Next Article in Journal
A Thermal, Mechanical, and Materials Framework for a Shape Memory Alloy Heat Engine for Thermal Management
Next Article in Special Issue
Influence of Carbonization Conditions on Structural and Surface Properties of K-Doped Mo2C Catalysts for the Synthesis of Methyl Mercaptan from CO/H2/H2S
Previous Article in Journal
Solution-Grown MAPbBr3 Single Crystals for Self-Powered Detection of X-rays with High Energies above One Megaelectron Volt
Previous Article in Special Issue
Oxidation of Methanol and Dichloromethane on TiO2-CeO2-CuO, TiO2-CeO2 and TiO2-CuO@VUKOPOR®A Ceramic Foams
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Ce-MnOx Composite Oxides via Coprecipitation and Their Catalytic Performance for CO Oxidation

1
College of Mechanical Engineering, Wuhan Polytechnic University, Wuhan 430048, China
2
Zhongye Changtian International Engineering Co., Ltd., Changsha 410205, China
3
State Key Laboratory of Material Processing and Die & Mould Technology, Wuhan 430074, China
4
School of Materials Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2158; https://doi.org/10.3390/nano13152158
Submission received: 16 June 2023 / Revised: 17 July 2023 / Accepted: 20 July 2023 / Published: 25 July 2023
(This article belongs to the Special Issue Nanocatalysts for Air Purification)

Abstract

:
Ce-MnOx composite oxide catalysts with different proportions were prepared using the coprecipitation method, and the CO-removal ability of the catalysts with the tested temperature range of 60–140 °C was investigated systematically. The effect of Ce and Mn ratios on the catalytic oxidation performance of CO was investigated using X-ray diffraction (XRD), X-ray energy dispersive spectroscopy (EDS), scanning electron microscopy (SEM), H2 temperature programmed reduction (H2-TPR), CO-temperature programmed desorption (CO-TPD), and in situ infrared spectra. The experimental results reveal that under the same test conditions, the CO conversion rate of pure Mn3O4 reaches 95.4% at 170 °C. Additionally, at 140 °C, the Ce-MnOx series composite oxide catalyst converts CO at a rate of over 96%, outperforming single-phase Mn3O4 in terms of catalytic performance. With the decrement in Ce content, the performance of Ce-MnOx series composite oxide catalysts first increase and then decrease. The Ce MnOx catalyst behaves best when Ce:Mn = 1:1, with a CO conversion rate of 99.96% at 140 °C and 91.98% at 100 °C.

1. Introduction

A fresh focus on CO emissions has evolved as the issue of air pollution garners increasing attention in many countries, and strict emission standards have also been proposed in plans. In East Asia, the main source of CO is industrial sources. However, the CO concentration in East Asia decreased with an annual trend of 0.41 ± 0.09% between 2005 and 2016, with China alone accounting for 84% of the CO reduction [1]. In Pakistan, carbon monoxide mainly comes from industrial production and transportation, with car emissions, fossil fuels, and biomass combustion being the main sources of CO, and pollutant levels in congested areas are higher than in open areas [2]. Catalytic oxidation [3], in which CO is oxidized to CO2 by a catalyst under certain conditions, is one of the most effective ways to remove CO in the face of these extremely stringent criteria and considerable emissions.
At the moment, the catalyst system used for CO oxidation can be categorized as noble metal catalysts (such as Pt, Au, and Pd) [4,5,6], as well as non-noble metal catalysts represented by metal oxides and composite metal oxide catalysts [7,8,9]. Non-noble metal catalysts have drawn a lot of interest in comparison to noble metal catalysts because of their low cost and abundant raw material sources. Manganese is frequently utilized in the field of catalysis because it possesses a variety of oxides in different valence states and a good redox capability [10,11,12]. Mobini et al. [13] synthesized and studied the effect of Mn supported on different metal oxides on CO low-temperature oxidation reaction using the coprecipitation method and sol-gel method. They emphasized that 20 wt% Mn/CeO2 has high dispersity, proper reducibility, and the maximum concentration of active site. The temperature corresponding to 50% CO conversion (T50) is about 142 °C. By using the citric acid-nitrate spontaneous combustion method, Mahnaz et al. [14] prepared LaMnO3 catalysts with calcination temperatures of 600 °C and 900 °C and discovered that the CO conversion rate reached 90% at 162 °C and 228 °C, respectively. Cerium oxide exhibits remarkable catalytic performance due to its fluorite-type structure, good oxygen storage capacity, and ability to undergo reversible transformation into Ce3+ and Ce4+ [15]. Venkataswamy et al. [16] employed the hydrothermal method, coprecipitation method, and gel sol method to fabricate catalysts with a fixed Ce:Mn ratio of 0.7:0.3. The result revealed that the as-prepared Ce0.7Mn0.3O2−δ catalyst by hydrothermal method showed the best catalytic activity, the temperature required for T90 was 132 °C, while that prepared by coprecipitation method and gel-sol method were 158 °C and 216 °C, respectively. To prepare a CO catalyst with a fixed cerium manganese ratio of 1:8, Ye et al. [17] employed citrate sol-gel (C), hydrothermal (H), and hydrothermal-citrate complexation (CH) methods, with T90 at 110 °C, CH-1:8 achieving the best CO-conversion rate, followed by H-1:8 (T90 = 142 °C) and C-1:8 (T90 = 205 °C). Zhang et al. [18] introduced monodisperse Mn-doped CeO2 nanoparticles using a template-free solvothermal method and found that the CeO2 nanoparticles prepared via Mn doping had higher CO catalytic oxidation activity than the pure CeO2 prepared by the same method, and the Ce0.93Mn0.07O2 catalyst made the CO conversion rate reach 100% at 255 °C. Numerous experimental studies have demonstrated the superior performance of Ce-MnOx composite oxide in the field of CO catalysts. It can be efficiently applied to reduce catalyst costs and prevent heating tail gas to meet reaction requirements in low-temperature CO catalytic oxidation. At the same time, the subsequent denitrification reactions benefit from temperature compensation provided by the heat generated during the CO catalytic oxidation reaction, greatly reduce energy consumption, and have great research significance.
Considering the previous studies, it is possible to synthesize effective cerium manganese CO catalysts via a couple of methods, but the impact of the cerium manganese ratio on catalytic performance is disregarded. The regulation of the cerium manganese ratio could alter the structure and morphology of the catalyst, thereby improving its performance. In order to investigate the influence of ratios on catalysts, the coprecipitation method, which has the advantages of a simple preparation process, low cost, and easy control of preparation conditions, is adopted.
In the early stage, catalysts such as Ce and Mn were considered one of the effective methods for reducing NOx [19,20]. However, Ce-Mn-O system catalysts have recently demonstrated exceptional efficacy in reducing NO, CO, and other pollutants. As a result, in this paper, Ce-MnOx composite oxide catalysts with different molar ratios are prepared using the coprecipitation method, and the influence of Ce-MnOx catalysts with different molar ratios on the catalytic performance of CO is studied, which showed better conversion rate with lower temperature required. XRD, EDS, SEM, H2-TPR, CO-TPD, and in situ infrared detection are employed to characterize the physical and chemical properties of the catalyst. CO-conversion rates achieved 91.98% at 100 °C and 99.96% at 140 °C, which resulted in a considerable decrease in the required reaction temperature as compared to the prior Ce-MnOx composite oxide catalyst.

2. Experimental Method

2.1. Experimental Reagent

Source of Ce: Ce(NO3)3·6H2O (99% purity), source of Mn: MnC4H6O4·4H2O (99% purity), and precipitating agent: NH3·H2O (25% purity) were purchased from Macklin Inc., Shanghai, China. N2 (99.9% concentration), CO (2% concentration), and O2 (99.9% concentration) were purchased from Changsha Fanggang Gas Co., Ltd. Changsha, China.

2.2. Preparation of CO Oxidation Catalyst

The Ce-MnOx composite oxide catalyst was synthesized via coprecipitation method. A certain amount of cerium nitrate hexahydrate and manganese acetate tetrahydrate were dissolved in deionized water using a mass ratio of Ce:Mn = X:1 (where X = 3, 2, 1, 1/2, 1/3). After 1 h of magnetic stirring, ammonia was slowly added drop by drop at a constant rate of 10 s/drop until the solution pH reached 10. After 2 h of continuous stirring, the solution was allowed to stand for another 10 h until sediment occurred. The sediment was subjected to several washes with suction filtration and deionized water until the pH reached 7, and the precursor was obtained. The precursor was then dried in a blast oven for 5 h at 120 °C and calcined for 4 h at 400 °C in a muffle furnace with a heating rate of 5 °C/min. After that, the prepared catalyst was then cooled naturally to room temperature. Finally, the sample was sieved with 40–60 mesh, and the screened particles were selected for catalyst activity test.

2.3. Catalyst Activity Test

In order to test the catalyst activity, a micro-fixed bed reactor was utilized. Ce-MnOx composite oxide catalyst particles were measured using a 5 mL cylinder, and a gas space velocity of 30,000 h−1 was set. N2 serves as the equilibrium gas in the mixed gas, which also contains an O2 concentration of 16% and CO concentration of 6000 ppm. The DX4000 flue gas analyzer manufactured by Gasmet Technologies Oy, Vantaa, Finland was utilized to analyze CO concentration before and after the reaction in real time. The performance of catalyst was examined between 60 °C and 140 °C, with the test temperature being adjusted step by step. The CO conversion rate of the catalyst was calculated by the following formula:
CO   conversion   rate   η = [ CO ] in [ CO ] out [ CO ] in × 100 %

2.4. Characterization of Catalysts

The phase analysis was detected by the D8 ADVANCE, Bruker, Karlsruhe, German using a Cu-Kα source with a scan rate of 5°/min and a step size of 0.02° and a 2θ angle ranging from 5 to 80°. The structure and element distribution of the sample were examined using Talos F200i, Thermo Fisher Scientific Inc, Massachusetts, America. The specific surface area was determined using Quadasorb SI-3, Anton Paar QuantaTec Inc., Boynton Beach, FL, USA. The surface microstructure was analyzed with JSM-6490LV, JEOL, Tokyo, Japan. Catalytic reduction capacity and surface CO adsorption–desorption were assessed using the Autochem II 2920, Micromeritics Instruments Corporation, Norcross, GA, USA. Detailed CO-TPD procedures were as follows: weigh and load 0.1 g sample into a U-shaped tube filled with quartz cotton. Raise the temperature at a rate of 10 °C/min to 180 °C, and purify the catalyst with constant temperature pre-treatment in He atmosphere for 60 min to remove impurities such as H2O and CO2. Afterward, after cooling to 50 °C, switch to a 9.84% CO/He mixture gas, and maintain blowing for 60 min to achieve adsorption saturation on the catalyst surface. Then, switch to He gas and maintain for 60 min to remove residual CO in the tube. Finally, continuously introduce He gas and heat up to 900 °C at a heating rate of 10 °C/min under the He atmosphere for testing. CO-DRIFTS was tested on the Nicolet 20, Thermo Fisher Scientific Inc., Waltham, MA, USA.

3. Results and Discussion

3.1. Catalyst Performance Evaluation

Figure 1 presents the results of the performance evaluation of Ce-MnOx composite oxide catalysts with varying proportions. As can be seen, the CO conversion efficiency increases with increasing catalyst temperature, while the catalyst performance first increases and then decreases with decreasing Ce content. The catalytic efficiency of a single MnOx catalyst is relatively poor, with a CO conversion efficiency of only 46.57% at 140 °C and 95.4% at 170 °C. However, the introduction of Ce significantly enhances the catalytic performance, drastically reducing the catalytic temperature. All of the Ce-MnOx catalysts achieve CO conversion rates above 96% at 140 °C, and the Ce-MnOx catalyst with the greatest catalytic activity has a Ce:Mn ratio of 1:1. At a temperature of 100 °C, the CO conversion rate reaches 91.98%, while the conversion rate of CO of single MnOx catalyst is only 7.66%. Furthermore, the Ce-MnOx catalyst demonstrates an exceptional CO conversion rate of 99.96% at a higher temperature of 140 °C. Further elaboration on the catalytic mechanism will be discussed in detail later.

3.2. XRD and EDS Characterization Results for Catalysts

The phase analysis of Ce-MnOx composite oxide catalysts in various ratios is illustrated in Figure 2. In the Ce-MnOx catalyst phase, only CeO2 can be observed when the Ce:Mn ratio steadily falls from Ce:Mn = 3:1 to Ce:Mn = 1:1, as shown in the figure, leaving the oxide peak of the Mn element undetected. As the Ce content continues to decrease, the Mn3O4 peak gradually increases, and the CeO2 peak gradually decreases. When Ce:Mn = 1:2 and Ce:Mn = 1:3, the main phase of the Ce-MnOx catalyst is Mn3O4, and the secondary phase is CeO2. A clear peak shift phenomenon appears when the phase diagram at 18° is slightly enlarged, indicating that MnOx is mainly highly dispersed and amorphous when the content of Ce is high and has entered the interior of the CeO2 lattice [21,22]. Due to the existence of defects and unstable states, amorphous MnOx has a high surface energy that is advantageous for the catalyst reaction since it promotes gas adsorption and reaction [23,24]. As a result, when Ce:Mn = 1:1, the Ce-MnOx catalyst contained a sizable amount of amorphous MnOx that facilitated the quick adsorption and desorption of CO on the catalytic surface, hence increasing CO conversion. The average grain size of test samples was calculated based on the XRD pattern and Scherrer equation, as shown in Table 1. With an increase in Mn content, the average grain size of Mn3O4 increased from 17.1 nm to 20.5 nm, indicating a rise in the crystallinity of Mn3O4 in catalysis with the decrease in Ce content. The average grain size of CeO2 shows a first decreasing and then increasing trend, and CeO2 has the smallest average particle size, measuring only 4.3 nm, when Ce:Mn is 1:1. This is due to the interaction between amorphous Mn3O4 and CeO2, which enables some Mnx+ ions to enter the CeO2 lattice interior and replace Ce4+ ions, thereby forming lattice defects, which decrease the crystallinity of the CeO2 lattice and result in a decrease in the average grain size of CeO2 [25]. The lower the crystallinity, the more uniformly the material is dispersed, which is beneficial to the improvement in catalyst activity [26].
In order to further observe the element distribution in the Ce-MnOx composite oxide catalyst, the catalyst element proportion with Ce:Mn = 1:1 was characterized by EDS, and the results are displayed in Figure 3. EDS analysis of Ce and Mn was applied to the selected region, and it was found that the atomic ratio of Ce and Mn is close to 1:1, which is in line with experimental expectations, and shows that Ce and Mn elements are evenly distributed in the catalyst. Combined with XRD, the oxide peak of Mn could not be detected in the phase, and it can be inferred that MnOx failed to form a crystal structure, which is mainly amorphous.

3.3. TEM Characterization of Catalyst

TEM characterization was conducted on three Ce-MnOx catalyst samples with different cerium to manganese molar ratios (3:1, 1:1, and 1:3). The results are presented in Figure 4, which demonstrates that lattice spacing of 0.318 nm and 0.325 nm corresponded to the (111) crystal plane of CeO2 for a molar ratio of cerium to manganese of 3:1. Cerium–manganese catalysts with a molar ratio of 1:1 have lattice spacing of 0.280 nm and 0.278 nm. In the report of Minakshi et al. [27], because the mean atomic number of CeO2 was higher and the defective crystal structure was lessened, they showed stronger diffraction and Kikuchi line contrast than MnO2. However, at this ratio, the amorphous Mnx+ ion in the Ce-MnOx composite oxide catalyst replaces the Ce4+ ion, making the CeO2 lattice distorted, resulting in the (111) lattice spacing of CeO2 becoming smaller. For a catalyst with a cerium–manganese molar ratio of 1:3, the lattice spacing is 0.274 nm and 0.433 nm. The former corresponds to the (111) crystal plane of the concave CeO2, and the latter corresponds to the (112) crystal plane of Mn3O4. The results of crystal plane analysis are consistent with those of phase analysis results provided in Section 3.2. It is observed that Mn oxide exists in an amorphous state when Ce content is high.

3.4. Analysis of the Microscopic Surface Structure of Catalyst

The surface morphology and structure of the Ce-MnOx catalyst were analyzed via SEM, as shown in Figure 5. The microstructure of the Ce-MnOx catalyst changes with the decrease of Ce content. Particles build up to form the catalyst at a Ce:Mn ratio of 3:1, resulting in a rough surface with varying pore sizes. When the Ce content is reduced to Ce:Mn = 1:1, the catalyst surface is smooth, and it can be clearly seen that the pore size is similar and evenly distributed. Ce content eventually drops to Ce:Mn = 1:3, at which point the catalyst surface is rough and has a large number of flocculent structures but comparatively few pores and smaller pore sizes. Uniform pores are visible in the morphology of the catalyst with Ce:Mn = 1:1, which results in well-scattered active sites that facilitate gas adsorption and desorption. As a result, the catalyst at a Ce:Mn of 1:1 displays the best performance.

3.5. Characterization of Catalyst H2 Temperature Programmed Reduction (H2-TPR)

The redox performance of the Ce-MnOx catalyst was evaluated by H2-TPR, and the results are shown in Figure 6. With the exception of Ce:Mn = 1:3, which exhibits three reduction peaks, only two reduction peaks are observed in other component samples for Ce-MnOx catalysts below 600 °C. The generation of highly dispersed MnOx clusters by a limited number of readily reduced manganese species results in a tiny peak at 208 °C when Ce:Mn = 1:3 [28,29]. The reduction of MnO2 and Mn2O3 to Mn3O4 results in the first reduction peak, which is located at 220 °C; the second reduction peak, which is located at 350 °C, is caused by the reduction of Mn3O4 to MnO and the reduction of oxygen on the surface of CeO2 [30]. As the content of the Ce element decreases from Ce:Mn = 3:1 to Ce:Mn = 1:1, the main reduction peaks of the catalyst all migrate to the low-temperature zone, and the intensity of the corresponding peaks increases. These findings imply that when the cerium concentration decreases, the interaction between cerium and manganese oxides is strengthened, promoting the reduction behavior of cerium and manganese in the sample and improving the fluidity of oxygen species. After that, the content of Ce element continues to decrease, the main reduction peak of the catalyst moves to the high-temperature zone, and the corresponding peak strength increases, and the second reduction peak is more obvious, which also reveals that the two main components of the catalyst in Ce:Mn = 1:2 and Ce:Mn = 1:3 with higher Mn content are Mn3O4. These findings are also consistent with the XRD phase analysis results found in Section 3.2.

3.6. Characterization of Catalyst with CO-Temperature Programmed Desorption (CO-TPD)

Figure 7 displays the process of CO adsorption, heating, and CO2 release on the surface of Ce-MnOx composite oxide catalysts in varying proportions. The catalysts exhibit two similar analytical peaks, named α and β peaks, which means that there are two different processes for adsorbing CO and desorbing CO2. This may be related to different oxygen species on the catalyst: α peak is the analytical peak of CO and active oxygen on the catalyst surface to generate CO2; β peak is the analytical peak of CO and catalyst lattice oxygen to generate CO2 [31]. The lowest temperature of 113.7 °C obtained as Ce:Mn is 1:1. A reduction in Ce concentration leads to a low-temperature shift and is followed by a high-temperature shift for the αpeak. Meanwhile, the β peak gradually moves to the high temperature with the decrease in Ce content, and the peak area increases continuously as well. The results reveal that with the decrease in Ce content, the lattice oxygen content gradually increases. The catalytic activity of the catalyst is generally related to the α peak [32]. The lower the α peak temperature, the more conducive to the evolution of CO2, and the higher the catalytic activity. It is consistent with the optimum catalytic activity of Ce-MnOx composite oxide catalyst at Ce:Mn of 1:1 that the analytical peak temperature for CO combined with surface active oxygen to generate CO2 is the lowest at Ce:Mn = 1:1. This might be due to the abundance of amorphous MnOx exposing more active catalyst sites, increasing the quantity of active oxygen on the surface and considerably enhancing the catalytic performance of the catalyst.

3.7. In Situ Infrared Diffuse Reflection Characterization of Catalysts

According to the results from the evaluation of catalytic activity, the Ce-MnOx composite oxide catalyst with Ce:Mn ratio of 1:1, which demonstrated the best performance, was adopted. A sample weighing 0.08 g was selected for testing. The sample was processed for 10 min at 120 °C in a nitrogen-purge environment to remove the residual water. After the treatment was completed, the sample was cooled to room temperature, and a mixed gas (N2 used as equilibrium gas, with 16% O2 concentration and CO concentration of 6000 ppm) was introduced and gradually heated up from 60 °C to 120 °C. In situ infrared spectroscopy was carried out when the temperature stabilized at each testing temperature, and the results are displayed in Figure 8. The absorption peak at 2343 cm−1 in the figure gradually increases with the increase in temperature, which is attributed to CO2 [32]. With the increment in temperature, the catalytic activity of the sample increases, and a considerable amount of CO is converted to CO2, resulting in an elevation of the peak here. The absorption peaks at 2119 cm−1 and 2181 cm−1 are CO absorption peaks, which gradually decrease with the increment in temperature [32,33,34]. The absorption peaks of 2895 cm−1 and 2922 cm−1 are attributed to the stretching vibration of the C-H bond [35]. The absorption peaks from 1200 cm−1 to 1800 cm−1 are connected to the vibration of carbonate species [35,36,37,38], and peak strength increases with the rising temperature.
Due to the low temperature at which the CO reaction is catalyzed by the Ce-MnOx catalyst, the L-H mechanism (Langmuir-the Hinshelwood mechanism) should be followed, which has a low activation energy requirement. Combined with in situ infrared spectroscopy and the L-H mechanism, it can be inferred that CO and O2 in the mixed gas are adsorbed on the catalyst surface, and the adsorbed [CO] and [O] are activated into active intermediate carbonate in the active center of the catalyst and then decomposed into CO2 [39,40].

4. Conclusions

(1)
The catalytic efficiency of pure-phase MnOx catalyst is 95.4% at 170 °C. While the performance of the Ce-doped catalyst is greatly improved, the Ce-MnOx catalyst can achieve more than 96% CO conversion rate at 140 °C. As the amount of Ce element diminishes, the catalyst performance initially rises and then falls. The best catalyst performance is achieved when the ratio of Ce to Mn is 1:1, where the CO-removal rate can reach 91.98% at 100°C and 99.96% at 140 °C.
(2)
The Ce-MnOx catalyst phase contains only CeO2 for Ce:Mn ≥ 1, and MnOx exists in an amorphous form; when Ce:Mn < 1, the Mn element mainly exists in the form of Mn3O4, and the secondary phase is CeO2. For Ce:Mn = 1:1, the Ce-MnOx catalyst has numerous amorphous MnOx. When Ce:Mn = 1:1, the Ce-MnOx catalyst has a large amount of amorphous MnOx and possesses excellent characteristics, such as low crystallinity and uniform element and pore distribution. Due to its distinguished reduction performance, CO adsorption, and desorption performance, the composite catalyst exhibits a good CO-conversion rate at lower temperatures.
(3)
The Ce-MnOx catalyst follows the L-H mechanism for CO catalysis. The adsorbed [CO] reacts with the adsorbed [O] to generate an active center on the surface of the catalyst and then generates the active intermediate carbonate, which is decomposed to generate CO2 subsequently.

Author Contributions

Conceptualization, J.Y.; methodology, J.Y. and J.K.; software, J.L.; validation, L.Y., X.Y., H.T. and K.Y.; formal analysis, Y.K.; investigation, W.L.; resources, Y.F.; data curation, Z.Y.; writing—original draft preparation, J.L.; writing—review and editing, J.L. and J.Y.; supervision, J.Y.; funding acquisition, J.Y. 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 number 51704221, and the Project of Science and Technology Program of Hubei Province, grant number 2022EHB020.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zheng, B.; Chevallier, F.; Ciais, P.; Yin, Y.; Deeter, M.N.; Worden, H.M.; Wang, Y.; Zhang, Q.; He, K. Rapid decline in carbon monoxide emissions and export from East Asia between years 2005 and 2016. Environ. Res. Lett. 2018, 13, 044007. [Google Scholar] [CrossRef] [Green Version]
  2. Aslam, A.; Ibrahim, M.; Mahmood, A.; Mubashir, M.; Sipra, H.F.K.; Shahid, I.; Ramzan, S.; Latif, M.T.; Tahir, M.Y.; Show, P.L. Mitigation of particulate matters and integrated approach for carbon monoxide remediation in an urban environment. J. Environ. Chem. Eng. 2021, 9, 105546. [Google Scholar] [CrossRef]
  3. Royer, S.; Duprez, D. Catalytic Oxidation of Carbon Monoxide over Transition Metal Oxides. Chemcatchem 2011, 3, 24–65. [Google Scholar] [CrossRef]
  4. Therrien, A.J.; Hensley, A.J.R.; Marcinkowski, M.D.; Zhang, R.; Lucci, F.R.; Coughlin, B.; Schilling, A.C.; McEwen, J.-S.; Sykes, E.C.H. An atomic-scale view of single-site Pt catalysis for low-temperature CO oxidation. Nat. Catal. 2018, 1, 192–198. [Google Scholar] [CrossRef]
  5. Xu, G.; Wang, R.; Yang, F.; Ma, D.; Yang, Z.; Lu, Z. CO oxidation on single Pd atom embedded defect-graphene via a new termolecular Eley-Rideal mechanism. Carbon 2017, 118, 35–42. [Google Scholar] [CrossRef]
  6. Zhou, X.; Shen, Q.; Yuan, K.; Yang, W.; Chen, Q.; Geng, Z.; Zhang, J.; Shao, X.; Chen, W.; Xu, G.; et al. Unraveling Charge State of Supported Au Single-Atoms during CO Oxidation. J. Am. Chem. Soc. 2018, 140, 554–557. [Google Scholar] [CrossRef] [PubMed]
  7. Sd, A.; Nsm, B. Selection of Manganese oxide catalysts for catalytic oxidation of Carbon monoxide at ambient conditions. Resour. Environ. Sustain. 2020, 1, 100003. [Google Scholar] [CrossRef]
  8. Dey, S.; Dhal, G.C.; Prasad, R.; Mohan, D. Effect of nitrate metal (Ce, Cu, Mn and Co) precursors for the total oxidation of carbon monoxide. Resour.-Effic. Technol. 2017, 3, 293–302. [Google Scholar] [CrossRef]
  9. Lykaki, M.; Pachatouridou, E.; Carabineiro, S.A.C.; Iliopoulou, E.; Andriopoulou, C.; Kallithrakas-Kontos, N.; Boghosian, S.; Konsolakis, M. Ceria nanoparticles shape effects on the structural defects and surface chemistry: Implications in CO oxidation by Cu/CeO2 catalysts. Appl. Catal. B Environ. 2018, 230, 18–28. [Google Scholar] [CrossRef]
  10. Sun, L.; Cao, Q.; Hu, B.; Li, J.; Hao, J.; Jing, G.; Tang, X. Synthesis, characterization and catalytic activities of vanadium–cryptomelane manganese oxides in low-temperature NO reduction with NH3. Appl. Catal. A Gen. 2011, 393, 323–330. [Google Scholar] [CrossRef]
  11. Yang, S.; Wang, C.; Li, J.; Yan, N.; Ma, L.; Chang, H. Low temperature selective catalytic reduction of NO with NH3 over Mn–Fe spinel: Performance, mechanism and kinetic study. Appl. Catal. B Environ. 2011, 110, 71–80. [Google Scholar] [CrossRef]
  12. Kim, Y.J.; Kwon, H.J.; Heo, I.; Nam, I.S.; Cho, B.K.; Jin, W.C.; Cha, M.S.; Yeo, G.K. Mn–Fe/ZSM5 as a low-temperature SCR catalyst to remove NOx from diesel engine exhaust. Appl. Catal. B Environ. 2012, 126, 9–21. [Google Scholar] [CrossRef]
  13. Mobini, S.; Meshkani, F.; Rezaei, M. Supported Mn catalysts and the role of different supports in the catalytic oxidation of carbon monoxide. Chem. Eng. Sci. 2019, 197, 37–51. [Google Scholar] [CrossRef]
  14. Ghiasi, M.; Malekzadeh, A. Structural Features of (Ce, La or Sr)(Mn or Co)O3 Nano-Perovskites as a Catalyst for Carbon Monoxide Oxidation. Acta Metall. Sin. (Engl. Lett.) 2014, 27, 635–641. [Google Scholar] [CrossRef]
  15. Nhiem, D.N.; Dai, L.M.; Van, N.D.; Lim, D.T. Catalytic oxidation of carbon monoxide over nanostructured CeO2–Al2O3 prepared by combustion method using polyvinyl alcohol. Ceram. Int. 2013, 39, 3381–3385. [Google Scholar] [CrossRef]
  16. Venkataswamy, P.; Jampaiah, D.; Mukherjee, D.; Aniz, C.U.; Reddy, B.M. Mn-doped Ceria Solid Solutions for CO Oxidation at Lower Temperatures. Catal. Lett. 2016, 146, 2105–2118. [Google Scholar] [CrossRef]
  17. Ye, Z.; Liu, Y.; Nikiforov, A.; Ji, J.; Zhao, B.; Wang, J. The research on CO oxidation over Ce-Mn oxides: The preparation method effects and oxidation mechanism. Chemosphere 2023, 336, 139130. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, X.Y.; Wei, J.J.; Yang, H.X.; Liu, X.F.; Liu, W.; Zhang, C.; Yang, Y.Z. One-Pot Synthesis of Mn-Doped CeO2 Nanospheres for CO Oxidation. Eur. J. Inorg. Chem. 2013, 2013, 4443–4449. [Google Scholar] [CrossRef]
  19. Chen, J.; Fu, P.; Lv, D.; Chen, Y.; Fan, M.; Wu, J.; Meshram, A.; Mu, B.; Li, X.; Xia, Q. Unusual positive effect of SO2 on Mn-Ce mixed-oxide catalyst for the SCR reaction of NOx with NH3. Chem. Eng. J. 2021, 407, 127071. [Google Scholar] [CrossRef]
  20. Jiang, L.; Liu, Q.; Ran, G.; Kong, M.; Ren, S.; Yang, J.; Li, J. V2O5-modified Mn-Ce/AC catalyst with high SO2 tolerance for low-temperature NH3-SCR of NO. Chem. Eng. J. 2019, 370, 810–821. [Google Scholar] [CrossRef]
  21. Thirupathi, B.; Smirniotis, P.G. Co-doping a metal (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) on Mn/TiO2 catalyst and its effect on the selective reduction of NO with NH3 at low-temperatures. Appl. Catal. B Environ. 2011, 110, 195–206. [Google Scholar] [CrossRef]
  22. Sang, M.L.; Park, K.H.; Hong, S.C. MnOx/CeO2–TiO2 mixed oxide catalysts for the selective catalytic reduction of NO with NH3 at low temperature. Chem. Eng. J. 2012, 195–196, 323–331. [Google Scholar] [CrossRef]
  23. Li, Q.; Li, X.; Li, W.; Zhong, L.; Zhang, C.; Fang, Q.; Chen, G. Effect of preferential exposure of anatase TiO2 {0 0 1} facets on the performance of Mn-Ce/TiO2 catalysts for low-temperature selective catalytic reduction of NOx with NH3. Chem. Eng. J. 2019, 369, 26–34. [Google Scholar] [CrossRef]
  24. Li, W.; Zhang, C.; Li, X.; Tan, P.; Zhou, A.L.; Fang, Q.Y.; Chen, G. Ho-modified Mn-Ce/TiO2 for low-temperature SCR of NOx with NH3: Evaluation and characterization. Chin. J. Catal. 2018, 39, 1653–1663. [Google Scholar] [CrossRef]
  25. Ma, X.Y.; Tian, G.L.; Cui, S.P.; Wang, Y.L. Different Precipitant Preparation of Nickel-Doped Mn/TiO2 Catalysts for Low-Temperature SCR of NO with NH3. In Materials Science Forum; Trans Tech Publications Ltd.: Bäch, Switzerland, 2018; pp. 976–984. [Google Scholar]
  26. Twigg, M.V.; Spencer, M.S. Deactivation of copper metal catalysts for methanol decomposition, methanol steam reforming and methanol synthesis. Top. Catal. 2003, 22, 191–203. [Google Scholar] [CrossRef]
  27. Minakshi, M.; Mitchell, D.R.G.; Carter, M.L.; Appadoo, D.; Nallathamby, K. Microstructural and spectroscopic investigations into the effect of CeO2 additions on the performance of a MnO2 aqueous rechargeable battery. Electrochim. Acta 2009, 54, 3244–3249. [Google Scholar] [CrossRef] [Green Version]
  28. Wu, X.; Liu, S.; Weng, D.; Lin, F.; Ran, R. MnOx-CeO2-Al2O3 mixed oxides for soot oxidation: Activity and thermal stability. J. Hazard. Mater. 2011, 187, 283–290. [Google Scholar] [CrossRef]
  29. Zhu, Y.; Wang, Q.; Lan, L.; Chen, S.H.; Zhang, J.J. Effect of surface manganese oxide species on soot catalytic combustion of Ce-Mn-O catalyst. J. Rare Earth 2022, 40, 1238–1246. [Google Scholar] [CrossRef]
  30. Li, H.F.; Lu, G.Z.; Dai, Q.G.; Wang, Y.Q.; Guo, Y.; Guo, Y.L. Efficient low-temperature catalytic combustion of trichloroethylene over flower-like mesoporous Mn-doped CeO2 microspheres. Appl. Catal. B Environ. 2011, 102, 475–483. [Google Scholar] [CrossRef]
  31. Li, J.; Lu, G.Z.; Wu, G.S.; Mao, D.S.; Guo, Y.L.; Wanga, Y.Q.; Yun, G.A. Effect of TiO2 crystal structure on the catalytic performance of Co3O4/TiO2 catalyst for low-temperature CO oxidation. Catal. Sci. Technol. 2014, 4, 1268–1275. [Google Scholar] [CrossRef]
  32. Wang, H.; Mao, D.; Qi, J.; Zhang, Q.H.; Ma, X.H.; Song, S.Y.; Gu, L.; Yu, R.B.; Wang, D. Hollow Multishelled Structure of Heterogeneous Co3O4-CeO2-x Nanocomposite for CO Catalytic Oxidation. Adv. Funct. Mater. 2019, 29, 1806588. [Google Scholar] [CrossRef]
  33. Binet, C.; Badri, A.; Lavalley, J.-C. A Spectroscopic Characterization of the Reduction of Ceria from Electronic Transitions of Intrinsic Point Defects. J. Phys. Chem. 2002, 98, 6392–6398. [Google Scholar] [CrossRef]
  34. Binet, C.; Daturi, M.; Lavalley, J.-C. IR study of polycrystalline ceria properties in oxidised and reduced states. Catal. Today 1999, 50, 207–225. [Google Scholar] [CrossRef]
  35. Dong, N.; Ye, Q.; Chen, M.; Cheng, S.; Dai, H. Sodium-treated sepiolite-supported transition metal (Cu, Fe, Ni, Mn, or Co) catalysts for HCHO oxidation. Chin. J. Catal. 2020, 41, 1734–1744. [Google Scholar] [CrossRef]
  36. Wang, S.R.; Wang, Y.J.; Jiang, J.Q.; Liu, R.; Li, M.Y.; Wang, Y.M.; Su, Y.; Zhu, B.L.; Zhang, S.M.; Huang, W.P.; et al. A DRIFTS study of low-temperature CO oxidation over Au/SnO2 catalyst prepared by co-precipitation method. Catal. Commun. 2009, 10, 640–644. [Google Scholar] [CrossRef]
  37. Gong, L.; Liu, C.; Liu, Q.; Dai, R.; Nie, X.; Lu, L.; Liu, G.; Hu, X. CuO/CeO2–MnO2 Catalyst Prepared by Redox Method for Preferential Oxidation of CO in H2-Rich Gases. Catal. Surv. Asia 2019, 23, 1–9. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Zhao, L.; Duan, J.; Bi, S. Insights into deNOx processing over Ce-modified Cu-BTC catalysts for the CO-SCR reaction at low temperature by in situ DRIFTS. Sep. Purif. Technol. 2020, 234, 116081. [Google Scholar] [CrossRef]
  39. Arena, F.; Di Chio, R.; Fazio, B.; Espro, C.; Spiccia, L.; Palella, A.; Spadaro, L. Probing the functionality of nanostructured MnCeOx catalysts in the carbon monoxide oxidation. Appl. Catal. B Environ. 2017, 210, 14–22. [Google Scholar] [CrossRef]
  40. Arena, F.; di Chio, R.; Espro, C.; Palella, A.; Spadaro, L. A definitive assessment of the CO oxidation pattern of a nanocomposite MnCeOx catalyst. React. Chem. Eng. 2018, 3, 293–300. [Google Scholar] [CrossRef]
Figure 1. Evaluation results of Ce-MnOx catalyst performance with different proportions.
Figure 1. Evaluation results of Ce-MnOx catalyst performance with different proportions.
Nanomaterials 13 02158 g001
Figure 2. XRD characterization of Ce-MnOx catalysts with different proportions.
Figure 2. XRD characterization of Ce-MnOx catalysts with different proportions.
Nanomaterials 13 02158 g002
Figure 3. The element distribution results of Ce:Mn = 1:1 catalysts for Ce, Mn.
Figure 3. The element distribution results of Ce:Mn = 1:1 catalysts for Ce, Mn.
Nanomaterials 13 02158 g003
Figure 4. TEM characterization of Ce-MnOx catalysts with different proportions: (a) Ce:Mn = 3:1; (b) Ce:Mn = 1:1; (c) Ce:Mn = 1:3.
Figure 4. TEM characterization of Ce-MnOx catalysts with different proportions: (a) Ce:Mn = 3:1; (b) Ce:Mn = 1:1; (c) Ce:Mn = 1:3.
Nanomaterials 13 02158 g004
Figure 5. SEM characterization of Ce-MnOx catalysts with different proportions: (a) Ce:Mn = 3:1; (b) Ce:Mn = 1:1; (c) Ce:Mn = 1:3.
Figure 5. SEM characterization of Ce-MnOx catalysts with different proportions: (a) Ce:Mn = 3:1; (b) Ce:Mn = 1:1; (c) Ce:Mn = 1:3.
Nanomaterials 13 02158 g005
Figure 6. H2-TPR characterization of Ce-MnOx catalysts with different proportions.
Figure 6. H2-TPR characterization of Ce-MnOx catalysts with different proportions.
Nanomaterials 13 02158 g006
Figure 7. CO-TPD characterization of Ce-MnOx catalysts with different proportions.
Figure 7. CO-TPD characterization of Ce-MnOx catalysts with different proportions.
Nanomaterials 13 02158 g007
Figure 8. CO-DRIFTS characterization of Ce-MnOx catalysts with Ce:Mn = 1:1.
Figure 8. CO-DRIFTS characterization of Ce-MnOx catalysts with Ce:Mn = 1:1.
Nanomaterials 13 02158 g008
Table 1. Average grain size of phase of double oxide catalyst.
Table 1. Average grain size of phase of double oxide catalyst.
Ce:MnThe Average Grain Size/nm
CeO2Mn3O4
3:16.50
2:14.70
1:14.30
1:25.017.1
1:36.320.5
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

Yang, J.; Li, J.; Kang, J.; Liu, W.; Kuang, Y.; Tan, H.; Yu, Z.; Yang, L.; Yang, X.; Yu, K.; et al. Preparation of Ce-MnOx Composite Oxides via Coprecipitation and Their Catalytic Performance for CO Oxidation. Nanomaterials 2023, 13, 2158. https://doi.org/10.3390/nano13152158

AMA Style

Yang J, Li J, Kang J, Liu W, Kuang Y, Tan H, Yu Z, Yang L, Yang X, Yu K, et al. Preparation of Ce-MnOx Composite Oxides via Coprecipitation and Their Catalytic Performance for CO Oxidation. Nanomaterials. 2023; 13(15):2158. https://doi.org/10.3390/nano13152158

Chicago/Turabian Style

Yang, Junsheng, Jie Li, Jiangang Kang, Wenkang Liu, Yijian Kuang, Hua Tan, Zhensen Yu, Liu Yang, Xuejin Yang, Kui Yu, and et al. 2023. "Preparation of Ce-MnOx Composite Oxides via Coprecipitation and Their Catalytic Performance for CO Oxidation" Nanomaterials 13, no. 15: 2158. https://doi.org/10.3390/nano13152158

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