Next Article in Journal
Effect of Groundwater Depression Cone on the Hydrochemical Evolution Process in the People’s Victory Canal Irrigation Area, China
Next Article in Special Issue
Recent Progress of Non-Pt Catalysts for Oxygen Reduction Reaction in Fuel Cells
Previous Article in Journal
Green Manufacturing-Oriented Polyetheretherketone Additive Manufacturing and Dry Milling Post-Processing Process Research
Previous Article in Special Issue
In Situ DRIFTS Study of Single-Atom, 2D, and 3D Pt on γ-Al2O3 Nanoflakes and Nanowires for C2H4 Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Co-Precipitated Mn0.15Ce0.85O2−δ Catalysts for NO Oxidation: Manganese Precursors and Mn-Ce Interactions

1
Key Laboratory of Advanced Materials of Ministry of Education of China, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
2
BOE Technology Group Co., Ltd., Beijing 100176, China
3
State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
4
National Engineering Laboratory for Mobile Source Emission Control Technology, China Automotive Technology & Research Center, Tianjin 300300, China
*
Author to whom correspondence should be addressed.
These two authors contributed equally to this work.
Processes 2022, 10(12), 2562; https://doi.org/10.3390/pr10122562
Submission received: 8 November 2022 / Revised: 29 November 2022 / Accepted: 1 December 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Synthesis and Application of Novel Nanocatalysts)

Abstract

:
Two Mn0.15Ce0.85O2−δ mixed oxides were synthesized by a co-precipitation method using Mn(NO3)2 and KMnO4 as the manganese precursors, respectively. Structural analyses by X-ray powder diffraction and Raman spectroscopy reveal the formation of MnOx-CeO2 solid solutions. The Mn0.15Ce0.85O2−δ catalyst prepared from the high-valent manganese precursor exhibits higher activity for the catalytic oxidation of NO. The advantage of KMnO4 is related to the improved redox property of the catalyst as supported by H2 temperature-programmed reduction (TPR) and O2 temperature-programmed desorption (TPD). The Mn-Ce interactions create more Mn4+, Ce3+ and oxygen vacancies on the KMnO4-synthesized mixed oxides based on the Raman and X-ray photoelectron spectra (XPS).

1. Introduction

NO oxidation is an important reaction for environmental protection, such as NO oxidative adsorption [1], fast selective catalytic reduction (SCR) of NO [2,3], lean NOx adsorption (LNT) [4], and NO2-assisted soot oxidation [5,6]. Pt catalysts are widely considered as the most efficient catalysts despite their high costs. Ceria is one of the cheap catalytic materials and plays important roles in many catalytic reactions owing to easy redox cycle between Ce4+ and Ce3+.
The redox property of ceria can be improved by interaction with transition metals, e.g., Fex+ [7], Mnx+ [8], Cox+ [9], and Cux+ [10]. Among the mixed oxides, MnOx-CeO2 catalysts are the most favorable ones due to a favorable synergistic effect between these two metal oxides. Nevertheless, their catalytic activities for NO oxidation have not been fully explored. Machida et al. [11] found that the enhanced NO oxidative adsorption of the mixed oxides is facilitated by both the high NO oxidation activity of MnOx and the strong NOx adsorption ability of CeO2. They pointed out that the active sites for NO oxidation were Mn species accompanied with reversible sorption/desorption of lattice oxygens based on XPS and O2-TPD studies [12]. Li et al. [13] evidenced the existence of a synergistic mechanism between the oxides in the MnOx-CeO2. They found that Mn-Ce-Ox improved the activities of NO oxidation due to the increased surface area and the enhanced dispersion of MnOx with the addition of Ce. In our previous study [14], this synergistic mechanism was further verified in a tight-contact mixture of MnO and CeO2 via a so-called long-ranged electronic interaction, which created more Mn4+ and oxygen vacancies compared with the loose-contact mixture of the oxides.
It is acceptable that such an electronic interaction would be much more significant in the solid solutions of MnOx and CeO2 synthesized by chemical methods. In this work, two Mn0.15Ce0.85O2−δ mixed oxides were synthesized by co-precipitation using different manganese precursors. Detailed microstructural and surface property characterizations were carried out to establish structure-performance relationships of the catalysts for NO oxidation.

2. Experimental

2.1. Catalyst Preparation

Two Mn0.15Ce0.85O2−δ (corresponding to a Mn/Ce molar ratio of 15/85) were synthesized by a co-precipitation method with different manganese precursors. KOH solution was added dropwise to a precursor solution of Mn(NO3)2 (aqueous solution, 50 wt.%, Aladdin, China) and Ce(NO3)3·6H2O (99.99 wt.%, Aladdin) to adjust the pH to 10.5 and stirred at 50 °C for 2 h. In order to ensure the complete removal of potassium ions, the precipitate was filtrated and washed at least four times. After drying the solid at 110 °C for 12 h and calcination at 500 °C for 6 h in air, the obtained catalyst was denoted as MnCe-L (the low-valent manganese sample). Another Mn0.15Ce0.85O2−δ catalyst was prepared by the same method except using KMnO4 (Aladdin) as the manganese precursor, and the obtained catalyst was denoted as MnCe-H (the high-valent manganese sample). Due to electronic interactions between Mnx+ and Cex+ and thermal oxidation/reduction of metal ions during calcination, different manganese and cerium ions co-existed in the final products.
For reference, Mn3O4 and CeO2 were synthesized by a similar precipitation method using Mn(NO3)2 and Ce(NO3)3·6H2O as the precursors, respectively. The physical mixture of Mn3O4 and CeO2 at a Mn/Ce molar ratio of 15/85 was prepared by mixing Mn3O4 and CeO2 powders using a spatula for two minutes. MnO2 (Aladdin) and MnO (Aladdin) were also used for reference.

2.2. Activity Measurement

Temperature-programmed oxidation (TPO) of NO was performed in a fixed-bed reactor connected to an infrared spectrometer Nicolet iS10 (Thermo Fisher, Waltham, MA, USA) with a gas mixture of 1000 ppm NO/10% O2/N2 (500 mL·min−1) as the reaction gas. In order to smooth airflow and heat dissipation, the catalyst powders (100 mg) were mixed with silica pellets (300 mg) using a spoon. Then, the mixture was placed in a tubular quartz reactor, and was heated to 700 °C at a rate of 10 °C·min−1.

2.3. Catalyst Characterization

X-ray diffraction (XRD) patterns were measured on a D8 ADVANCE diffractometer (Bruker, Germany) using Cu Kα radiation (λ = 0.15418 nm). The XRD patterns ranged between 10° and 80° at 0.02° intervals and recorded at a scanning speed of 4°·min−1.
Specific surface areas of the samples were determined on an automatic surface analyzer (F-Sorb 3400, Gold APP Instrument, Beijing, China) by the four-point Brunauer–Emmett–Teller (BET) method. Prior to the experiment, the samples were degassed at 200 °C for 2 h to remove the molecules adsorbed on the sample surface.
X-ray fluorescence analysis (XRF) was performed on a Shimadzu instrument (1800 Kyoto, Japan) to determine the elemental contents of the samples.
Raman spectra were collected on a confocal micro-Raman apparatus (Aurora J300, IDSpec, Hong Kong, China) using an Ar+ laser with a wavelength of 632.8 nm.
X-ray photoelectron spectra (XPS) were performed on a PHI-Quantera SXM system equipped with Al Kα radiation. The data was corrected based on C 1s (284.8 eV).
Temperature-programmed reduction (TPR) of H2 was conducted on an Auto Chem II Chemisorption Analyzer (Micromeritics, USA). A 50 mg sample was pretreated with a He flow (50 mL·min−1) at 450 °C for 30 min. After being cooled down to room temperature, the sample was heated in 10% H2/Ar (50 mL·min−1) to 900 °C with a rate of 10 °C·min−1.
Temperature-programmed desorption (TPD) of O2 was carried out on the Auto Chem II Chemisorption Analyzer. After purging in He at 300 °C for 1 h, 200 mg sample was exposed to 10% O2/He at 50 °C for 1 h. The reactor was heated to 900 °C at a rate of 10 °C·min−1 after flushed by He for half an hour.

3. Results and Discussion

3.1. NO Oxidation Activity

Figure 1 shows the NO oxidation activities of the mixed oxide catalysts and reference samples. The dotted NO2 profile was drawn by the thermodynamic equilibrium of NO + O2 Processes 10 02562 i001 NO2. Pure CeO2 shows low NO conversions (<30%) within the whole temperature range, which is ascribed to its relatively poor redox behavior. The NO oxidation activity of manganese oxides increases with the valance of the metal ions, i.e., MnO2 > Mn3O4 > MnO. Specifically, Mn3O4 exhibits a quite strong ability to oxidize NO to NO2 and reaches the maximum NO conversion (59%) at 337 °C. The mixture of Mn3O4 and CeO2 shows NO conversions between the two monoxide components. Given the same composition of the mixed oxides, both the co-precipitated samples present much higher catalytic performance than the mechanical mixture. MnCe-H even shows slightly higher activity than Mn3O4, while MnCe-L has somewhat lower NO conversions at the temperatures lower than 360 °C. Considering the dominating species of Ce in the mixed oxides, these facts demonstrate strong synergistic effects between manganese oxide and ceria in the co-precipitated samples, which depend importantly on the preparation methods and the precursor adopted. Although MnO2 exhibits higher NO oxidation activity than Mn0.15Ce0.85O2−δ mixed oxides, the poor thermal stability [15], poor sulfur resistance [16] and relatively high cost limit its industrial applications. It is well known that phase transformation of MnO2 to Mn2O3 occurs readily at 200–720 °C [17], while Mn0.15Ce0.85O2−δ catalysts maintain high NO oxidation activity after calcination at 650 °C [18].

3.2. Structural Properties

The powder XRD patterns of the samples are shown in Figure 2a. The diffraction peaks of the precipitated manganese oxide and ceria correlate well with the characteristics of tetragonal Mn3O4 and cubic CeO2, respectively [19]. Only diffraction peaks of ceria are observed in the patterns of two Mn0.15Ce0.85O2−δ mixed oxides. The absence of any MnOx-related peaks implies highly dispersed state of MnOx clusters in the matrix of ceria or the formation of MnOx-CeO2 solid solutions [20]. Table 1 lists the lattice constants and mean crystallite sizes of the samples calculated according to Cohen’s method and Williamson-Hall equation, respectively. Apparently, the calculated lattice constants of ceria in the mixed oxides are smaller than that of pure ceria. The shrinkage of the ceria crystal cell is attributed to the incorporation of smaller manganese ions (Mn4+: 0.053 nm; Mn3+: 0.065 nm; Mn2+: 0.083 nm) into the ceria (Ce3+: 0.114 nm; Ce4+: 0.097 nm) lattice to form metastable pseudo-solid solutions [11,21]. It herein indicates the substitution of cerium by manganese in the crystal cell of ceria in the co-precipitated mixed oxides. Furthermore, MnCe-L presents a larger lattice parameter than MnCe-H, which is associated with the incorporation of a larger amount of large Mn2+ cations in the former mixed oxides. As listed in Table 1, Mn3O4 presents an exceptionally lower BET surface area owing to significant grain growth of manganese oxide [12]. It is also noted that MnCe-H shows a higher specific surface area than MnCe-L due to the smaller crystallites. No potassium was detected by XRF, and thus its influence can be excluded. The atomic ratio of Mn/(Mn + Ce) is close to 0.15 for both the mixed oxides, correlating with the theoretical value.
Raman spectroscopy was applied to obtain additional information of both M-O band and lattice defects, and the results are shown in Figure 2b. The bands at 485 and 660 cm−1 can be attributed to the stretching mode of the Mn-O lattice of Mn3O4 [22]. CeO2 exhibits a distinct at 462 cm−1 assigned to a vibration mode of F2g symmetry. This band shifts towards lower wavenumbers to different degrees for the mixed oxides. Such red shifts are related to the creation of more oxygen vacancies, or the expansion of the crystal cell [23]. According to the XRD results, the latter possibility can be excluded. Thus, the red shifts correspond to nonstoichiometry of CeO2−δ. The greater shift degree of MnCe-H is attributed to a more significant change of CeO2 environment interacted by the substituted high-valent Mn4+ and Mn3+ cations. The ceria band is symmetrical and remains undisturbed by the manganese oxide feature at 485 cm−1. Additionally, the spectrum of the Mn3O4 + CeO2 mixture (not shown) exhibits not only the characteristics of CeO2 at 456 cm−1 but also that of Mn3O4 at 653 cm−1. Herein, the bands at 635 and 647 cm−1 are associated with the oxygen vacancies induced by the generation of Ce3+ [19,24], which show a similar shift trend with the main band. These oxygen vacancies are conducive to the diffusion of lattice oxygen, further promoting the oxidation reaction.
Figure 3 shows the deconvoluted XPS spectra of the two Mn0.15Ce0.85O2−δ mixed oxides. As shown in Figure 3a, the Mn 2p3/2 spectra reveal three main peaks at 640.5, 641.7 and 642.6 eV attributed to the presence of Mn2+, Mn3+ and Mn4+, respectively [25,26]. The peak at 646.0 eV is the shakeup satellite of divalent Mn bound to O as the charge-transfer compound [24]. The relative percentage of Mn2+ was calculated by the peak area ratio of Mn2+/(Mn2+ + Mn3+ + Mn4+). Obviously, the Mn(NO3)2-derived MnCe-L processes more Mn2+ (40.6%) than the KMnO4-dervied one (27.6%). That is, the choice of the Mn precursor affects the oxidation state of manganese ions in the obtained products.
Figure 3b shows the XPS spectra of Ce 3d fitted with eight peaks, including those of Ce4+ 3d3/2 (u, u″ and u‴), Ce4+ 3d5/2 (v, v″ and v‴), Ce3+ 3d3/2 (u′), and Ce3+ 3d5/2 (v′). The relative percentage of Ce3+ was calculated by the peak area ratio of v′/(v + v′ + v″ + v‴). The obtained surface Ce3+/Ce ratio on MnCe-H (25.7%) is somewhat higher than that on MnCe-L (23.8%). The changes are not so significant for the multiple splitting of the Ce 3d signals, which may be related to the dominant content of Ce in the mixed oxides. According to Machida’s work [11], it is plausible for Mn3+ to substitute Ce4+ in the fluorite structure when considering their structural similarity, although their ionic radius are quite different. In the present work, the introduction of more Mn3+ instead of Mn2+ in MnCe-H produces more Ce3+ due to the electronic equilibrium and structural stability since Ce3+ has a lower oxidation state and a larger ionic radius than Ce4+.
The formation of Ce3+ can also produce more oxygen vacancy and adsorbed oxygen species, which is confirmed by the XPS spectra of O 1s in Figure 3c. The peaks at 529.4 and 531.9 eV are characteristic of lattice oxygen (Olatt) and surface adsorbed oxygen (Oads), respectively [27]. The relative percentage of Oads was calculated by the peak area ratio of Oads/(Olatt + Oads). More adsorbed oxygen (33.1%) is detected on the surface of MnCe-H than on MnCe-L (22.3%), accompanied with a high ratio of Ce3+.

3.3. Redox Properties

The reducibility of metal ions was investigated by H2-TPR. Figure 4a shows the reduction profiles of Mn3O4, CeO2 and the co-precipitated mixed oxides. CeO2 has two typical peaks at 475 and 692 °C assigned to the reduction of surface/subsurface Ce4+ and bulk Ce4+, respectively. Mn3O4 shows a broad peak at 394 °C due to reduction of Mn3+ to Mn2+. The calculated H2 consumption (1720 μmol·g−1) is much smaller than the theoretical value (4370 μmol·g−1) for the reduction of Mn3O4 to MnO, which is mainly caused by the difficult reduction of large manganese oxides [28]. A shoulder at 168 °C belongs to the reduction of surface MnO2 clusters to Mn3O4 [29], although they cannot be detected by XRD.
Compared with Mn3O4, MnCe-L and especially MnCe-H show a similar but much stronger peak at 177–199 °C, which implies the generation of more Mn4+ in the mixed oxides. The calculated H2 consumption for the first deconvoluted peak in MnCe-H (624 μmol·g−1) is larger than that in MnCe-L (568 μmol·g−1). Meanwhile, the second peak shifts towards lower temperatures (276–321 °C). The calculated H2 consumption for this peak is 804 and 989 μmol·g−1 for MnCe-H and MnCe-L, respectively. These data are larger than the theoretical value (321 μmol·g−1) by assuming that all manganese ions in the mixed oxides exist in the form of Mn3O4 and MnO is the final product. Herein, it is ascribed not only to the reduction of Mn3O4 to MnO but also to that of CeO2 to Ce2O3 promoted by Mnx+ in the solid solutions or at the interface. Similarly, the promoted reduction of Ce4+ may also contribute to the first peak. Those small peaks at 361–458 °C are suggested to be associated with the reduction of less affected Mn3O4 and CeO2. The high-temperature peaks at 713–729 °C are assigned to the reduction of bulk oxygen from ceria unpromoted with Mn.
The O2-TPD technique was used to determine the reactivity of oxygen species of metal oxides. As shown in Figure 4b, the peak at 107 °C is related to desorption of superoxide ion O2 on the Mn3O4 surface [30]. The peaks at 345 and 437 °C are assigned to the desorption of chemisorbed oxygen species [31]. The peaks range from 450 to 650 °C are associated with the surface lattice oxygen species [25]. Overlapped broad peaks, which are attributed to the reduction of superoxide ions O2 and peroxide ions O22−/O on the surface, are observed for CeO2 [32]. The desorption curves of the mixed oxides are more complicated. The distinct peaks at 230–253 °C and 392–423 °C are attributed to the desorption of surface-active oxygen species bound to oxygen vacancies (i.e., O2− and O) and the coordinately unsaturated surface lattice oxygen, respectively [33]. It can be seen by integrating the peak area that more surface-active oxygen species are produced on MnCe-H than on MnCe-L. It implies that the strong metal-metal interaction in the MnCe-H results in more lattice defects (such as oxygen vacancies) and facilitates the creation of active O22−/O. The peak at around 560 °C is related to the coordinately saturated surface lattice oxygen species. The desorption peaks at 765–771 °C are attributed to lattice oxygen ion O2− from CeO2 promoted by Mnx+ [34]. It suggests that the incorporation of Mn into ceria enhances the mobility of lattice oxygen. These results are consistent with the H2-TPR results.

4. Conclusions

Mn(NO3)2 and KMnO4 were applied as the manganese precursors to synthesize Mn0.15Ce0.85O2−δ mixed oxides by coprecipitation, respectively. Manganese existed mainly in a pseudo-solid solution with a fluorite-type structure of CeO2. The application of KmnO4 leads to a higher proportion of Mn4+ and Mn3+ in the mixed oxide than that in the Mn(NO3)2-dervied one. The incorporation of Mn4+ and Mn3+ into the ceria lattice results in the formation of more Ce3+ due to the electronic equilibrium and structural stability. Herein, more oxygen vacancies and active oxygen species are created on the former catalyst, resulting in a maximum NO conversion of 64% achieved at 331 °C for NO oxidation.

Author Contributions

Experimental preparation and operation, analysed data, writing—original draft, and employment of software, Y.G. and B.J.; conceptualization, writing—review and editing, and supervision, X.W.; analysed data and funding acquisition, Z.L., R.R. and D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Engineering Laboratory for Mobile Source Emission Control Technology (No. NELMS2020A08) and the Key Laboratory of Advanced Materials of Ministry of Education of China (Grant No. ADV22-18).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, H.; Zhou, C.; Galvez, M.; Costa, P.; Chen, Y. MnOx-CeO2 mixed oxides as the catalyst for NO-assisted soot oxidation: The key role of NO adsorption/desorption on catalytic activity. Appl. Surf. Sci. 2018, 462, 678–684. [Google Scholar] [CrossRef]
  2. Zhao, S.; Peng, J.; Ge, R.; Wu, S.; Zeng, K.; Huang, H.; Yang, K.; Sun, Z. Research progress on selective catalytic reduction (SCR) catalysts for NOx removal from coal-fired flue gas. Fuel Process. Technol. 2022, 236, 107432. [Google Scholar] [CrossRef]
  3. Irfan, M.F.; Goo, J.H.; Kim, S.D. Co3O4 based catalysts for NO oxidation and NOx reduction in fast SCR process. Appl. Catal. B Environ. 2008, 78, 267–274. [Google Scholar] [CrossRef]
  4. Song, J.H.; Park, D.C.; You, Y.-W.; Kim, Y.J.; Lee, J.H.; Heo, I.; Kim, D.H. Promotive effects of Ba addition on lean NOx reduction by CO over IrRu/Al2O3 catalyst. Chem. Eng. J. 2023, 452, 139331. [Google Scholar] [CrossRef]
  5. Wasalathanthri, N.D.; SantaMaria, T.M.; Kriz, D.A.; Dissanayake, S.L.; Kuo, C.-H.; Biswas, S.; Suib, S.L. Mesoporous manganese oxides for NO2 assisted catalytic soot oxidation. Appl. Catal. B Environ. 2017, 201, 543–551. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, P.; Li, Z.; Ao, C.; Zhang, L.; Lei, L. Interactive effects of NOx synergistic and hydrothermal aging on soot catalytic combustion in Ce-based catalysts. Combust. Flame 2022, 245, 112289. [Google Scholar] [CrossRef]
  7. Tang, A.; Hu, L.; Yang, X.; Jia, Y.; Zhang, Y. Promoting effect of the addition of Ce and Fe on manganese oxide catalyst for 1,2-dichlorobenzene catalytic combustion. Catal. Commun. 2016, 82, 41–45. [Google Scholar] [CrossRef]
  8. Du, J.; Qu, Z.; Dong, C.; Song, L.; Qin, Y.; Huang, N. Low-temperature abatement of toluene over Mn-Ce oxides catalysts synthesized by a modified hydrothermal approach. Appl. Surf. Sci. 2018, 433, 1025–1035. [Google Scholar] [CrossRef]
  9. Jin, B.; Wu, X.; Weng, D.; Liu, S.; Yu, T.; Zhao, Z.; Wei, Y. Roles of cobalt and cerium species in three-dimensionally ordered macroporous CoxCe1-xOδ catalysts for the catalytic oxidation of diesel soot. J. Colloid Interface Sci. 2018, 532, 579–587. [Google Scholar] [CrossRef]
  10. Wang, Q.; Li, L.; Huang, T.; Ding, J.; Lu, Y.; Liang, B.; Liu, H.; Li, G. Sequencing the CuOx active species for CO preferential oxidation at low-temperature over CeO2-CuO composite catalysts. Chem. Eng. J. 2023, 452, 139467. [Google Scholar] [CrossRef]
  11. Machida, M.; Uto, M.; Kurogi, D.; Kijima, T. Solid-gas interaction of nitrogen oxide adsorbed on MnOx-CeO2: A DRIFTS study. J. Mater. Chem. 2001, 11, 900–904. [Google Scholar] [CrossRef]
  12. Machida, M.; Uto, M.; Kurogi, D.; Kijima, T. MnOx-CeO2 binary oxides for catalytic NOx sorption at low temperatures, sorptive removal of NOx. Chem. Mater. 2000, 12, 3158–3164. [Google Scholar] [CrossRef]
  13. Li, H.; Tang, X.; Yi, H.; Yu, L. Low-temperature catalytic oxidation of NO over Mn-Ce-Ox catalyst. J. Rare Earths 2010, 28, 64–68. [Google Scholar] [CrossRef]
  14. Wu, X.; Yu, H.; Weng, D.; Liu, S.; Fan, J. Synergistic effect between MnO and CeO2 in the physical mixture: Electronic interaction and NO oxidation activity. J. Rare Earths 2013, 31, 1141–1147. [Google Scholar] [CrossRef]
  15. Xu, Q.; Fang, Z.; Chen, Y.; Guo, Y.; Wang, L.; Wang, Y.; Zhang, J.; Wang, Z. Titania-Samarium-Manganese Composite Oxide for the Low-Temperature Selective Catalytic Reduction of NO with NH3. Environ. Sci. Technol. 2020, 54, 2530–2538. [Google Scholar] [CrossRef]
  16. Fang, X.; Liu, Y.; Cheng, Y.; Cen, W. Mechanism of Ce-Modified Birnessite-MnO2 in Promoting SO2 Poisoning Resistance for Low-Temperature NH3-SCR. ACS Catal. 2021, 11, 4125–4135. [Google Scholar] [CrossRef]
  17. Zhao, B.; Ran, R.; Wu, X.; Weng, D. Phase structures, morphologies, and NO catalytic oxidation activities of single-phase MnO2 catalysts. Appl. Catal. A Gen. 2016, 514, 24–34. [Google Scholar] [CrossRef]
  18. Wu, X.; Liu, S.; Weng, D.; Lin, F. Textural-structural properties and soot oxidation activity of MnOx-CeO2 mixed oxides. Catal. Commun. 2011, 12, 345–348. [Google Scholar] [CrossRef]
  19. Liao, Y.; Fu, M.; Chen, L.; Wu, J.; Huang, B.; Ye, D. Catalytic oxidation of toluene over nanorod-structured Mn–Ce mixed oxides. Catal. Today 2013, 216, 220–228. [Google Scholar] [CrossRef]
  20. Xiao, Y.; Wu, X.; Liu, S.; Wan, J.; Li, M.; Weng, D.; Tong, C. Modification of PdO/CeO2-ZrO2 catalyst by MnOx for water-gas shift reaction: Redox property and valence state of Pd. J. Mater. Sci. 2016, 51, 5377–5387. [Google Scholar] [CrossRef]
  21. Zhu, Y.; Wang, Q.; Lan, L.; Chen, S.; Zhang, J. Effect of surface manganese oxide species on soot catalytic combustion of Ce–Mn–O catalyst. J. Rare Earths 2022, 40, 1238–1246. [Google Scholar] [CrossRef]
  22. Silva, G.C.; Almeida, F.S.; Dantas, M.S.S.; Ferreira, A.M.; Ciminelli, V.S. Raman and IR spectroscopic investigation of as adsorbed on Mn3O4 magnetic composites. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 100, 161–165. [Google Scholar] [CrossRef] [PubMed]
  23. McBride, J.R.; Hass, K.C.; Poindexter, B.D.; Weber, W.H. Raman and X-ray studies of Ce1−xRExO2−y, where RE = La, Pr, Nd, Eu, Gd, and Tb. J. Appl. Phys. 1994, 76, 2435. [Google Scholar] [CrossRef]
  24. You, X.; Sheng, Z.; Yu, D.; Yang, L.; Xiao, X.; Wang, S. Influence of Mn/Ce ratio on the physicochemical properties and catalytic performance of graphene supported MnOx-CeO2 oxides for NH3-SCR at low temperature. Appl. Surf. Sci. 2017, 423, 845–854. [Google Scholar] [CrossRef]
  25. 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]
  26. Iwanowski, R.; Heinonen, M.; Janik, E. X-ray photoelectron spectra of zinc-blende MnTe. Chem. Phys. Lett. 2004, 387, 110–115. [Google Scholar] [CrossRef]
  27. Zhang, L.; Shu, H.; Jia, Y.; Lei, Z.; Bai, F.; Kuang, W.; Qi, L.; Shang, J.; Chao, W. Study on denitration and sulfur removal performance of Mn-Ce supported fly ash catalyst. Chemosphere 2021, 270, 128646. [Google Scholar]
  28. Ji, Y.; Duan, A.; Jiang, G.; Liu, J. Comparative study on the formation and reduction of bulk and Al2O3-supported cobalt oxides by H2-TPR technique. J. Phys. Chem. C 2009, 113, 7186–7199. [Google Scholar] [CrossRef]
  29. Lin, R.; Liu, W.-P.; Zhong, Y.-J.; Luo, M.-F. Co Oxidation Activity And Tpr Characterization Of Ag-Mn Complex Oxide Catalysts. React. Kinet. Catal. Lett. 2001, 72, 289–295. [Google Scholar] [CrossRef]
  30. Liang, H.; Hong, Y.; Zhu, C.; Li, S.; Chen, Y.; Liu, Z.; Ye, D. Influence of partial Mn-substitution on surface oxygen species of LaCoO3 catalysts. Catal. Today 2013, 201, 98–102. [Google Scholar] [CrossRef]
  31. Jampaiah, D.; Velisoju, V.; Devaiah, D.; Singh, M.; Mayes, E.; Coyle, V.; Reddy, B.; Bansal, V.; Bhargava, S. Flower-like Mn3O4/CeO2 microspheres as an efficient catalyst for diesel soot and CO oxidation: Synergistic effects for enhanced catalytic performance. Appl. Surf. Sci. 2019, 473, 209–221. [Google Scholar] [CrossRef]
  32. Lin, X.; Li, S.; He, H.; Wu, Z.; Wu, J.; Chen, L.; Ye, D.; Fu, M. Evolution of oxygen vacancies in MnOx-CeO2 mixed oxides for soot oxidation. Appl. Catal. B Environ. 2018, 223, 91–102. [Google Scholar] [CrossRef]
  33. Sun, H.; Yu, X.; Yang, X.; Ma, X.; Lin, M.; Shao, C.; Zhao, Y.; Wang, F.; Ge, M. Au/Rod-like MnO2 catalyst via thermal decomposition of manganite precursor for the catalytic oxidation of toluene. Catal. Today 2019, 332, 153–159. [Google Scholar] [CrossRef]
  34. Liang, Q.; Wu, X.; Weng, D.; Xu, H. Oxygen activation on Cu/Mn–Ce mixed oxides and the role in diesel soot oxidation. Catal. Today 2008, 139, 113–118. [Google Scholar] [CrossRef]
Figure 1. NO2 evolution during NO-TPO tests over (a) Mn0.15Ce0.85O2−δ mixed oxides and (b) reference oxide catalysts. Reaction conditions: NO = 1000 ppm, O2 = 10%, N2 in balance, GHSV = 30,000 h−1.
Figure 1. NO2 evolution during NO-TPO tests over (a) Mn0.15Ce0.85O2−δ mixed oxides and (b) reference oxide catalysts. Reaction conditions: NO = 1000 ppm, O2 = 10%, N2 in balance, GHSV = 30,000 h−1.
Processes 10 02562 g001
Figure 2. (a) XRD patterns and (b) Raman spectra of (1) MnCe-L, (2) MnCe-H, (3) CeO2 and (4) Mn3O4.
Figure 2. (a) XRD patterns and (b) Raman spectra of (1) MnCe-L, (2) MnCe-H, (3) CeO2 and (4) Mn3O4.
Processes 10 02562 g002
Figure 3. XPS spectra of (a) Mn 2p3/2, (b) Ce 3d and (c) O 1s on (1) MnCe-L and (2) MnCe-H.
Figure 3. XPS spectra of (a) Mn 2p3/2, (b) Ce 3d and (c) O 1s on (1) MnCe-L and (2) MnCe-H.
Processes 10 02562 g003
Figure 4. (a) H2-TPR and (b) O2-TPD profiles of (1) MnCe-L, (2) MnCe-H, (3) CeO2 and (4) Mn3O4.
Figure 4. (a) H2-TPR and (b) O2-TPD profiles of (1) MnCe-L, (2) MnCe-H, (3) CeO2 and (4) Mn3O4.
Processes 10 02562 g004
Table 1. Structural properties of the catalysts.
Table 1. Structural properties of the catalysts.
CatalystLattice Parameter (nm)Crystallite Size (nm)SBET (m2·g−1)
MnCe-La = b = c = 0.53876.996
MnCe-Ha = b = c = 0.53824.2126
CeO2a = b = c = 0.53946.798
Mn3O4a = b = 0.5754; c = 0.943236.323
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, Y.; Jin, B.; Wu, X.; Li, Z.; Ran, R.; Weng, D. Co-Precipitated Mn0.15Ce0.85O2−δ Catalysts for NO Oxidation: Manganese Precursors and Mn-Ce Interactions. Processes 2022, 10, 2562. https://doi.org/10.3390/pr10122562

AMA Style

Gao Y, Jin B, Wu X, Li Z, Ran R, Weng D. Co-Precipitated Mn0.15Ce0.85O2−δ Catalysts for NO Oxidation: Manganese Precursors and Mn-Ce Interactions. Processes. 2022; 10(12):2562. https://doi.org/10.3390/pr10122562

Chicago/Turabian Style

Gao, Yuxi, Baofang Jin, Xiaodong Wu, Zhenguo Li, Rui Ran, and Duan Weng. 2022. "Co-Precipitated Mn0.15Ce0.85O2−δ Catalysts for NO Oxidation: Manganese Precursors and Mn-Ce Interactions" Processes 10, no. 12: 2562. https://doi.org/10.3390/pr10122562

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