Next Article in Journal
Determining the Composite Structure of Au-Fe-Based Submicrometre Spherical Particles Fabricated by Pulsed-Laser Melting in Liquid
Previous Article in Journal
First-Principles Study of Thermo-Physical Properties of Pu-Containing Gd2Zr2O7
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Ce–Mn Composite Oxides with Enhanced Catalytic Activity for Removal of Benzene through Oxalate Method

1
Department of Chemistry, School of Chemistry and Biological Engineering, University of Science and Technology, Beijing 100083, China
2
CAS Key Laboratory of Standardization and Measurement for Nanotechnology, CAS Center for Excellence in Nanaoscience, National Center for Nanoscience and Technology, Beijing 100190, China
3
State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
*
Authors to whom correspondence should be addressed.
Those authors contributed equally to this work.
Nanomaterials 2019, 9(2), 197; https://doi.org/10.3390/nano9020197
Submission received: 10 January 2019 / Revised: 30 January 2019 / Accepted: 30 January 2019 / Published: 3 February 2019

Abstract

:
The catalytic activities of CeO2-MnOx composite oxides synthesized through oxalate method were researched. The results exhibited that the catalytic properties of CeO2-MnOx composite oxides were higher than pure CeO2 or MnOx. When the Ceat/Mnat ratio was 3:7, the catalytic activity reached the best. In addition, the activities of CeO2-MnOx synthesized through different routes over benzene oxidation were also comparative researched. The result indicated that the catalytic property of sample prepared by oxalate method was better than others, which maybe closely related with their meso-structures. Meanwhile, the effects of synergistic interaction and oxygen species in the samples on the catalytic ability can’t be ignored.

1. Introduction

Volatile organic compounds (VOCs) produced by industrial manufacturing are an important class of air pollutants. Among the VOCs, aromatic compounds are one of the major hazardous pollutants, in which benzene is considered to be one of the representing aromatic materials extensively applied in the industry. The complete catalytic oxidation of benzene is often studied as a model reaction, characteristic of the catalytic combustion of VOCs due to its chemical stability [1,2,3,4,5]. The selection of catalysts is important to catalytic degradation of benzene. At present, both classes of catalysts i.e., noble metals and transition metal oxides, have been widely studied for the degradation of VOCs [6,7,8,9]. However, the usage of noble-metal-based catalysts is limited due to high cost, low thermal stability and sensitivity to poisoning. Transition metal oxide-based catalysts are suitable alternative because of higher thermal stability and lower price [10]. In certain cases, transition metal oxides can be actually more active than noble metal catalysts [11].
Ceria (CeO2), as a typical rare earth oxide, was investigated in heterogeneous catalysis field due to its high oxygen storage capacity. It can provide active oxygen species to ensure the catalytic reaction. More recently, CeO2-based composite oxides were employed for VOCs removal and obtain satisfied results, especially Ce-Mn composites [12,13,14]. CeO2-MnOx can be applied as heterogeneous catalysts for the abatement of contaminants in the liquid and gas phases, such as the catalytic reduction of NO and oxidation of acrylic acid and formaldehyde, which exhibit much higher catalytic activity than those of pure MnOx and CeO2 [15,16,17,18].
In our previous work [19], Mn element was also doped or mixed with CeO2 to obtain Ce-Mn composites through hydrothermal method. Their catalytic behaviours over benzene oxidation were researched, among which, all of Ce-Mn composites exhibited higher activity than MnOx and CeO2. However, the best conversion temperature of benzene oxidation over Ce-Mn composite was ca. 375 °C, which could not be applied in the moderate or lower temperature range (100–200 °C) [19], even if supporting noble metal species [20]. Therefore, we need to prepare catalysts with better performance through adjusting microstructure. In this article, we report a series of Ce-Mn composites synthesized through different routes. Their catalytic activities over benzene oxidation are researched comparatively so as to acquire more active catalyst in the temperature range of 100 and 200 °C. Their microstructures are also analyzed in detail so that understanding the elements influencing the activity of samples.

2. Experimental

2.1. The Preparation of Ce-Mn Composite Oxides

The chemicals used in this work, including Ce(NO3)3·6H2O (99%), Mn(CH3COO)2·4H2O, Na2CO3, C2H2O4·2H2O, NaOH (98%), and ethanol, were purchased from Beijing Chemicals Company (Beijing, China). A series of Ce-Mn composite oxides were synthesized by carbonate method. Briefly, Ce(NO3)3·6H2O and Mn(CH3COO)2·4H2O in appreciate amounts were dissolved in a 100 mL H2O and mixed with a 0.24 M C2H2O4·2H2O solution under strong stirring. Then, the mixture was stirred for another 1 h. The precipitates were collected by centrifugation, washed with distilled water and ethanol several times. The obtained materials, labeled as CexMn1−x (where x refers to the Ce/(Ce + Mn) atomic ratio) were dried at 80 °C overnight and calcined at 450 ℃ for 2 h with a heating rate of 2 °C·min−1. Pure CeO2 and MnOx were also prepared using the similar process as reference. In order to compare the catalytic activity, Ce-Mn composite oxides were also synthesized through carbonate method and hydrothermal method [19,21].

2.2. Characterization Technique

The crystal phase of the materials was characterized on X-ray diffraction (XRD, Philips, Amsterdam, The Netherlands) equipped with a Cu Kα radiation source (λ = 0.154187 nm) at a scanning rate of 0.03 °/s (2θ from 10° to 90°). The assignment of the crystalline phases was based on the ICSD data base (CeO2 no. 81-0792; Mn3O4 no. 80-0382; Mn2O3 no. 89-4836). The morphology images of catalysts were recorded on a scanning electron microscopy (SEM, JEOL JSM-6700F, Tokyo, Japan) operating at 15 kV and 10 μA. The microstructures of catalysts were examined using transmission electron microscopy (TEM, JEOL JEM-2010F) with an accelerating voltage of 200 kV.
The BET specific surface area (SBET) was measured by physical adsorption of N2 at the liquid nitrogen temperature using an Autosorb-1 analyzer (Quantachrome, Boynton Beach, FL, USA). Before measurement, the samples were degassed at 300 °C for 4 h under vacuum. Surface composition was determined by X-ray photoelectron spectroscopy (XPS, VG Scientific, Waltham, MA, USA) using an ESCALab220i-XL electron spectrometer from VG Scientific with a monochromatic Al Kα radiation. The binding energy (BE) was referenced to the C1s line at 284.8 eV from adventitious carbon.
Hydrogen temperature-programmed reduction (H2-TPR) was performed with a U-type quartz reactor equipped with Automated Catalyst Characterization System (Autochem 2920, MICROMERITICS). A 50 mg sample (40–60 mesh) was loaded and pretreated with a 5% O2 and 95% He mixture (30 mL/min) at 150 °C for 1 h and cooled to 50 °C under He flow. The samples were then heated to 900 °C at a rate of 10 °C/min under the flow of a 10% H2 and 90% Ar mixture (30 mL/min).

2.3. Catalytic Activity Tests

Activity tests for catalytic oxidation of benzene over CexMn1−x composite catalysts were performed in a continuous-flow fixed-bed reactor under atmospheric pressure, containing 100 mg of catalyst samples (40–60 mesh). A standard reaction gas containing 1000 ppm benzene and 20% O2 in N2 was fed with a total flow rate of 100 mL/min. The weight hourly space velocity (WHSV) was typically 60,000 mL·g−1·h−1. The reactants and the products were analyzed on-line using a GC/MS 6890N gas chromatograph (Hewlett-Packard, Palo Alto, CA, USA) interfaced to a 5973N mass selective detector (Hewlett-Packard, Palo Alto, USA) with a HP-5MS capillary column (30 m × 0.25 mm × 0.25 μm) and another GC (GC112A, Shangfen, Shanghai, China) with a carbon molecule sieve column. The conversion of benzene (Xbenzene, %) was calculated as follows:
X benzene = C benzene ( in ) C benzene ( out ) C benzene ( in ) × 100 %
where, Cbenzene (in) (ppm) and Cbenzene (out) (ppm) are the concentrations of benzene in the inlet and outlet gas, respectively.

3. Results and Discussion

3.1. Catalytic Oxidation Activity of CexMn1−x Composite Oxides for Benzene

The catalytic performance of CeO2, MnOx and CexMn1−x catalysts was evaluated through the oxidation of benzene. The catalytic conversion of benzene as a function of the temperature, 100–400 °C, is shown in Figure 1a. It can be acquired that MnOx exhibits the least active followed by CeO2. With the Mn element adding into CeO2, the activity increases monotonically up to a Mn content of 70 at.% and Ce0.3Mn0.7 is the most active among all catalysts achieving complete benzene conversion at ca. 200 °C. The MnOx and CeO2 catalysts synthesized through oxalate route also achieve full conversion at ca. 300 °C. In addition, the activities of Ce0.3Mn0.7 synthesized through different routes over benzene oxidation are comparative researched in order to identify the advantage of oxalate route (Figure 1b). The result indicates that the sample exhibits higher activity than that of corresponding samples synthesized by hydrothermal or carbonate routes, which is probably related with the microstructure of catalyst. For the purposes of comparison, the reaction temperatures T10%, T50%, T90% (corresponding to the benzene conversion = 10%, 50%, 90%) used to evaluate the performances of the catalysts are summarized in Table 1.

3.2. Characterization of CexMn1−x Catalysts

Figure 2a shows the XRD patterns of the samples in the angular range 10–70 2θ. The diffraction peaks at 2θ = 28.5, 33.0, 47.4, 56.4 and 59.2 in the XRD profile of the pure cerium oxide clearly demonstrate the presence of cubic fluorite structure of CeO2 (JCPDS 81-0792). For pure MnOx, the XRD pattern exhibits complex diffraction peaks which can be considered as a mixture of crystalline Mn3O4 (JCPDS 80-0382) and Mn2O3 (JCPDS 89-4836). The data demonstrates that manganese oxide is a multivalent framework manganese (2+ and 3+), which indicates the existence of various manganese oxides. For CexMn1−x catalyst, the XRD patterns exhibit more complex. The XRD patterns of the CexMn1−x mixed oxide (x ≥ 0.5) do not show any diffraction of manganese oxides, and only broad reflections attributed to CeO2 are observed, which is possible to be related with the formation of solid solution between MnOx and CeO2. The diffraction patterns of CexMn1−x mixed oxides at x ≤ 0.3 show crystallization of Mn3O4 and Mn2O3 except that of CeO2.
The formation of solid solution between MnOx and CeO2 can be further evidenced by the fact that the characteristic diffraction peak of CeO2 in the composite oxides is slightly shifted to higher values of the Bragg angles, compared with the pure CeO2 (Figure 2b). Since the ionic radius of Mn2+ (0.091 nm) and Mn3+ (0.066 nm) are both smaller than that of the Ce4+ (0.1098 nm), the incorporation of Mn2+ or Mn3+ into the CeO2 lattice to form CeO2-MnOx solid solution would result in remarkable decrease in the lattice parameter of CeO2 in the CexMn1−x composite oxide. Meanwhile the O vacancy is also easier to form in order to balance charge due to Ce4+ replaced by Mn2+ or Mn3+ in the CexMn1−x catalyst. The oxygen vacancy is beneficial to catalytic activities of CexMn1−x catalyst [19].
The SEM images of as-prepared CexMn1−x oxide catalysts are presented in Figure 3a. For pure oxide CeO2, it can be seen that the catalyst is composed of many thin flakes, which are overlapped together to form butterfly-like structure (inset picture). The thickness of every flake is about 200 nm and the length can extend to several micrometers. In the SEM image of MnOx, a lot of grains with ellipsoid-like morphology are seen clearly and the size is ca.10 μm. At the surface of every grain, deep ravines are also obviously observed (inset picture). For CexMn1−x composite catalysts, the morphology changes gradually with the Mn content increasing. When the Mn theoretical content reaches 50% (Ce0.5Mn0.5), a few of bulk-like particles with size of several micrometers can be detected except thin flakes. When the theoretical content of Mn ion reaches to 70% (Ce0.3Mn0.7), a large number of grains possessing layered structure can be observed. In addition, it can be acquired that Ce, Mn and O elements are dispersed together homogeneously through element distribution over CexMn1−x composite catalysts.
Through the discussion over catalytic activities of CexMn1−x, it has been acquired that Ce0.3Mn0.7 presents the highest activity and preparation route is also important to the property of catalyst. Therefore, the TEM images of Ce0.3Mn0.7 prepared through different technology routes are shown so as to analyze their microstructures (Figure 3b). In the TEM image of Ce0.3Mn0.7 synthesized by oxalate route, the catalyst is composed of thin flakes. At the surface of flake, some mesoporous structures can be observed and the size of mesoporous is ca. 2 nm. As we known, the oxalate chains chelated in the corresponding precursors are easy to be decomposed into COx and H2O with calcination in air, which can leave behind large numbers of voids due to the release of gaseous COx and H2O [21]. Meanwhile, some primary nanoparticles are assembled together thereby forming porous structure, which is beneficial to absorb and desorb the gas due to the formation of massive active sites and decrease of mass transfer effect [22,23]. For Ce0.3Mn0.7 synthesized by carbonate route, some grains with dumbbell shape can be seen, which are composed of some stacked nanoparticles with the size of 1–2 nm. In the TEM of Ce0.3Mn0.7 prepared through hydrothermal method, a lot of nanorods with the diameter of ca. 10 nm and the length of 300–400 nm are observed, which is also composed of some assembled nanoparticles.
In addition, the N2 adsorption–desorption isotherms and the pore size distribution of the as-prepared catalysts are displayed in Figure S1. The data show that the isotherms of the as-prepared materials possess type IV characteristics with well-developed H3 type hysteresis loops. The result indicates that the CexMn1−x composite catalysts possess porous structure, which is consistent with the result of SEM. The porous structure can facilitate the adsorption and diffusion of reactive molecules, thus greatly reducing limitations of inter-phase mass transfer and enhancing their catalytic activities [21].
The oxidation state of catalyst surface species was examined by XPS analysis. Figure 4 exhibits XPS patterns of Ce 3d, Mn 2p, Mn 3s and O 1s for samples, respectively. In the Ce 3d spectrum of support (Figure 4a), six peaks labeled as V0 (882.1 eV), V1 (888.7 eV), V2 (898.1 eV), V0′ (900.7 eV), V1′ (907.1 eV) and V2′ (916.3 eV) can be identified as characteristic of Ce4+ 3d final states [24,25]. The high BE doublet (V2/V2′) is attributed to the final state of Ce(IV)3d94f0O2p6, doublet V1/V1′ is originated from the state of Ce(IV)3d94f1O2p5, and doublet V0/V0′ corresponds to the state of Ce(IV)3d94f2O2p4. The character peaks of Ce3+ are also observed at 903.4/884.7 eV and 897.6/879.3 eV labeled as U1/U1′ and U0/U0′, respectively [26]. The amount of Ce3+ is estimated to be 11.05, 10.89, 10.05 and 5.65% for CeO2, Ce0.7Mn0.3, Ce0.5Mn0.5 and Ce0.3Mn0.7, which can be calculated according to the Equation (2). Therefore, Ce species in the CexMn1−x composite oxides exist mainly in tetravalent oxidation state.
X C e 3 + = A C e 3 + S C e A ( C e 3 + + C e 4 + ) S C e × 100 %
where X C e 3 + is the percentage content of Ce3+, A is the integrate area of characteristic peak in the XPS pattern, S is sensitivity factors (S = 7.399).
Figure 4 shows the Mn 2p XPS spectra of CexMn1−x composite oxides, in which Mn 2p doublet can be distinguished. The binding energies of the Mn 2p3/2 component appear at 641.7 eV and those for Mn 2p1/2 appear at 653.2 eV. The BE values of the Mn 2p3/2 (641.7 eV) and spin–orbit splitting (11.7 eV) are well matching with the reported values of the trivalent manganese [27]. The shoulder of the Mn 2p3/2 component at 640.7 eV is attributed to Mn2+ species [28]. The XPS results do not provide any evidence for the presence of Mn4+ species (642.2–643 eV) [29,30]. In order to determine the chemical states of Mn further, Mn 3s XPS spectra of CexMn1−x are analyzed (Figure 4c). The spin−orbit splitting value (ΔEs) between the two doublets was 5.44 eV for all samples, closing to the value of 5.1 for the standard sample of α-Mn2O3. The ΔEs of MnO is about 6.3 eV, indicating that the oxidation status of Mn is predominantly tervalent [31,32]. The average oxidation state of Mn was calculated using formula AOS = 8.95 − 1.13 × ΔEs [33]. The data was calculated to be 2.80 that fall in between the average oxidation state of Mn (+2.67) in Mn3O4 and the state of Mn (+3) in Mn2O3. Therefore, the element Mn in the CexMn1−x catalysts is existed in the form of Mn3O4 and Mn2O3, which is consistent with the data of XRD. In addition, the peak of Mn 2p3/2 (641.5 eV) for CexMn1−x composites is shifted to lower binding energy comparing with pure MnOx, which may be a consequence of interaction between CeO2 and MnOx [19]. It is worthy to note that the pattern and the intensity of Ce0.7Mn0.3 is similar and less weaker compared with other CexMn1−x composites. The corresponding XPS spectrum can be seen in the Figure S2.
The XPS O1s spectra (Figure 4d) show a main peak at a binding energy of 529.1–529.9 eV, corresponding to lattice oxygen of CeO2 and MnOx phases(O2−; denoted as Oα) [27,30]. A broad shoulder at the higher binding energy region (531.3–531.8 eV) is ascribed to defective oxides or oxygen species of the surface carbonates and hydroxide (denoted as Oβ) [18,34]. It is worthy to note that the peak corresponding to lattice oxygen in CexMn1−x composite catalyst with higher Mn content tends to shift toward higher BE value than that of pure CeO2 (from 529.1 eV to 529.6 eV), which suggests that the environments of oxygen change with increasing Mn content. This appearance is also attributed to the interaction between CeO2 and MnOx. In addition, the content of Oα is calculated according to the Equation (3) and listed in Table 2. The data shows that the Ce0.3Mn0.7 sample possesses more lattice oxygen species, confirming that the mobility and availability of lattice oxygen species are enhanced due to the synergistic effects of CeO2 and MnOx in Ce0.3Mn0.7 [35].
X O α = A O α S O A ( O α + O β ) S O × 100 %
where X O α is the percentage content of Oα, A is the integrate area of characteristic peak in the XPS pattern, S is sensitivity factors (S = 0.711).
In order to check the redox properties of the new series of CexMn1−x systems, TPR of all the catalysts were carried out. Figure 5a shows the H2-TPR profiles of CeO2, MnOx and CexMn1−x composite oxides. Similar to previous findings [36,37,38], pure CeO2 exhibits two reduction peaks at around 405 °C and about 719 °C. The former low-temperature reduction is due to the removal of surface oxygen and the later high-temperature reduction is related to the oxygen species in bulk CeO2. The H2-TPR profile of pure MnOx shows two strong reduction peaks at 167 °C and 330 °C, respectively, with an area ratio of the lower to the higher temperature hydrogen consumption of about 1:2.42. The actual hydrogen consumption of two reduction peaks is 0.92343 and 1.88118 mmol/g and the corresponding ratio is 1:2.04. As we known, Mn2O3 proceed two-step reduction, the low-temperature reduction peak represented the reduction of Mn2O3 to Mn3O4 and the high-temperature reduction peak referred to the further reduction of Mn3O4 to MnO [39]. The theoretical hydrogen consumption ratio is 1:2 as seen in Formula (4) and (5), which is less than the fitted value (2.42) and actual data (2.04). It indicates that the hydrogen gas is more consumed at higher temperature, which is attributed to the extra existence of Mn3O4 phase. This is in agreement with the XRD data, which show that the crystalline phase of pure MnOx corresponds to Mn2O3 and Mn3O4.
3 M n 2 O 3 + H 2 2 M n 3 O 4 + H 2 O
M n 3 O 4 + H 2 3 M n O + H 2 O
In contrast to pure CeO2 and MnOx, the reduction profiles of CexMn1−x catalysts are more complicated. For CexMn1−x, the TPR profiles consist of three overlapping peaks at lower temperature and one peak at higher temperature. According to the reduction characteristics of pure MnOx and CeO2, it can be deduced that the lower temperature peaks (182/351 °C, 189/340 °C, 169/343 °C) are assigned to the two-step reduction of Mn2O3. The peaks at 693 °C or 714 °C are attributed to the oxygen species in bulk CeO2. It is worthy to note that another obvious peak at lower temperature (295 °C, 283 °C or 232 °C) as shown in Figure 5b is possible to be caused by the synergistic effect between Mn2+/Mn3+ and Ce4+, which is related with CeO2-MnOx solid solution. It can facilitate the mobility of the oxygen species in the CexMn1−x composite oxide. Therefore, their catalytic activities over benzene are higher than pure CeO2 and MnOx. Additionally, the reduction temperature is lower and the peak corresponding to the oxygen species in bulk CeO2 disappear in the TPR pattern of Ce0.3Mn0.7 compared with those of Ce0.7Mn0.3/Ce0.5Mn0.5. This indicates that the reduction of the manganese oxide and the cerium oxide in Ce0.3Mn0.7 is promoted more obvious, which results in the highest activity over benzene in the CexMn1−x composite oxides. In the view of hydrogen consumption, the same conclusion can be also obtained. Ce0.3Mn0.7 consumed the most hydrogen gas (4.88447 mmol/g) among CexMn1−x composite oxides indicating that Ce0.3Mn0.7 possesses more oxygen species, which are beneficial to benzene oxidation reaction. Therefore, Ce0.3Mn0.7 presents higher activity compared with other CexMn1−x.

3.3. Factors Influencing the Catalytic Activity

Through the analysis, it has been acquired that Ce0.3Mn0.7 performs excellently in catalytic oxidation of benzene among the CexMn1−x catalysts. Many factors are considered to determine catalytic performances. Firstly, the existence of Ce-Mn solid solution results in the formation of oxygen vacancies due to incorporate Mn into CeO2 crystal lattice, which can enhance their activities. Secondly, the large numbers of active sites will be introduced after removal of oxalate chains, because of the formation of porous structure which can also facilitate the adsorption and diffusion of organic molecules, thus reducing limitations of interphase mass transfer and promoting their catalytic activities. Thirdly, the better reducibility at low temperature also plays a great role in the catalytic activity. Through the analysis of TPR, it has been acquired that the reductions of CexMn1−x catalysts start at the lower temperature compared with CeO2, which indicate that the catalysts possess more highly reducible surface species such as absorbed oxygen. Additionally, the existence of special peak caused by the strong interaction between Ce and Mn compared with MnOx also result in the enhancement of reducibility. Finally, the oxidation of organic molecules over transition metal oxide or mixed metal oxide catalysts involves two identical mechanisms: a Langmuir–Hinshelwood mechanism at lower temperature and a Mars–van Krevelen mechanism with increasing reaction temperature [28,40]. At lower temperature, the adsorbed oxygen species with higher activity can enhance the adsorption and oxidation of VOCs. With the temperature rising, the adsorbed organic molecules are oxidized by the oxygen of metal oxides, which can be replenished by gas phase oxygen. Therefore, the adsorbed oxygen species will have an important role to play in determining its catalytic activity. As displayed in Table 2, the Ce0.3Mn0.7 exhibits a higher content of Ce4+ and Mn3+. The high-valence of cerium and manganese ions are preferred to adsorb more active oxygen species to attend in the reaction, thereby Ce0.3Mn0.7 possessed higher activity.

4. Conclusions

A series of CexMn1−x composite oxides, CeO2 and MnOx were synthesized through oxalate method and the complete catalytic oxidation of benzene were examined. The results indicated that CexMn1−x catalysts exhibited better activities comparing with pure CeO2 or MnOx, among which the catalytic activity reached the best when the Ceat/Mnat optimum ratio was 3:7. In order to identify the advantage of oxalate route, Ce-Mn composite oxides were also synthesized through carbonate method and hydrothermal method. The results indicated that the samples prepared by oxalate route exhibited higher activities, which were probably related with the microstructure of catalyst. Additionally, the influence of oxygen vacancy and synergistic effect in the benzene catalytic oxidation can’t be also ignored.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/9/2/197/s1, Figure S1: N2 adsorption-desorption isotherms and pore-size distribution of CeO2, MnOx and CexMn1−x catalysts. Figure S2: X-ray photoelectron spectra in the Mn 2p (a) and Mn 3s (b) regions for the Ce0.7Mn0.3 catalysts.

Author Contributions

Conceptualization: M.Y., Z.W., Q.W.; methodology: M.Y., G.S.; software: M.L.; validation: M.Y., Z.W.; formal analysis: M.Y., G.S.; investigation: M.Y., G.S.; resources: Z.W., Q.W.; data curation: M.Y., G.S.; writing—original draft preparation: M.Y., G.S.; writing—review and editing: Z.W., Q.W.; visualization: M.L.; supervision: Z.W., Q.W.; project administration: Z.W.; funding acquisition: Y.C.

Funding

This research was financially supported by the National Key R&D Program of China (2016YFC0207100). We also acknowledge the National Natural Science Foundation of China (No. 51402061).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Garcia, T.; Solsona, B.; Cazorla-Amoros, D.; Linares-Solano, A.; Taylor, S.H. Total oxidation of volatile organic compounds by vanadium promoted palladium-titania catalysts: Comparison of aromatic and polyaromatic compounds. Appl. Catal. B Environ. 2006, 62, 66–76. [Google Scholar] [CrossRef]
  2. Tang, X.F.; Xu, Y.D.; Shen, W.J. Promoting effect of copper on the catalytic activity of MnOx-CeO2 mixed oxide for complete oxidation of benzene. Chem. Eng. J. 2008, 144, 175–180. [Google Scholar] [CrossRef]
  3. Einaga, H.; Futamura, S. Comparative study on the catalytic activities of alumina-supported metal oxides for oxidation of benzene and cyclohexane with ozone. React. Kinet. Catal. Lett. 2004, 81, 121–128. [Google Scholar] [CrossRef]
  4. Einaga, H.; Ogata, A. Benzene oxidation with ozone over supported manganese oxide catalysts: Effect of catalyst support and reaction conditions. J. Hazard. Mater. 2009, 164, 1236–1241. [Google Scholar] [CrossRef] [PubMed]
  5. Einaga, H.; Teraoka, Y.; Ogat, A. Benzene oxidation with ozone over manganese oxide supported on zeolite catalysts. Catal. Today 2011, 164, 571–574. [Google Scholar] [CrossRef]
  6. Liotta, L.F.; Ousmane, M.; Di Carlo, G.; Pantaleo, G.; Deganello, G.; Boreave, A.; Giroir-Fendler, A. Catalytic Removal of Toluene over Co3O4-CeO2 Mixed Oxide Catalysts: Comparison with Pt/Al2O3. Catal. Lett. 2009, 127, 270–276. [Google Scholar] [CrossRef]
  7. Liotta, L.F.; Ousmane, M.; Di Carlo, G.; Pantaleo, G.; Deganello, G.; Marci, G.; Retailleau, L.; Giroir-Fendler, A. Total oxidation of propene at low temperature over Co3O4-CeO2 mixed oxides: Role of surface oxygen vacancies and bulk oxygen mobility in the catalytic activity. Appl. Catal. A Gen. 2008, 347, 81–88. [Google Scholar] [CrossRef]
  8. Lojewska, J.; Kolodziej, A.; Lojewski, T.; Kapica, R.; Tyczkowski, J. Structured cobalt oxide catalyst for VOC combustion. Part I: Catalytic and engineering correlations. Appl. Catal. A Gen. 2009, 366, 206–211. [Google Scholar] [CrossRef]
  9. Walerczyk, W.; Zawadzki, M. Structural and catalytic properties of Pt/ZnAl2O4 as catalyst for VOC total oxidation. Catal. Today 2011, 176, 159–162. [Google Scholar] [CrossRef]
  10. Zwinkels, M.F.M.; Jaras, S.G.; Menon, P.G.; Griffin, T.A. Catalytic Materials for High-Temperature Combustion. Catal. Rev.-Sci. Eng. 1993, 35, 319–358. [Google Scholar] [CrossRef]
  11. Lahousse, C.; Bernier, A.; Grange, P.; Delmon, B.; Papaefthimiou, P.; Ioannides, T.; Verykios, X. Evaluation of gamma-MnO2 as a VOC removal catalyst: Comparison with a noble metal catalyst. J. Catal. 1998, 178, 214–225. [Google Scholar] [CrossRef]
  12. Dai, Q.G.; Wang, X.Y.; Lu, G.Z. Low-temperature catalytic combustion of trichloroethylene over cerium oxide and catalyst deactivation. Appl. Catal. B Environ. 2008, 81, 192–202. [Google Scholar] [CrossRef]
  13. Kim, H.J.; Choi, S.W.; Inyang, H.I. Catalytic oxidation of toluene in contaminant emission control systems using Mn-Ce/γ-Al2O3. Environ. Technol. 2008, 29, 559–569. [Google Scholar] [CrossRef]
  14. Saqer, S.M.; Kondarides, D.I.; Verykios, X.E. Catalytic Activity of Supported Platinum and Metal Oxide Catalysts for Toluene Oxidation. Top. Catal. 2009, 52, 517–527. [Google Scholar] [CrossRef]
  15. Qi, G.S.; Yang, R.T. Performance and kinetics study for low-temperature SCR of NO with NH3 over MnOx-CeO2 catalyst. J. Catal. 2003, 217, 434–441. [Google Scholar] [CrossRef]
  16. Qi, G.S.; Yang, R.T.; Chang, R. MnOx-CeO2 mixed oxides prepared by co-precipitation for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  17. Silva, A.M.T.; Marques, R.R.N.; Quinta-Ferreira, R.M. Catalysts based in cerium oxide for wet oxidation of acrylic acid in the prevention of environmental risks. Appl. Catal. B Environ. 2004, 47, 269–279. [Google Scholar] [CrossRef] [Green Version]
  18. Tang, X.F.; Li, Y.G.; Huang, X.M.; Xu, Y.D.; Zhu, H.Q.; Wang, J.G.; Shen, W.J. MnOx-CeO2 mixed oxide catalysts for complete oxidation of formaldehyde: Effect of preparation method and calcination temperature. Appl. Catal. B Environ. 2006, 62, 265–273. [Google Scholar] [CrossRef]
  19. Wang, Z.; Shen, G.L.; Li, J.Q.; Liu, H.D.; Wang, Q.; Chen, Y.F. Catalytic removal of benzene over CeO2-MnOx composite oxides prepared by hydrothermal method. Appl. Catal. B Environ. 2013, 138, 253–259. [Google Scholar] [CrossRef]
  20. Wang, Z.; Yang, M.; Shen, G.L.; Liu, H.D.; Chen, Y.F.; Wang, Q. Catalytic removal of benzene over CeO2-MnOx composite oxides with rod-like morphology supporting PdO. J. Nanopart. Res. 2014, 16, 2367. [Google Scholar] [CrossRef]
  21. Tang, W.X.; Wu, X.F.; Li, S.D.; Li, W.H.; Chen, Y.F. Porous Mn-Co mixed oxide nanorod as a novel catalyst with enhanced catalytic activity for removal of VOCs. Catal. Commun. 2014, 56, 134–138. [Google Scholar] [CrossRef]
  22. Zhang, R.Z.; Dai, H.X.; Du, Y.C.; Zhang, L.; Deng, J.G.; Xia, Y.S.; Zhao, Z.X.; Meng, X.; Liu, Y.X. P123-PMMA Dual-Templating Generation and Unique Physicochemical Properties of Three-Dimensionally Ordered Macroporous Iron Oxides with Nanovoids in the Crystalline Walls. Inorg. Chem. 2011, 50, 2534–2544. [Google Scholar] [CrossRef]
  23. Chen, X.; Shen, Y.F.; Suib, S.L.; O’Young, C.L. Characterization of manganese oxide octahedral molecular sieve (M-OMS-2) materials with different metal cation dopants. Chem. Mater. 2002, 14, 940–948. [Google Scholar] [CrossRef]
  24. Pfau, A.; Schierbaum, K.D. The Electronic-Structure of Stoichiometric and Reduced CeO2 Surfaces—An XPS, UPS and HREELS Study. Surf. Sci. 1994, 321, 71–80. [Google Scholar] [CrossRef]
  25. Xiao, W.D.; Guo, Q.L.; Wang, E.G. Transformation of CeO2 (111) to Ce2O3 (0001) films. Chem. Phys. Lett. 2003, 368, 527–531. [Google Scholar] [CrossRef]
  26. Lenormand, F.; Hilaire, L.; Kili, K.; Krill, G.; Maire, G. Oxidation-State of Cerium in Cerium-Based Catalysts Investigated by Spectroscopic Probes. J. Phys. Chem. 1988, 92, 2561–2568. [Google Scholar] [CrossRef]
  27. 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]
  28. Santos, V.P.; Pereira, M.F.R.; Orfao, J.J.M.; Figueiredo, J.L. The role of lattice oxygen on the activity of manganese oxides towards the oxidation of volatile organic compounds. Appl. Catal. B Environ. 2010, 99, 353–363. [Google Scholar] [CrossRef]
  29. Rao, T.; Shen, M.Q.; Jia, L.W.; Hao, J.J.; Wang, J. Oxidation of ethanol over Mn-Ce-O and Mn-Ce-Zr-O complex compounds synthesized by sol-gel method. Catal. Commun. 2007, 8, 1743–1747. [Google Scholar] [CrossRef]
  30. Dula, R.; Janik, R.; Machej, T.; Stoch, J.; Grabowski, R.; Serwicka, E.M. Mn-containing catalytic materials for the total combustion of toluene: The role of Mn localisation in the structure of LDH precursor. Catal. Today 2007, 119, 327–331. [Google Scholar] [CrossRef]
  31. Han, Y.F.; Chen, F.X.; Zhong, Z.Y.; Ramesh, K.; Chen, L.W.; Widjaja, E. Controlled synthesis, characterization, and catalytic properties of Mn2O3 and Mn3O4 nanoparticles supported on mesoporous silica SBA-15. J. Phys. Chem. B 2006, 110, 24450–24456. [Google Scholar] [CrossRef]
  32. Luo, Y.; Deng, Y.Q.; Mao, W.; Yang, X.J.; Zhu, K.K.; Xu, J.; Han, Y.F. Probing the Surface Structure of α-Mn2O3 Nanocrystals during CO Oxidation by Operando Raman Spectroscopy. J. Phys. Chem. C 2012, 116, 20975–20981. [Google Scholar] [CrossRef]
  33. Galakhov, V.R.; Demeter, M.; Bartkowski, S.; Neumann, M.; Ovechkina, N.A.; Kurmaev, E.Z.; Logachevskaya, N.I.; Mukovskii, Y.M.; Mitchell, J.; Ederer, D.L. Mn 3s exchange splitting in mixed-valence manganites. Phys. Rev. B 2002, 65, 113102. [Google Scholar] [CrossRef]
  34. Wang, H.; Zhu, H.Q.; Qin, Z.F.; Liang, F.X.; Wang, G.F.; Wang, J.G. Deactivation of a Au/CeO2-Co3O4 catalyst during CO preferential oxidation in H2-rich stream. J. Catal. 2009, 264, 154–162. [Google Scholar] [CrossRef]
  35. Wu, Y.S.; Zhang, Y.X.; Liu, M.; Ma, Z.C.C. Complete catalytic oxidation of o-xylene over Mn-Ce oxides prepared using a redox-precipitation method. Catal. Today 2010, 153, 170–175. [Google Scholar] [CrossRef]
  36. Trovarelli, A. Catalytic properties of ceria and CeO2-containing materials. Catal. Rev. 1996, 38, 439–520. [Google Scholar] [CrossRef]
  37. Aneggi, E.; Boaro, M.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Insights into the redox properties of ceria-based oxides and their implications in catalysis. J. Alloy Compd. 2006, 408, 1096–1102. [Google Scholar] [CrossRef]
  38. Tana; Zhang, M.L.; Li, J.; Li, H.J.; Li, Y.; Shen, W.J. Morphology-dependent redox and catalytic properties of CeO2 nanostructures: Nanowires, nanorods and nanoparticles. Catal. Today 2009, 148, 179–183. [Google Scholar] [CrossRef]
  39. Tang, X.F.; Chen, J.L.; Li, Y.G.; Li, Y.; Xu, Y.D.; Shen, W.J. Complete oxidation of formaldehyde over Ag/MnOx-CeO2 catalysts. Chem. Eng. J. 2006, 118, 119–125. [Google Scholar] [CrossRef]
  40. Tang, W.X.; Wu, X.F.; Li, D.Y.; Wang, Z.; Liu, G.; Liu, H.D.; Chen, Y.F. Oxalate route for promoting activity of manganese oxide catalysts in total VOCs’ oxidation: Effect of calcination temperature and preparation method. J. Mater. Chem. A 2014, 2, 2544–2554. [Google Scholar] [CrossRef]
Figure 1. C6H6 conversion (%, (a)) over CeO2, MnOx and CexMn1−x catalysts synthesized by oxalate method; C6H6 conversion comparison as a function of reaction temperature over Ce3Mn7 through different reaction route (b).
Figure 1. C6H6 conversion (%, (a)) over CeO2, MnOx and CexMn1−x catalysts synthesized by oxalate method; C6H6 conversion comparison as a function of reaction temperature over Ce3Mn7 through different reaction route (b).
Nanomaterials 09 00197 g001
Figure 2. XRD patterns of all the samples: (a) wild angle patterns, and (b) Enlarged-zone patterns. Crystalline phases detected: (o) CeO2, (□) Mn3O4, (∆) Mn2O3.
Figure 2. XRD patterns of all the samples: (a) wild angle patterns, and (b) Enlarged-zone patterns. Crystalline phases detected: (o) CeO2, (□) Mn3O4, (∆) Mn2O3.
Nanomaterials 09 00197 g002
Figure 3. SEM images and EDX mapping of CeO2, MnOx, Ce0.5Mn0.5 and Ce0.3Mn0.7 (a); TEM images of Ce0.3Mn0.7 synthesized through different routes (b).
Figure 3. SEM images and EDX mapping of CeO2, MnOx, Ce0.5Mn0.5 and Ce0.3Mn0.7 (a); TEM images of Ce0.3Mn0.7 synthesized through different routes (b).
Nanomaterials 09 00197 g003
Figure 4. X-ray photoelectron spectra in the Ce 3d (a), Mn 2p (b), Mn 3s (c), O 1s (d) regions for the CexMn1−x catalysts.
Figure 4. X-ray photoelectron spectra in the Ce 3d (a), Mn 2p (b), Mn 3s (c), O 1s (d) regions for the CexMn1−x catalysts.
Nanomaterials 09 00197 g004
Figure 5. H2-TPR of CeO2, MnOx, CexMn1−x (a) and Enlarged picture of the red circle area (b).
Figure 5. H2-TPR of CeO2, MnOx, CexMn1−x (a) and Enlarged picture of the red circle area (b).
Nanomaterials 09 00197 g005
Table 1. Catalytic activities of CeO2, MnOx and CexMn1−x catalysts.
Table 1. Catalytic activities of CeO2, MnOx and CexMn1−x catalysts.
CatalystBenzene Oxidation Activity
T10% (°C)T50% (°C)T90% (°C)
CeO2145200257
Ce0.7Mn0.3153200220
Ce0.5Mn0.5128185218
Ce0.3Mn0.7117156190
MnOx194241273
Ce0.3Mn0.7 (Hydrothermal)200265350
Ce0.3Mn0.7 (Carbonate)150220250
Table 2. XPS results of CexMn1−x samples.
Table 2. XPS results of CexMn1−x samples.
SampleOOα/(Oα + Oβ) (%)Ce (%)Mn (%)
OαOβCe3+/(Ce3+ + Ce4+)Mn2+/(Mn2+ + Mn3+)
CeO2529.0531.377.1611.05-
Ce0.7Mn0.3529.1531.381.0110.8916.23
Ce0.5Mn0.5529.1531.382.0410.0515.12
Ce0.3Mn0.7529.2531.583.185.6513.42
MnOx529.6531.678.47-10.69

Share and Cite

MDPI and ACS Style

Yang, M.; Shen, G.; Liu, M.; Chen, Y.; Wang, Z.; Wang, Q. Preparation of Ce–Mn Composite Oxides with Enhanced Catalytic Activity for Removal of Benzene through Oxalate Method. Nanomaterials 2019, 9, 197. https://doi.org/10.3390/nano9020197

AMA Style

Yang M, Shen G, Liu M, Chen Y, Wang Z, Wang Q. Preparation of Ce–Mn Composite Oxides with Enhanced Catalytic Activity for Removal of Benzene through Oxalate Method. Nanomaterials. 2019; 9(2):197. https://doi.org/10.3390/nano9020197

Chicago/Turabian Style

Yang, Min, Genli Shen, Mi Liu, Yunfa Chen, Zhen Wang, and Qi Wang. 2019. "Preparation of Ce–Mn Composite Oxides with Enhanced Catalytic Activity for Removal of Benzene through Oxalate Method" Nanomaterials 9, no. 2: 197. https://doi.org/10.3390/nano9020197

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