanowire-supported catalysts with 10 wt.% CuO and 3 wt.% MOx (MOx = Fe2O3, Co3O4, NiO) were denoted as the 10CuO-3MOx/α-MnO2 (MOx = Fe2O3, Co3O4, NiO). Meanwhile, the loading amount of the Co3O4 (wt.%) was subsequently changed in the same way. The obtained catalysts with loading amounts of 10 wt.% CuO and y wt.% Co3O4 were denoted as the 10CuO-yCo3O4/α-MnO2 (y = 0, 1, 3, 5, 7, 10). As for the influence of the calcination temperature, the catalysts with loading of the 10 wt.% CuO and 3 wt.% Co3O4 were calcinated at different temperatures and named 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C).

4.3. Catalyst Characterizations

The X-ray diffraction (XRD) patterns of the catalysts were recorded on an X-ray power diffractometer (XRD-6100) from the Shimadzu Corporation by using the Cu Kα rays (0.15046 nm), 40 KV tube voltage, and 40 mA tube current. The 2θ scanning range was from 10° to 80°, and the scanning speed was controlled at 3°/min. The scanning electron microscope (SEM) images were taken on an Apreo S Hivac instrument (Thermo Fisher Science, Waltham, MA, USA) with the accelerating voltage of 5 kV. The nitrogen physisorption was performed on an Autosorb-iQ-AG-MP instrument (Quantachrome, Boynton Beach, FL, USA) at liquid nitrogen temperature (−196 °C). The samples were degassed at 300 °C for 3 h to remove the surface-adsorbed water and impurities before the regular test. The specific surface areas of the catalysts were calculated by the Brunauer-Emmett-Teller (BET) method, and the pore size distribution and pore volume were calculated from the adsorption branch of the isotherm by the Barrett-Joyner-Halenda (BJH) method in the range of 0–1.0 P/P0. The X-ray photoelectron spectroscopy (XPS) measurements were performed on a Thermo Science K-Alpha + spectrometer (Thermo Fisher Science, Waltham, MA, USA). For the XPS measurement, the penetration depth of each catalyst was about 1–2 nm. The binding energy (BE) was calibrated by using C 1s = 284.8 eV as the standard. A fixed-bed reactor was used to conduct the hydrogen temperature-programmed reduction (H2-TPR) experiment. The hydrogen consumption profile was recorded and analyzed with the online LC-D200 mass spectrometer (TILON GRP TECHNOLOGY LIMITED, Shanghai, China). The mixture of H2 (1.2 mL/min) and Ar (23.7 mL/min) was introduced into the reactor. For each test, 100 mg of catalyst was loaded. After the hydrogen signal baseline line (m/z = 2) was stabilized, the H2-TPR test was performed with a heating rate of 20 °C/min from room temperature to 800 °C.

4.4. Catalyst Evaluation

The activity of the catalysts was evaluated in a fixed-bed reactor equipped with a quartz tube (i.d. = 8.00 mm). For each test, 100 mg catalyst was placed in the center of the quartz tube. The reactant gases, with a composition of 1% CO, 21% O2, 79% N2 (20 mL/min), were introduced into the reactor with the gas hourly space velocity (GHSV) of 12,000 mL·g−1·h−1. The catalytic activities of CO oxidation over different catalysts were tested in the specified temperature range. Each catalyst was tested three times at each temperature. The outlet gases were analyzed online by using the GC-680 Perkin Elmer gas chromatograph equipped with the thermal conductivity detector (TCD). The 24 h stability tests of catalysts were carried out at 80 °C with the GHSV of 12,000 mL·g−1·h−1.
The catalytic performances of CO oxidation over these catalysts were stated in the form of the CO conversion. The calculated formula is listed below:
XCO = (CCO, Inlet − CCO, Outlet)/CCO, Inlet × 100%
where XCO is the conversion rate of CO; and CCO, Inlet (ppm), and CCO, Outlet (ppm) are CO flowing into and out of the reactor, respectively.
In order to evaluate the intrinsic rate of CO oxidation on these catalysts, the calculated formula of the specific surface area normalization reaction rate is listed below [25]:
r norm ( mol · m 2 · s 1 ) = C   inlet · F m cat · S BET · ln ( 1 1 X CO )
where rnorm (mol·m−2·s−1) is the normalized reaction rate, F (mol·s−1) is the CO flow rate, mcat (g) is the mass of catalyst, and SBET (m2·g−1) is the BET surface area.

5. Conclusions

In this work, the novel α-MnO2 nanowire was prepared by the one-step hydrothermal method. A series of transition metal oxides (Fe2O3, Co3O4, NiO) promoted the CuO-based catalyst supported on the α-MnO2 nanowire and were synthesized by the co-precipitation method for the CO oxidation reaction. The effects of the transition metal oxide type, the loading amount, and the calcination temperature on the CO oxidation reaction had been systematically investigated. It was found that the 10CuO-3Co3O4/α-MnO2-400 catalyst showed the highest reactivity with T90 = 75 °C. It was found that the 10CuO-3Co3O4/α-MnO2-400 catalyst possessed the largest specific surface area and exposed more active sites, which could further enhance the catalytic activity. The 10CuO-3NiO/α-MnO2-400 catalyst had the highest surface-adsorbed oxygen content and normalized reaction rate. This indicated that the surface-adsorbed oxygen was the surface-active oxygen involved in the oxidation of CO. It was found that the 10CuO-3Co3O4/α-MnO2-400 catalyst suffered from some deactivation during the 12 h stability test, which might be caused by the thermal sintering and agglomeration of the CuO active sites and α-MnO2 nanowire support. This should be the key consideration when designing the CuO-based CO oxidation catalyst in the future.

Author Contributions

Investigation, H.Z., H.S., Y.Z., Y.C., C.-e.W. and J.Q.; formal analysis, H.Z., H.S., Y.Z., Y.C., Y.X., C.-e.W. and J.X.; conceptualization, H.Z., Y.Z., H.S. and J.Q.; writing—original draft preparation, Y.Z. and L.X.; writing—review and editing, C.P., L.X. and M.C.; supervision, C.P., L.X. and M.C.; funding acquisition, L.X. and M.C.; project administration, L.X. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 22276098, 21976094, and 22176100), the National Key Research and Development Project (Grant No.2018YFC0213802), the Jiangsu Province “Carbon Peak and Carbon Neutrality Science and Technology Innovation Special Fund (The Third Batch)—Industry Foresight and Key Core Technology Research (Grant No. BE2022033-2), and the Postgraduate Research and Practice Innovation Program of Jiangsu Province (SJCX22_0367 and KYCX22_1216).

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schnitzhofer, R.; Beauchamp, J.; Dunkl, J.; Wisthaler, A.; Weber, A.; Hansel, A. Long-term measurements of CO, NO, NO2, benzene, toluene and PM10 at a motorway location in an Austrian valley. Atmos. Environ. 2008, 42, 1012–1024. [Google Scholar] [CrossRef]
  2. Kunkalekar, R.K.; Salker, A.V. Low temperature carbon monoxide oxidation over nanosized silver doped manganese dioxide catalysts. Catal. Commun. 2010, 12, 193–196. [Google Scholar] [CrossRef]
  3. Zheng, B.; de Beurs, K.M.; Owsley, B.C.; Henebry, G.M. Scaling relationship between CO pollution and population size over major US metropolitan statistical areas. Landsc. Urban Plan. 2019, 187, 191–198. [Google Scholar] [CrossRef]
  4. Wu, X.; Lang, J.; Sun, Z.; Jin, F.; Hu, Y.H. Photocatalytic conversion of carbon monoxide: From pollutant removal to fuel production. Appl. Catal. B Environ. 2021, 295, 120312. [Google Scholar] [CrossRef]
  5. Yoo, S.; Lee, E.; Jang, G.H.; Kim, D.H. Effect of Pd precursors on the catalytic properties of Pd/CeO2 catalysts for CH4 and CO oxidation. Mol. Catal. 2022, 533, 112791. [Google Scholar] [CrossRef]
  6. Schubert, M.M.; Hackenberg, S.; van Veen, A.C.; Muhler, M.; Plzak, V.; Behm, R.J. CO Oxidation over Supported Gold Catalysts—“Inert” and “Active” Support Materials and Their Role for the Oxygen Supply during Reaction. J. Catal. 2001, 197, 113–122. [Google Scholar] [CrossRef]
  7. Min, B.K.; Friend, C.M. Heterogeneous Gold-Based Catalysis for Green Chemistry: Low-Temperature CO Oxidation and Propene Oxidation. ChemInform 2007, 107, 2709–2724. [Google Scholar] [CrossRef]
  8. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, G.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, G.; Zhao, Y.; Fu, G.; Duchesne, P.; Gu, L.; Zheng, Y.; Weng, X.; Chen, M.; Pao, C. Interfacial effects in iron-nickel hydroxide-platinum nanoparticles enhance catalytic oxidation. Science 2014, 344, 495–499. [Google Scholar] [CrossRef] [Green Version]
  10. Liu, K.; Wang, A.; Zhang, T. Recent Advances in Preferential Oxidation of CO Reaction over Platinum Group Metal Catalysts. ACS Catal. 2012, 2, 1165–1178. [Google Scholar] [CrossRef]
  11. Zhang, L.; Liu, H.; Huang, X.; Sun, X.; Jiang, Z.; Schlögl, R.; Su, D. Stabilization of Palladium Nanoparticles on Nanodiamond-Graphene Core-Shell Supports for CO Oxidation. Angew. Chem. 2015, 127, 16049–16052. [Google Scholar] [CrossRef]
  12. Gonzalez-A, E.; Rangel, R.; Solís-Garcia, A.; Venezia, A.M.; Zepeda, T.A. FTIR investigation under reaction conditions during CO oxidation over Ru(x)-CeO2 catalysts. Mol. Catal. 2020, 493, 111086. [Google Scholar] [CrossRef]
  13. Yadav, P.K.; Kumari, S.; Naveena, U.; Deshpande, P.A.; Sharma, S. Insights into the substitutional chemistry of La1-xSrxCo1-yMyO3 (M = Pd, Ru, Rh, and Pt) probed by in situ DRIFTS and DFT analysis of CO oxidation. Appl. Catal. A Gen. 2022, 643, 118768. [Google Scholar] [CrossRef]
  14. Bunluesin, T.; Cordatos, H.; Gorte, R.J. Study of CO Oxidation Kinetics on Rh/Ceria. J. Catal. 1995, 157, 222–226. [Google Scholar] [CrossRef]
  15. Dey, S.; Dhal, G.C. Applications of silver nanocatalysts for low-temperature oxidation of carbon monoxide. Inorg. Chem. Commun. 2019, 110, 107614. [Google Scholar] [CrossRef]
  16. Du, Y.; Meng, Q.; Wang, J.; Yan, J.; Fan, H.; Liu, Y.; Dai, H. Three-dimensional mesoporous manganese oxides and cobalt oxides: High-efficiency catalysts for the removal of toluene and carbon monoxide. Micropor. Mesopor. Mat. 2012, 162, 199–206. [Google Scholar] [CrossRef]
  17. Xiao, M.; Zhang, X.; Yang, Y.; Cui, X.; Chen, T.; Wang, Y. M (M = Mn, Co, Cu)-CeO2 catalysts to enhance their CO catalytic oxidation at a low temperature: Synergistic effects of the interaction between Ce3+-Mx+-Ce4+and the oxygen vacancy defects. Fuel 2022, 323, 124379. [Google Scholar] [CrossRef]
  18. Bi, F.; Zhang, X.; Xiang, S.; Wang, Y. Effect of Pd loading on ZrO2 support resulting from pyrolysis of UiO-66: Application to CO oxidation. J. Colloid Interf. Sci. 2020, 573, 11–20. [Google Scholar] [CrossRef] [PubMed]
  19. Dey, S.; Dhal, G.C. Catalytic Conversion of Carbon Monoxide into Carbon Dioxide over Spinel Catalysts: An Overview. Mater. Sci. Energy Technol. 2019, 2, 575–588. [Google Scholar] [CrossRef]
  20. Kaplin, I.Y.; Lokteva, E.S.; Golubina, E.V.; Shishova, V.V.; Maslakov, K.I.; Fionov, A.V.; Isaikina, O.Y.; Lunin, V.V. Efficiency of manganese modified CTAB-templated ceria-zirconia catalysts in total CO oxidation. Appl. Surf. Sci. 2019, 485, 432–440. [Google Scholar] [CrossRef]
  21. Niu, X.; Lei, Z. Copper doped manganese oxides to produce enhanced catalytic performance for CO oxidation. J. Environ. Chem. Engi. 2019, 7, 103055. [Google Scholar] [CrossRef]
  22. Liu, H.; Li, X.; Dai, Q.; Zhao, H.; Chai, G.; Guo, Y.; Guo, Y.; Wang, L.; Zhan, W. Catalytic oxidation of chlorinated volatile organic compounds over Mn-Ti composite oxides catalysts: Elucidating the influence of surface acidity. Appl. Catal. B Environ. 2021, 282, 119577. [Google Scholar] [CrossRef]
  23. Ye, Z.; Giraudon, J.-M.; Nuns, N.; Simon, P.; De Geyter, N.; Morent, R.; Lamonier, J.-F. Influence of the preparation method on the activity of copper-manganese oxides for toluene total oxidation. Appl. Catal. B Environ. 2018, 223, 154–166. [Google Scholar] [CrossRef]
  24. Lin, Y.; Tian, H.; Qian, J.; Yu, M.; Hu, T.; Lassi, U.; Chen, Z.; Wu, Z. Biocarbon-directed vertical δ-MnO2 nanoflakes for boosting lithium-ion diffusion kinetics. Mater. Today Chem. 2022, 26, 101023. [Google Scholar] [CrossRef]
  25. Yang, W.; Su, Z.; Xu, Z.; Yang, W.; Peng, Y.; Li, J. Comparative study of α-, β-, γ- and δ-MnO2 on toluene oxidation: Oxygen vacancies and reaction intermediates. Appl. Catal. B Environ. 2019, 260, 118150. [Google Scholar] [CrossRef]
  26. Hayashi, E.; Yamaguchi, Y.; Kamata, K.; Tsunoda, N.; Kumagai, Y.; Oba, F.; Hara, M. Effect of MnO2 Crystal Structure on Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid. J. Am. Chem. Soc. 2019, 141, 890–900. [Google Scholar] [CrossRef] [PubMed]
  27. Robinson, D.M.; Go, Y.B.; Mui, M.; Gardner, G.; Zhang, Z.; Mastrogiovanni, D.; Garfunkel, E.; Li, J.; Greenblatt, M.; Dismukes, G.C. Photochemical Water Oxidation by Crystalline Polymorphs of Manganese Oxides: Structural Requirements for Catalysis. J. Am. Chem. Soc. 2013, 135, 3494–3501. [Google Scholar] [CrossRef] [PubMed]
  28. Liang, S.; Teng, F.; Bulgan, G.; Zong, R.; Zhu, Y. Effect of Phase Structure of MnO2 Nanorod Catalyst on the Activity for CO Oxidation. J. Phys. Chem. C 2008, 112, 5307–5315. [Google Scholar] [CrossRef]
  29. Tian, F.; Li, H.; Zhu, M.; Tu, W.; Lin, D.; Han, Y. Effect of MnO2 Polymorphs’ Structure on Low-Temperature Catalytic Oxidation: Crystalline Controlled Oxygen Vacancy Formation. ACS Appl. Mater. Interfaces 2022, 14, 18525–18538. [Google Scholar] [CrossRef] [PubMed]
  30. Yang, R.; Fan, Y.; Ye, R.; Tang, Y.; Cao, X.; Yin, Z.; Zeng, Z. MnO2-Based Materials for Environmental Applications. Adv. Mater. 2021, 33, 2004862. [Google Scholar] [CrossRef] [PubMed]
  31. Jokar, R.; Alavi, S.M.; Rezaei, M.; Akbari, E. Catalytic performance of copper oxide supported α-MnO2 nanowires for the CO preferential oxidation in H2-rich stream. Int. J. Hydrogen Energy 2021, 46, 32503–32513. [Google Scholar] [CrossRef]
  32. Qian, K.; Qian, Z.; Hua, Q.; Jiang, Z.; Huang, W. Structure–activity relationship of CuO/MnO2 catalysts in CO oxidation. Appl. Surf. Sci. 2013, 273, 357–363. [Google Scholar] [CrossRef]
  33. Sun, R.L.; Zhang, S.R.; An, K.; Song, P.F.; Liu, Y. Cu1.5Mn1.5O4 spinel type composite oxide modified with CuO for synergistic catalysis of CO oxidation. J. Fuel Chem. Technol. 2021, 49, 799–808. [Google Scholar] [CrossRef]
  34. Sophiana, I.C.; Topandi, A.; Iskandar, F.; Devianto, H.; Nishiyama, N.; Budhi, Y.W. Catalytic oxidation of benzene at low temperature over novel combination of metal oxide based catalysts: CuO, MnO2, NiO with Ce0.75Zr0.25O2 as support. Mater. Today Chem. 2020, 17, 100305. [Google Scholar] [CrossRef]
  35. Gao, J.; Jia, C.; Zhang, L.; Wang, H.; Yang, Y.; Hung, S.-F.; Hsu, Y.Y.; Liu, B. Tuning chemical bonding of MnO2 through transition-metal doping for enhanced CO oxidation. J. Catal. 2016, 341, 82–90. [Google Scholar] [CrossRef]
  36. Ivanov, K.I.; Kolentsova, E.N.; Dimitrov, D.Y.; Petrova, P.T.; Tabakova, T.T. Alumina Supported Cu-Mn-Cr Catalysts for CO and VOCs oxidation. Int. Sch. Sci. Res. Innov. 2015, 9, 605–612. [Google Scholar]
  37. Krämer, M.; Schmidt, T.; Stöwe, K.; Maier, W.F. Structural and catalytic aspects of sol–gel derived copper manganese oxides as low-temperature CO oxidation catalyst. Appl. Catal. A Gen. 2006, 302, 257–263. [Google Scholar] [CrossRef]
  38. Huang, N.; Qu, Z.; Dong, C.; Qin, Y.; Duan, X. Superior performance of α@β-MnO2 for the toluene oxidation: Active interface and oxygen vacancy. Appl. Catal. A Gen. 2018, 560, 195–205. [Google Scholar] [CrossRef]
  39. Wang, Y.; Wu, J.; Wang, G.; Yang, D.; Ishihara, T.; Guo, L. Oxygen vacancy engineering in Fe doped akhtenskite-type MnO2 for low-temperature toluene oxidation. Appl. Catal. B Environ. 2021, 285, 119873. [Google Scholar] [CrossRef]
  40. Rosso, J.J.; Hochella, M.F. Natural Iron and Manganese Oxide Samples by XPS. Surf. Sci. Spectra 1996, 4, 253–265. [Google Scholar] [CrossRef]
  41. Li, J.; Zhu, P.; Zuo, S.; Huang, Q.; Zhou, R. Influence of Mn doping on the performance of CuO-CeO2 catalysts for selective oxidation of CO in hydrogen-rich streams. Appl. Catal. A Gen. 2010, 381, 261–266. [Google Scholar] [CrossRef]
  42. Michael, A.S. Mn2O3 by XPS. Surf. Sci. Spectra 1999, 6, 39–46. [Google Scholar] [CrossRef]
  43. Liu, W.; Wang, S.; Cui, R.; Song, Z.; Zhang, X. Enhancement of catalytic combustion of toluene over CuMnOx hollow spheres prepared by oxidation method. Micropor. Mesopor. Mat. 2021, 326, 111370. [Google Scholar] [CrossRef]
  44. Wang, Y.; Liu, K.; Wu, J.; Hu, Z.; Huang, L.; Zhou, J.; Ishihara, T.; Guo, L. Unveiling the Effects of Alkali Metal Ions Intercalated in Layered MnO2 for Formaldehyde Catalytic Oxidation. ACS Catal. 2020, 10, 10021–10031. [Google Scholar] [CrossRef]
  45. Wang, Y.; Yang, D.; Li, S.; Zhang, L.; Zheng, G.; Guo, L. Layered copper manganese oxide for the efficient catalytic CO and VOCs oxidation. Chem. Eng. J. 2019, 357, 258–268. [Google Scholar] [CrossRef]
  46. Zeng, J.; Xie, H.; Zhang, H.; Huang, M.; Liu, X.; Zhou, G.; Jiang, Y. Insight into the effects of oxygen vacancy on the toluene oxidation over α-MnO2 catalyst. Chemosphere 2022, 291, 135890. [Google Scholar] [CrossRef] [PubMed]
  47. Xu, J.; Wu, W.; Gao, E.; Zhu, J.; Yao, S.; Li, J. Revealing the role of oxygen vacancies on α-MnO2 of different morphologies in CO oxidation using operando DRIFTS-MS. Appl. Surf. Sci. 2023, 618, 156643. [Google Scholar] [CrossRef]
  48. Liu, W.; Xiang, W.; Chen, X.; Song, Z.; Gao, C.; Tsubaki, C.; Zhang, X. A novel strategy to adjust the oxygen vacancy of CuO/MnO2 catalysts toward the catalytic oxidation of toluene. Fuel 2022, 312, 122975. [Google Scholar] [CrossRef]
  49. Yang, W.; Zhu, Y.; You, F.; Yan, L.; Ma, Y.; Lu, C.; Gao, P.; Hao, Q.; Li, W. Insights into the surface-defect dependence of molecular oxygen activation over birnessite-type MnO2. Appl. Catal. B Environ. 2018, 233, 184–193. [Google Scholar] [CrossRef] [Green Version]
  50. Mo, S.; Zhang, Q.; Li, J.; Sun, Y.; Ren, Q.; Zou, S.; Zhang, Q.; Lu, J.; Fu, M.; Mo, D.; et al. Highly efficient mesoporous MnO2 catalysts for the total toluene oxidation: Oxygen-Vacancy defect engineering and involved intermediates using in situ DRIFTS. Appl. Catal. B Environ. 2020, 264, 118464. [Google Scholar] [CrossRef]
  51. Wang, J.; Li, J.; Jiang, C.; Zhou, P.; Zhang, P.; Yu, J. The effect of manganese vacancy in birnessite-type MnO2 on room-temperature oxidation of formaldehyde in air. Appl. Catal. B Environ. 2017, 204, 147–155. [Google Scholar] [CrossRef]
  52. Papavasiliou, J.; Avgouropoulos, G.; Ioannides, T. Combined steam reforming of methanol over Cu–Mn spinel oxide catalysts. J. Catal. 2007, 251, 7–20. [Google Scholar] [CrossRef]
  53. Liu, T.; Yao, Y.; Wei, L.; Shi, Z.; Han, L.; Yuan, H.; Li, B.; Dong, L.; Wang, F.; Sun, C. Preparation and Evaluation of Copper–Manganese Oxide as a High-Efficiency Catalyst for CO Oxidation and NO Reduction by CO. J. Phys. Chem. C 2017, 121, 12757–12770. [Google Scholar] [CrossRef]
  54. Yuan, H.L.; Wang, Y.Q.; Zhou, S.M.; Liu, L.S.; Chen, X.L.; Lou, S.Y.; Yuan, R.J.; Hao, Y.M.; Li, N. Low-Temperature Preparation of Superparamagnetic CoFe2O4 Microspheres with High Saturation Magnetization. Nanoscale Res. Lett. 2010, 5, 1817–1821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Wu, Y.; Guo, Q.; Liu, H.; Wei, S.; Wang, L. Effect of Fe doping on the surface properties of δ-MnO2 nanomaterials and its decomposition of formaldehyde at room temperature. J. Environ. Chem. Eng. 2022, 10, 108277. [Google Scholar] [CrossRef]
  56. Yadav, R.S.; Havlica, J.; Masilko, J.; Kalina, L.; Wasserbauer, J.; Hajdúchová, M.; Enev, V.; Kuřitka, I.; Kožáková, Z. Impact of Nd3+ in CoFe2O4 spinel ferrite nanoparticles on cation distribution, structural and magnetic properties. J. Magn. Magn. Mater. 2016, 399, 109–117. [Google Scholar] [CrossRef]
  57. Hoang, M.; Hughes, A.E.; Turney, T.W. An XPS study of Ru-promotion for Co/CeO2 Fischer-Tropsch catalyst. Appl. Surf. Sci. 1993, 72, 55–65. [Google Scholar] [CrossRef]
  58. Oh, Y.S.; Roh, H.S.; Jun, K.W.; Baek, Y.S. A highly active catalyst, Ni/Ce–ZrO2/θ-Al2O3, for on-site H2 generation by steam methane reforming: Pretreatment effect. Int. J. Hydrogen Energy 2003, 28, 1387–1392. [Google Scholar] [CrossRef]
  59. Hu, X.; Chen, J.; Li, S.; Chen, Y.; Qu, W.; Ma, Z.; Tang, X. The Promotional Effect of Copper in Catalytic Oxidation by Cu-Doped α-MnO2 Nanorods. J. Phys. Chem. C 2019, 124, 701–708. [Google Scholar] [CrossRef]
  60. Dong, C.; Wang, H.; Ren, Y.; Qu, Z. Effect of alkaline earth metal promoter on catalytic activity of MnO2 for the complete oxidation of toluene. J. Environ. Sci. 2021, 104, 102–112. [Google Scholar] [CrossRef] [PubMed]
  61. Feng, C.; Jiang, F.; Xiong, G.; Chen, C.; Wang, Z.; Pan, Y.; Fei, Z.; Lu, Y.; Li, X.; Zhang, R.; et al. Revelation of Mn4+-Osur-Mn3+ active site and combined Langmuir-Hinshelwood mechanism in propane total oxidation at low temperature over MnO2. Chem. Eng. J. 2023, 451, 138868. [Google Scholar] [CrossRef]
  62. Park, J.H.; Kang, D.C.; Park, S.J.; Shin, C.H. CO oxidation over MnO2 catalysts prepared by a simple redox method: Influence of the Mn (II) precursors. J. Ind. Eng. Chem. 2015, 25, 250–257. [Google Scholar] [CrossRef]
  63. Du, C.; Zhao, S.; Tang, X.; Yu, Q.; Gao, F.; Zhou, Y.; Liu, J.; Yi, H. Synthesis of ultra-thin 2D MnO2 nanosheets rich in oxygen vacancy defects and the catalytic oxidation of n-heptane. Appl. Surf. Sci. 2022, 606, 154846. [Google Scholar] [CrossRef]
  64. Cui, Y.; Song, H.; Shi, Y.; Ge, P.X.; Chen, M.D.; Xu, L.L. Enhancing the Low-Temperature CO Oxidation over CuO-Based α-MnO2 Nanowire Catalysts. Nanomaterials 2022, 12, 2083. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Powder XRD patterns of the (a) 10CuO-3MOx/α-MnO2-400 catalysts doped with different transition metals (MOx = Fe2O3, Co3O4, NiO); (b) 10CuO-yCo3O4/α-MnO2-400 catalysts doped with different contents of Co3O4 (y = 1, 3, 5, 7, 10); (c) 10CuO-3Co3O4/α-MnO2-T catalysts calcined at different temperatures (T = 200, 400, 500, 600 °C).
Figure 1. Powder XRD patterns of the (a) 10CuO-3MOx/α-MnO2-400 catalysts doped with different transition metals (MOx = Fe2O3, Co3O4, NiO); (b) 10CuO-yCo3O4/α-MnO2-400 catalysts doped with different contents of Co3O4 (y = 1, 3, 5, 7, 10); (c) 10CuO-3Co3O4/α-MnO2-T catalysts calcined at different temperatures (T = 200, 400, 500, 600 °C).
Catalysts 13 00588 g001
Figure 2. SEM-EDS images of the (a) α-MnO2 nanowire, (b) 10CuO/α-MnO2-400, (c,d) 10CuO-3Fe2O3/α-MnO2-400, (e,f) 10CuO-3Co3O4/α-MnO2-400, and (g,h) 10CuO-3NiO/α-MnO2-400.
Figure 2. SEM-EDS images of the (a) α-MnO2 nanowire, (b) 10CuO/α-MnO2-400, (c,d) 10CuO-3Fe2O3/α-MnO2-400, (e,f) 10CuO-3Co3O4/α-MnO2-400, and (g,h) 10CuO-3NiO/α-MnO2-400.
Catalysts 13 00588 g002aCatalysts 13 00588 g002b
Figure 3. (a) Nitrogen adsorption-desorption isotherms and (b) BJH pore size distribution of the 10CuO/α-MnO2-400 and 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Figure 3. (a) Nitrogen adsorption-desorption isotherms and (b) BJH pore size distribution of the 10CuO/α-MnO2-400 and 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Catalysts 13 00588 g003
Figure 4. (a) Mn 2p spectra, (b) O 1s spectra, and (c) Cu 2p spectra of the 10CuO/α-MnO2, 10CuO-3Fe2O3/α-MnO2, 10CuO-3Co3O4/α-MnO2, and 10CuO-3NiO/α-MnO2 (T = 400 °C) catalysts.
Figure 4. (a) Mn 2p spectra, (b) O 1s spectra, and (c) Cu 2p spectra of the 10CuO/α-MnO2, 10CuO-3Fe2O3/α-MnO2, 10CuO-3Co3O4/α-MnO2, and 10CuO-3NiO/α-MnO2 (T = 400 °C) catalysts.
Catalysts 13 00588 g004
Figure 5. (a) Fe 2p, (b) Co 2p, and (c) Ni 2p spectra of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Figure 5. (a) Fe 2p, (b) Co 2p, and (c) Ni 2p spectra of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Catalysts 13 00588 g005
Figure 6. H2-TPR profiles of the (a) 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO), (b) 10CuO-yCo3O4/α-MnO2-400 (y = 0, 1, 3, 5, 7, 10), and (c) 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) catalysts.
Figure 6. H2-TPR profiles of the (a) 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO), (b) 10CuO-yCo3O4/α-MnO2-400 (y = 0, 1, 3, 5, 7, 10), and (c) 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) catalysts.
Catalysts 13 00588 g006
Figure 7. (a) CO conversions and (b) normalized reaction rates over the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts under the reaction conditions: CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm.
Figure 7. (a) CO conversions and (b) normalized reaction rates over the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts under the reaction conditions: CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm.
Catalysts 13 00588 g007
Figure 8. CO conversions over the (a) 10CuO-yCo3O4/α-MnO2-400 (y = 0, 1, 3, 5, 7, 10), and (b) 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) catalysts under the reaction conditions: CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm.
Figure 8. CO conversions over the (a) 10CuO-yCo3O4/α-MnO2-400 (y = 0, 1, 3, 5, 7, 10), and (b) 10CuO-3Co3O4/α-MnO2-T (T = 200, 400, 500, 600 °C) catalysts under the reaction conditions: CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm.
Catalysts 13 00588 g008
Figure 9. (a) The 12 h stability test of CO oxidation over the 10CuO-3Co3O4/α-MnO2-400 catalyst under the conditions: 80 °C, CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm; (b) The XRD patterns of the fresh and the spent 10CuO-3Co3O4/α-MnO2-400 catalysts before and after the stability test.
Figure 9. (a) The 12 h stability test of CO oxidation over the 10CuO-3Co3O4/α-MnO2-400 catalyst under the conditions: 80 °C, CO/O2/N2 = 1/20/79, GHSV = 12,000 mL·g−1·h−1, 1 atm; (b) The XRD patterns of the fresh and the spent 10CuO-3Co3O4/α-MnO2-400 catalysts before and after the stability test.
Catalysts 13 00588 g009
Table 1. Structural properties of the 10CuO/α-MnO2-400 and 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Table 1. Structural properties of the 10CuO/α-MnO2-400 and 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
CatalystSpecific Surface Area (m2/g)Pore Volume (cm3/g)Average
Pore Diameter (nm)
Isotherm Type
10CuO/MnO2720.093.1IV H3
10CuO-3Fe2O3/MnO2640.103.1IV H3
10CuO-3Co3O4/MnO2840.183.1IV H3
10CuO-3NiO/MnO2520.103.1IV H3
Table 2. The ratio of Mn3+/(Mn3+ + Mn4+), Oads/Olatt, and Cu+/(Cu2+ + Cu+) of the 10CuO-3MOx/α-MnO2(MOx = Fe2O3, Co3O4, NiO) and the 10CuO/α-MnO2 catalyst.
Table 2. The ratio of Mn3+/(Mn3+ + Mn4+), Oads/Olatt, and Cu+/(Cu2+ + Cu+) of the 10CuO-3MOx/α-MnO2(MOx = Fe2O3, Co3O4, NiO) and the 10CuO/α-MnO2 catalyst.
SampleMn3+/(Mn3+ + Mn4+)Oads/OlattCu+/(Cu2+ + Cu+)
10CuO/α-MnO20.4140.3240.551
10CuO-3Fe2O3/α-MnO20.4060.2830.644
10CuO-3Co3O4/α-MnO20.6060.3000.563
10CuO-3NiO/α-MnO20.5410.3520.688
Table 3. XPS electronic binding energies of surface elements of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
Table 3. XPS electronic binding energies of surface elements of the 10CuO-3MOx/α-MnO2-400 (MOx = Fe2O3, Co3O4, NiO) catalysts.
SampleCu 2p3/2O 1sMn 2p3/2
10CuO/α-MnO2933.6529.7642.1
10CuO-3Fe2O3/α-MnO2933.7529.7642.1
10CuO-3Co3O4/α-MnO2933.7529.8642.3
10CuO-3NiO/α-MnO2934.2529.9642.4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Zhang, Y.; Song, H.; Cui, Y.; Xue, Y.; Wu, C.-e.; Pan, C.; Xu, J.; Qiu, J.; Xu, L.; et al. Transition Metal (Fe2O3, Co3O4 and NiO)-Promoted CuO-Based α-MnO2 Nanowire Catalysts for Low-Temperature CO Oxidation. Catalysts 2023, 13, 588. https://doi.org/10.3390/catal13030588

AMA Style

Zhang H, Zhang Y, Song H, Cui Y, Xue Y, Wu C-e, Pan C, Xu J, Qiu J, Xu L, et al. Transition Metal (Fe2O3, Co3O4 and NiO)-Promoted CuO-Based α-MnO2 Nanowire Catalysts for Low-Temperature CO Oxidation. Catalysts. 2023; 13(3):588. https://doi.org/10.3390/catal13030588

Chicago/Turabian Style

Zhang, Haiou, Yixin Zhang, Huikang Song, Yan Cui, Yingying Xue, Cai-e Wu, Chao Pan, Jingxin Xu, Jian Qiu, Leilei Xu, and et al. 2023. "Transition Metal (Fe2O3, Co3O4 and NiO)-Promoted CuO-Based α-MnO2 Nanowire Catalysts for Low-Temperature CO Oxidation" Catalysts 13, no. 3: 588. https://doi.org/10.3390/catal13030588

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