Home Effect of calcination temperature on rare earth tailing catalysts for catalytic methane combustion
Article Open Access

Effect of calcination temperature on rare earth tailing catalysts for catalytic methane combustion

  • Ran Zhao , ZiChen Tian and Zengwu Zhao EMAIL logo
Published/Copyright: December 2, 2020
Become an author with De Gruyter Brill

Abstract

Bayan Obo tailings are rich in rare earth elements (REEs), iron, and other catalytic active substances. In this study, mine tailings were calcined at different temperatures and tested for the catalytic combustion of low-concentration methane. Upon calcination at 600°C, high catalytic activity was revealed, with 50% CH4 conversion at 587°C (space velocity of 12,000 mL/g h). The physicochemical properties of catalysts were characterized using thermogravimetric analysis, X-ray diffraction, scanning electron microscopy, hydrogen temperature-programmed reduction (H2-TPR), and X-ray photoelectron spectroscopy (XPS). Compared to the raw ore sample, the diffraction peak intensity of Fe2O3 increased post calcination, whereas that of CeCO3F decreased. A porous structure appeared after the catalyst was calcined at 600°C. Additionally, Fe, Ce, Ti, and other metal elements were more highly dispersed on the catalyst surface. H2-TPR results revealed a broadening of the reduction temperature range for the catalyst calcined at 600°C and an increase in the reduction peak. XPS analysis indicated the presence of Ce in the form of Ce3+ and Ce4+ oxidation states and the coexistence of Fe in the form of Fe2+ and Fe3+. Moreover, XPS revealed a higher surface Oads/Olatt ratio. This study provides evidence for the green reuse of Bayan Obo mine tailings in secondary resources.

1 Introduction

Coal mine gas is an associated gas in coal seams. The main component of coal mine gas is methane (CH4), which is often released into the atmosphere during mining through ventilation pipes. The greenhouse contribution of CH4 to global warming is the emission of harmful NOx, CO, and other hydrocarbons. However, the synthesis of a reasonably designed catalyst with good low-temperature ignition performance and high-temperature thermal stability remains a challenge [6,7]. Although noble metal catalysts, particularly, palladium-based catalysts exhibit high activity in the catalytic, 21 times higher than the greenhouse effect of carbon dioxide (CO2) [1,2,3,4,5]. Therefore, effectively eliminating low-concentration CH4 is highly significant for environmental protection. The catalytic combustion of CH4 has many advantages over thermal combustion: reduction in ignition temperature, complete combustion, and reduction in combustion of CH4; their price and easy sintering at high temperatures limit their industrial application [8,9,10]. Therefore, searching for a nonnoble metal catalyst to completely oxidize methane at low temperature is important.

The different metal components in a mixed metal oxide can interact with each other, e.g., through electron coordination or for structural stability, which increases their catalytic activity compared to the corresponding single-component catalyst. Currently, catalysts containing Cu and Mn exhibit high catalytic activity in the combustion of CH4. Fe and Co catalysts modified by the addition of rare earth elements (REEs) as additives have attracted research attention [11,12,13]. Zhang et al. prepared an Mn–Ni catalyst using a coprecipitation method and investigated its role in the catalytic combustion of low-concentration CH4 [14]. Li et al. showed that a Ce1−xFexO2−δ catalyst exhibits not only improved reducibility but also increased lattice oxygen, thereby increasing the catalytic oxidation capacity of the catalyst [15]. Zhang et al. showed that the addition of CeO2, ZrO2, La2O3, and CeO2–ZrO2 as additives can significantly improve the specific surface area of the catalyst, promote the dispersion of each active component, and increase the surface oxygen concentration; thus, the activity of the Cu–Mn–O/Al2O3/COR catalyst is significantly improved [16].

The Bayan Obo deposit in China is the largest REE resource worldwide [17]. The mine tailings in Bayan Obo are generated via separation and flotation. Till date, 160 million tons of mine tailings have been stockpiled in the Bayan Obo tailings dams. Using or recycling these resources is difficult owing to many complex factors, such as low grade, fine grain size, and a complex mineral composition [18,19]. However, the mine tailings still contain REEs and transition metal elements, of which REEs and iron oxides are a common raw material for catalyst preparation. Elements, such as Fe, Ce, and Mn, present in the mine trailing may display synergistic effects with other components to increase the catalytic activity in methane combustion. Therefore, this study uses mine tailings as a raw material to prepare a catalyst for the catalytic combustion of low-concentration CH4. This reduces greenhouse gas emission and realizes the reuse of mine tailing resources.

2 Experiment

2.1 Catalyst preparation

The raw materials used in this experiment were mine tailings containing REEs (henceforth referred to as “mine tailings”) from the Bayan Obo area [20]. Tables 1 and 2 display the main elements present in the mine tailings. The composition of the mine tailings is complex; they contain numerous metal elements, including RE oxides (REO) (5.88%) and Fe2O3 (27.7%). Table 3 displays the major mineral species in the mine tailings, including iron ore, fluorite, bastnaesite, monazite, barite, and ankerite.

Table 1

Chemical composition of Bayan Obo mine tailings

ComponentsSiO2Fe2O3CaOTiO2Na2OMnO2MgOAl2O3BaOFREOOthers
Amount (wt%)11.927.727.21.001.281.963.311.262.708.925.886.89
Table 2

Content of RE oxides in Bayan Obo mine tailings

ComponentsREOCeO2Pr2O3Nd2O3La2O3
Amount (wt%)5.883.010.331.101.44
Table 3

Content of major mineral species in Bayan Obo mine tailings

Mineral speciesHematite/MagnetiteBastnaesiteFluoriteAnkeriteBariteMonaziteApatiteAmphibolePyroxene
Amount (wt%)30.019.8317.586.518.633.722.793.511.76

The following steps were performed for catalyst preparation: a certain amount of tailings containing REEs were crushed, ground, sieved, and dried and were then divided using 100–200, 200–300, 300–400, and 400–500 meshes. The particle size with the best catalytic activity was calcined at 400°C, 500°C, 600°C, and 700°C for 4 h, and the resulting samples were labeled as 2, 3, 4, and 5, respectively. The original REE-containing mine tailing sample of dolomite was labeled as 1.

2.2 Catalytic activity tests

Figure 1 shows the setup for measuring the activity for catalytic oxidation of methane. The catalysts’ activity was evaluated using a continuous-flow fixed-bed quartz tube microreactor (diameter 10 mm). The reaction conditions were as follows: 2% CH4, 18% O2, and N2 as the balance gas, ordinary atmospheric pressure, a volumetric space velocity (SV) of 20,000 mL/g h, and 500 mg of catalyst. The reaction temperature was increased from 300°C to 750°C at a heating rate of 5°C/min, and data were recorded at each temperature point that is stabilized for 30 min. Finally, the methane content was monitored online via gas chromatography (Agilent 8890B) using a thermal conductivity detector (TCD). The catalytic efficiency of methane was calculated as follows:

(1)η=ω0ω1ω0×100%

where η is the conversion rate of CH4, ω0 is the methane content before the reaction, and ω1 is the methane content after stabilization. The activity of the methane combustion catalyst was preliminarily evaluated using three performance indicators: the temperature at which methane conversion is 10%, 50%, and 100%, referred to as T10, T50, and T90, respectively.

Figure 1 Experimental setup for the catalytic oxidation of methane. (1) N2/CH4 mixed gas cylinders; (2) O2 gas cylinders; (3) pressure gauge; (4) pressure-reducing valve; (5) mass flow meter; (6) reaction valve; (7) quartz tube; (8) thermocouple; (9) furnace wall; (10) catalyst; (11) quartz cotton; (12) gas chromatograph; (13) exhaust gas.
Figure 1

Experimental setup for the catalytic oxidation of methane. (1) N2/CH4 mixed gas cylinders; (2) O2 gas cylinders; (3) pressure gauge; (4) pressure-reducing valve; (5) mass flow meter; (6) reaction valve; (7) quartz tube; (8) thermocouple; (9) furnace wall; (10) catalyst; (11) quartz cotton; (12) gas chromatograph; (13) exhaust gas.

The kinetic parameters for CH4 combustion were measured at a conversion rate below 20%. The reaction rate of CH4 (r, mol/gcat s) conversion was calculated as follows:

(2)r=VCH4ηWcat

where VCH4·η, and Wcat are the CH4 flow rate (mol/s) and the CH4 conversion catalyst weight, respectively. According to the Arrhenius formula, the activation energy (Ea) of the catalyst for CH4 combustion can be obtained from the slope of a linear plot of ln r versus 1/T, according to the following equation:

(3)lnr=Ea/RT+C

2.3 Characterization of catalyst

To elucidate the relation between structure and catalytic activity, the catalysts were characterized. Herein, several techniques were applied. The thermogravimetric (TG) analysis curve was measured using STA449C, NETZSCH (Germany). The online mass spectrometer (MS) (HPR20) used in this study was manufactured by HIDEN (UK). X-ray fluorescence (XRF) was obtained using a Rigaku ZSX primus (Japan). XRF samples were prepared on glass disk melts in lithium tetraborate using automatic fusion apparatus with a mass ratio of 1:15. The operating voltage and current were 60 kV and 60 mA, respectively. XRF spectrometry was performed under vacuum. X-ray diffraction (XRD) results were recorded on a Bruker D8 Advance X-ray diffractometer, wherein the radiation source was Cu-Kα, the scanning angle range was 20°–80° with a scanning speed of 3°/min, and the voltage and current were 40 kV and 40 mA, respectively. Hydrogen temperature-programmed reduction (H2-TPR) measurement was performed on a temperature rise chemical adsorption instrument (PCA-1200) with a TCD. Scanning electron microscopy (SEM) was performed using a Zeiss Sigma-500 field scanning electron microscope with a voltage of 30.0 kV. X-ray photoelectron spectroscopy (XPS) measurements were conducted on a PHI Quantera II instrument. Calcination was mainly performed in a VTL1600 vertical tube furnace under an air atmosphere; therein, the inner diameter of the corundum tube is 80 mm, its length is 130 mm, and the heating element is a silicon and molybdenum rod.

3 Results and discussion

3.1 Thermogravimetric-mass spectrum analysis of catalyst

Figure 2a shows the TG curves of the REE-containing mine tailings. Between 200°C and 400°C, the TG curve shows a small weight loss phenomenon caused by the preliminary decomposition of cerium fluorocarbon or carbonate during calcination. Between 400°C and 700°C, the TG curve shows continuous and rapid weight loss caused by the further decomposition of fluorocarbon–cerium ore into the oxides of the REEs by heating [18,19,20]. When the mine tailings are heated, they enter an endothermic state. Between 650°C and 930°C, weight loss continues to occur and an endothermic peak is present, which is caused by the sintering or melting of the REE-rich tailings.

Figure 2 Thermogravimetric analysis of the REE-containing tailings: (a) thermogravimetric analysis and (b) CO2 mass spectrum signal.
Figure 2

Thermogravimetric analysis of the REE-containing tailings: (a) thermogravimetric analysis and (b) CO2 mass spectrum signal.

During the calcination process, the RECO3F completely decomposed to REOF, as the rare earth fluoride oxide generates phase separation to produce rare earth oxide and rare earth fluoride as the calcination temperature increases. Rare earth oxides include CeO2 and other substances that are effective for catalysis. The presence of CeO2 is important as it promotes the catalytic combustion of methane [20]. In summary, the catalyst’s calcination changes the mineral phase of the mine tailings. TG provides support for determining the optimal calcination temperature, with the disappearance of the fluorocarbon cerium ore peak representing the complete decomposition of the fluorocarbon cerium ore.

Figure 2b shows the mass spectrum of the REE-rich tailings. The first peak at 520°C is attributed to the decomposition of cerium fluorocarbon. The following equation shows the decomposition of the fluorocarbon cerium ore [20]:

(4)RECO3F+O2REOF+CO2
(5)REOFCe0.75Nd0.25O1.875+(Ce,Pr)La2O3F3

The second peak at 680°C is attributed to the decomposition of MgCO3 in iron dolomite. The decomposition of dolomite is divided into two steps: CaMg(CO3)2 decomposes in the dolomite sample to CaCO3 and MgO at approximately 630–680°C, followed by [22]:

(6)CaMg(CO3)2CaCO3+MgO+CO2

3.2 Catalytic performance of methane combustion

The particle size of the mineral significantly influences the catalytic combustion of methane. As shown in Figure 3, the catalytic activity of the mine tailings with particle size above 300 mesh is significantly higher than that with particle size below 300 mesh. Typically, the catalytic performance of methane combustion increases with decreasing particle size of the original ore due to the increase in the relative dispersion with decreasing particle size, which produces a large contact area between the active components and CH4. The catalytic activity of the mine tailings with the particle size of 300–400 mesh and 400–500 mesh was similar. However, the catalytic activity of tailings containing particles of 300–400 mesh size was slightly higher, and thus, these tailings were used further in the experiment.

Figure 3 Effect of the original mineral particle size on the methane conversion.
Figure 3

Effect of the original mineral particle size on the methane conversion.

As shown in Figure 4, the effects of the calcination temperature of the catalysts and the reaction temperature on the catalytic combustion of methane were studied herein. The activity of the catalytic combustion of methane increases as the transformation temperature decreases. Table 4 summarizes the corresponding results. The catalytic activity increases with the calcination temperature. At a calcination temperature of 600°C, the catalyst is optimized. Compared to the original mine tailings, T10 and T90 are decreased by 56°C and 85°C, respectively. The catalytic activity of the catalyst decreases at a calcination temperature of 700°C. In summary, the catalytic activity decreases as follows: sample 4 > sample 3 > sample 2 > sample 1 > sample 5. Cerium oxide will form at approximately 500°C; however, at higher calcination temperatures, structural changes will occur, which will reduce the specific surface area.

Figure 4 (a) Effect of calcination temperature on methane conversion; (b) Arrhenius plots for the oxidation of methane over the catalysts.
Figure 4

(a) Effect of calcination temperature on methane conversion; (b) Arrhenius plots for the oxidation of methane over the catalysts.

Table 4

Catalytic activities and Ea of the catalysts for CH4 combustion

SamplesT10/°CT50/°CT90/°CEa (kJ/mol)
Sample 1499659748112.24
Sample 2485637697111.56
Sample 3468623686108.94
Sample 443558666397.02
Sample 5606702737136.77

According to literature, methane oxidation has a first-order kinetic model (with a methane concentration of less than 20%) [6,10]. The apparent activation energy (Ea) of samples 1, 4, and 5 was calculated, and the correlation coefficients of the obtained curves were all above 0.989, showing strong linearity. As shown in Figure 4b and Table 4, the Ea for sample 4 is 97.02 kJ/mol, which is lower than that of sample 1 (112.24 kJ/mol) and sample 5 (136.77 kJ/mol). Thus, sample 4 has the highest catalytic activity for CH4 combustion.

3.3 XRD characterization analysis

XRD patterns of the original mine tailings and the subsequent catalysts calcined at different temperatures are shown in Figure 5a and b. The mineral composition of the mine tailings is extremely complex. According to literature, these REE-containing mine tailings contain 71 elements and 172 minerals [18]. The main components are iron mineral and fluorocarbons, of which the primary constituent is the fluorocarbon cerium ore.

Figure 5 XRD patterns of REE-rich tailings: (a) original mine tailings and (b) mine tailings calcined at different temperatures.
Figure 5

XRD patterns of REE-rich tailings: (a) original mine tailings and (b) mine tailings calcined at different temperatures.

Figure 5a shows that CaF2, Fe2O3, and CeCO3F are the main minerals in the mine tailings before calcination. The intensity of the diffraction peak of CaF2 is the strongest, indicating that the CaF2 content is high. The intensity of the CeCO3F peak gradually decreases with increasing calcination temperature (Figure 5b). At the calcination temperature of 700°C, the diffraction peak of the fluorocarbon cerium ore disappears. According to literature, the fluorocarbon cerium ore is completely decomposed into CeO2 at 700°C [21]. Furthermore, Fe3O4 in the mine tailings is converted into Fe2O3. Both CeO2 and Fe2O3 can promote the catalytic combustion of methane. However, the conversion rate of methane decreases after calcination at 700°C.

3.4 Properties and microstructure analysis

As shown in the SEM results (Figure 6a), the original mine tailings have an irregular shape but smooth surface. Post calcination at 500°C, cracks appear on the surface of the mineral (Figure 6b). Post calcination at 600°C, further cracks are observed on the mineral, with the formation of a pore structure that increases the mineral’s specific surface area. At 700°C, the entire surface of the tailings is flocculent. In conclusion, surface roughness increases with the calcination temperature, resulting in a flocculated surface. This effect increases the specific surface area of the catalysts and exposes active sites, increasing the contact area between methane gas and the catalyst [22,23].

Figure 6 SEM characterization of rare earth tailings before and after calcination: (a) original mine tailings, (b) tailings calcined at 500°C, (c) tailings calcined at 600°C, (d) tailings calcined at 700°C.
Figure 6

SEM characterization of rare earth tailings before and after calcination: (a) original mine tailings, (b) tailings calcined at 500°C, (c) tailings calcined at 600°C, (d) tailings calcined at 700°C.

As shown in Figure 7, energy-dispersive X-ray spectroscopy (EDS) mapping of the catalyst calcined at 600°C reveals that Fe, Ce, Ti, O, and other active metal elements are evenly distributed on the catalyst surface.

Figure 7 EDS mapping of tailings calcined at 600°C.
Figure 7

EDS mapping of tailings calcined at 600°C.

3.5 XPS and oxygen species analysis

Figure 8a shows the Ce 3d XPS spectra for samples 1–5. Cerium mainly exists as Ce3+ and Ce4+ in the samples. The characteristic peaks of Ce4+ approximately appear at the following binding energies: 882.20, 888.60, 898.00, 907.20, and 916.15 eV, labeled as v0, v1, v2, v3, v4, and v5, respectively. The characteristic peaks of Ce3+ approximately appear at the following binding energies: 884.40, 880.00, 903.90, and 899.30 eV, labeled as u0, u1, u2, and u3, respectively [24,25,26]. The spectra for the mine tailings contain the characteristic peaks of both Ce3+ and Ce4+ before and after calcination. The conversion of Ce3+ to Ce4+ occurs in the calcined samples. Ce3+ is beneficial owing to the production of oxygen vacancies and unsaturated chemical bonds, which can promote the adsorption and migration of oxygen and thereby improve the redox performance of the catalyst. The relative contents of Ce3+/Ce4+ are presented in Table 5. Sample 4 contained the highest Ce3+ content, which confirms the above viewpoint. Therefore, the catalyst calcined at 600°C exhibits better catalytic performance owing to the presence of both Ce3+ and Ce4+.

Figure 8 XPS spectra of catalysts: (a) Ce 3d, (b) Fe 2p, (c) O 1 s.
Figure 8

XPS spectra of catalysts: (a) Ce 3d, (b) Fe 2p, (c) O 1 s.

Table 5

Surface compositions and oxidation states of the samples.

Ce3dFe2pO1s
Ce3+Ce4+Ce3+/Ce4+Fe2+Fe3+Fe2+/Fe3+OadsOlattOads/Olatt
Sample 10.520.481.080.210.790.300.820.184.56
Sample 20.210.790.270.090.910.100.880.127.33
Sample 30.230.770.300.170.830.200.920.0811.50
Sample 40.24 0.760.310.360.640.560.950.0519.00
Sample 50.220.780.280.190.810.200.800.204.00

Figure 8b shows the Fe 2p XPS spectra for samples 1–5. The binding energies of the characteristic peaks of Fe3+ are observed at approximately 711 (W1) and 725 eV (W2). The binding energy for Fe2+ is at approximately 718–721 eV [27]. From the spectrum, we can observe that Fe exists in the form of both Fe2+ and Fe3+. Interconversion between Fe2+ and Fe3+ provides unstable oxygen vacancies and increases mobility of the lattice oxygen species on the catalyst surface. Fe2+ is easily reduced; thus, Fe2+ is beneficial to the catalytic performance of the catalyst. Table 5 shows that the tailings calcined at 600°C contain more Fe2+. The Fe2+/Fe3+ ratio in sample 4 was the highest among the five samples and reached 0.56, resulting in good catalytic performance.

Figure 8c shows the O 1s XPS spectra for samples 1–5. Two species of oxygen can be distinguished here. The characteristic peak for lattice oxygen (Olatt) can be found at approximately 529.5–530.5 eV, and the characteristic peak for adsorbed energy (Oads) can be found at approximately 531.5–533.0 eV [28]. According to the literature [29,30], lattice oxygen and adsorbed oxygen can be interconverted, whereby the reaction with adsorbed molecules on the catalyst surface can be achieved by converting lattice oxygen to adsorbed oxygen. The Oads/Olatt ratio can be used to describe the ratio of oxygen species: the higher the ratio, the more reactive oxygen species are present on the catalyst surface and the better is the activity. Table 5 shows a range of Oads/Olatt values for the samples in the following order: sample 4 (19.00) > sample 3 (11.50) > sample 2 (7.33) > sample 1 (4.56) > sample 5 (4.00). Therefore, the high content of Ce3+, Fe2+, and Oads contributes to the improvement of catalytic activity.

3.6 Reducibility and stability of catalyst

High mobility of oxygen in surface and bulk contributes to the high activity for hydrocarbon oxidation. H2-TPR diagram shows (Figure 9) that sample 1 has only one hydrogen consumption peak at 665°C, samples 2 and 3 have two reduction peaks, and sample 4 has four reduction peaks. The first reduction peak of samples 2, 3, and 4 appears at approximately 400°C, and the first reduction peak of sample 4 appears at approximately 500°C; these peaks are due to the surface-active oxygen reduction of the Ce4+–O–Ce4+ bond. In samples 1–4, a reduction peak appears at approximately 650°C, which can be attributed to FeO [31,32,33]. The reduction peaks of samples 4 and 5 at approximately 650–800°C can be attributed to the surface lattice oxygen [34]. Among them, sample 4 has the largest number of reduction peaks, which are at low temperatures, revealing that the redox properties of the catalyst calcined at 600°C are relatively good, thus improving the corresponding catalytic performance.

Figure 9 H2-TPR diagrams before and after the calcination of mine tailings.
Figure 9

H2-TPR diagrams before and after the calcination of mine tailings.

Figure 10 shows the stability experiment results for samples 1, 3, and 4. The samples were evaluated on-stream for 3,500 min at T90 (686°C). The methane conversion of sample 1 began to decline significantly around 260 min and consistently declined until 3,500 min. The methane conversion dropped from 63% to 31%. However, the methane conversion for samples 3 and 4 only started to gradually decline after 1,000 min. After 3500 min, the methane conversion rate remained above 80%. Therefore, a higher calcination temperature can increase the stability of the catalyst.

Figure 10 Methane combustion with time-on-stream for samples 1, 3, and 4.
Figure 10

Methane combustion with time-on-stream for samples 1, 3, and 4.

4 Conclusion

Mine tailings from an REE mine were used to prepare catalysts for the catalytic combustion of low-concentration CH4. The mine tailings were calcined at 400°C, 500°C, 600°C, and 700°C to prepare the catalysts. Calcination at 600°C produced catalysts that exhibited the lowest apparent activation energy (Ea is 97.02 kJ/mol) along with the highest performance for the catalytic combustion of methane (T50 of 586°C and T90 of 663°C). To further investigate the structural changes post calcination, various characterization methods were performed. The diffraction peak intensity of Fe2O3 in the mine tailings was enhanced, whereas the CeCO3F peak intensities were reduced. Moreover, a series of beneficial properties were afforded upon the calcination of the catalyst. Cracks and holes appeared on the surface, which increased the specific surface area. Fe, Ce, Ti, O, and other active metal elements were evenly distributed throughout the catalyst. A high surface Oads/Olatt ratio was achieved, thus enhancing the reducibility. Ce coexisted in both Ce3+ and Ce4+ oxidation states, and Fe coexisted in the Fe2+ and Fe3+ oxidation states as active components in the mine tailings, which improved the catalytic activity. However, at 700°C, sintering occurred, which lowered the catalytic performance.


tel: +86-04-72-595-1507, fax: +86-04-72-595-1507

Acknowledgments

The investigation was financially supported by Inner Mongolia Science and Technology Plan (2017CXYD-3 and 0901051701), open project for key basic research of the Inner Mongolia Autonomous Region (20140201), and Inner Mongolia Major Basic Research Open Project (0406091701).

References

[1] Lu SM. A global survey of gas hydrate development and reserves: specifically in themarine field. Renew Sustain Energy Rev. 2015;41:884–900.10.1016/j.rser.2014.08.063Search in Google Scholar

[2] Toso A, Colussi S, Padigapaty S, De Leitenburg C, Trovarelli A. High stability and activity of solution combustion synthesized Pd-based catalysts for methane combustion in presence of water. Appl Catal B. 2018;230(15):237–45.10.1016/j.apcatb.2018.02.049Search in Google Scholar

[3] Hong E, Kim C, Lim DH, Cho HJ, Shin CH. Catalytic methane combustion over Pd/ZrO2 catalysts: effects of crystalline structure and textural properties. Appl Catal B. 2018;232:544–52.10.1016/j.apcatb.2018.03.101Search in Google Scholar

[4] Hu RS, Bai YQ, Du YH, Zhang HM, Du YF, Zhang JG, et al. Surface structure and catalytic performance of Sr-doped La2NiAlO6 double perovskite catalysts for methane combustion. J Rare Earths. 2015;33(12):1284–92.10.1016/S1002-0721(14)60558-5Search in Google Scholar

[5] Shi LM, Chu W, Qu FF, Luo SZ. Low-temperature catalytic combustion of methane over MnOx–CeO2 mixed oxide catalysts: effect of preparation method. Catal Lett. 2007;113(1–2):59–64.10.1007/s10562-006-9012-6Search in Google Scholar

[6] Li JJ, Hu RS, Zhang JG, Meng WW, Du YF, Si YF, et al. Influence of preparation methods of La2CoMnO6/CeO2 on the methane catalytic combustion. Sustain Energy Fuels. 2016;178:148–54.10.1016/j.fuel.2016.03.041Search in Google Scholar

[7] Yang XW, Du CH, Guo YL, Guo Y, Wang L, Wang YS, et al. and Al2O3 supported hybrid Pd-CeO2 colloidal spheres and its enhanced catalytic performances for methane combustion. J Rare Earths. 2019;37(07):714–9.10.1016/j.jre.2018.10.005Search in Google Scholar

[8] Jia L, Chen YL, Liu XJ, Chen XH, Zheng Y, Huang F, et al. Microstructural property regulation and performance in methane combustion reaction of ordered mesoporous alumina supported palladium-cobalt bimetallic catalysts. Appl Catal B. 2020;263:118–269.Search in Google Scholar

[9] Xu P, Wu ZX, Deng JG, Liu YX, Xie SH, Guo GS, et al. Catalytic performance enhancement by alloying Pd with Pt on ordered mesoporous manganese oxide for methane combustion. Chin J Catal. 2017;38(1):92–105.10.1016/S1872-2067(16)62567-6Search in Google Scholar

[10] Fan XH, Li LM, Jing FL, Li J, Chu W. Effects of preparation methods on CoAlOx/CeO2 catalysts for methane catalytic combustion. Energy Wkly N. 2019;225:588–95.10.1016/j.fuel.2018.03.006Search in Google Scholar

[11] Li DY, Li KZ, Xu RD, Wang H, Tian D, Wei YG, et al. Ce1-xFexO2−-δ catalysts for catalytic methane combustion: role of oxygen vacancy and structural dependence. Catal Today. 2018;318(15):73–85.10.1016/j.cattod.2017.12.015Search in Google Scholar

[12] Hu Z, Qiu S, You Y, Guo Y, Guo YL, Wang L, et al. Hydrothermal synthesis of NiCeOx nanosheets and its application to the total oxidation of propane. Appl Catal B. 2018;225(5):110–20.10.1016/j.apcatb.2017.08.068Search in Google Scholar

[13] Zhang YG, Qin ZF, Wang GF, Zhu HQ, Dong M, Li SN, et al. Catalytic performance of MnOx–NiO composite oxide in lean methane combustion at low temperature. Appl Catal B. 2013;129(17):172–81.10.1016/j.apcatb.2012.09.021Search in Google Scholar

[14] Zhang JJ, Li JW, Zhu JQ, Wang Y, Chen BH. Effect of promoter on the performance of Cu–Mn complex oxide monolithic catalysts for lean methane catalytic combustion. Chin J Catal. 2011;32(8):1380–6.10.3724/SP.J.1088.2011.10323Search in Google Scholar

[15] Zhang B, Liu CJ, Li CL, Jiang MF. A novel approach for recovery of rare earths and niobium from Bayan Obo tailings. Min Eng. 2014;65(15):17–23.10.1016/j.mineng.2014.04.011Search in Google Scholar

[16] Ma PQ, Gao YS, Xu LZ. Comprehensive utilization and environmental protection of Bayan Obo resources in Baotou. Decis Mak Consult (Chin). 2009;2:88–91.Search in Google Scholar

[17] Zheng Q, Wu WY, Bian X. Investigations on mineralogical characteristics of rare earth minerals in Bayan Obo tailings during the roasting process. J Rare Earths. 2017;35(3):300–8.10.1016/S1002-0721(17)60913-XSearch in Google Scholar

[18] Lin G, Zhang LiB, Yin SH, Peng JH, Li SW, Xie F. Study on the calcination experiments of rare earth carbonates using microwave heating. Green Process Synth. 2015;4(4):329–36.10.1515/gps-2015-0040Search in Google Scholar

[19] Sun SC, Wu ZY, Gao B, Bian X, Wu WY, Tu GF. Effect of CaO on fluorine in the decomposition of REFCO3. J Rare Earths. 2007;25(4):508–11.10.1016/S1002-0721(07)60465-7Search in Google Scholar

[20] Zhang SR, Tu GF, Ren CZ, Zhang CX, Li CC. Study on the decomposition behavior of bastnasite. ChJ Rare Met (Chin). 1998;22(3):26–8.Search in Google Scholar

[21] Tu GF, Zhang SR, Ren CZ, Xing CZ, Zhang CX. Thermal decomposition reaction kinetics model of powdered bastnaesite. J Rare Earths (Chin). 2000;18(04):265–7.Search in Google Scholar

[22] Galaai H, Pijolat M, Nahdi K, Trabelsi-Ayadi M. Mechanism of growth of MgO and CaCO3 during a dolomite partial decomposition. Solid State Ion. 2007;178(15–18):1039–47.10.1016/j.ssi.2007.05.013Search in Google Scholar

[23] Wang ZF, Huang YN, Luo HJ, Gong ZJ, Zhang K, Li N, et al. Denitrification performance of rare earth tailings-based catalysts. Green Process Synth. 2019;8(1):865–72.10.1515/gps-2019-0057Search in Google Scholar

[24] Alexandrou M, Nix RM. The growth, structure and stability of ceria over layers on Pd (111). vol. 321, North-Holland; 1994. p. 47–57.10.1016/0039-6028(94)90025-6Search in Google Scholar

[25] Burroughs P, Hamnett A, Orchard AF, Thornton G. Satellite structure in the X-ray photoelectron spectra of some binary and mixed oxides of lanthanum and cerium. Chem Soc Dalton Trans. 1976;17:1686–98.10.1039/dt9760001686Search in Google Scholar

[26] Wang Z, Shen GL, Li JQ, Liu HD, Wang Q, Chen YF. Catalytic removal of benzene over CeO2-MnOx composite oxides prepared by hydrothermal method. Appl Catal B. 2013;138:253–9.10.1016/j.apcatb.2013.02.030Search in Google Scholar

[27] Liu FD, He H, Ding Y, Zhang CB. Effect of manganese substitution on the structure ang activity of iron titanate catalyst for the selective catalytic reduction of NO with NH3. Appl Catal B. 2009;93(1–2):194–204.10.1016/j.apcatb.2009.09.029Search in Google Scholar

[28] Xia YS, Dai HX, Zhang L, Deng JG, He H, Au CT. Ultrasound-assisted nanocasting fabrication and excellent catalytic performance of three-dimensionally ordered mesoporous chromia for the combustion of formaldehyde, acetone, and methanol. Appl Catal B. 2010;100:229–37.10.1016/j.apcatb.2010.07.037Search in Google Scholar

[29] Wang B, Weng D, Wu XD, Ran R. Modification of Pd-CeO2 catalyst by different treatments: effect on the structure and CO oxidation activity. Appl Surf Sci. 2011;257:3878–83.10.1016/j.apsusc.2010.11.083Search in Google Scholar

[30] Hu Z, Liu XF, Meng DM, Guo Y, Guo YL, Lu GZ. Effect of ceria crystal plane on the physicochemical and catalytic properties of Pd/ceria for CO and propane oxidation. ACS Catal. 2016;6:2265.10.1021/acscatal.5b02617Search in Google Scholar

[31] Shi XY, He H, Xie LJ. The effect of Fe species distribution and acidity of Fe-ZSM-5 on thehydrothermal stability and SO2 and hydrocarbons durability in NH3-SCR reaction. Chinese. J Catal. 2015;36:649–56.10.1016/S1872-2067(14)60268-0Search in Google Scholar

[32] Brandenberger S, Krocher O, Casapu M, Tissler A, Althoff R. Hydrothermal deactivation of Fe-ZSM-5 catalysts for the selective catalytic reduction of NO with NH3. Appl Catal B. 2011;101(3–4):649–59.10.1016/j.apcatb.2010.11.006Search in Google Scholar

[33] Chu W, Chernavskii P, Gengembre L, Pankina G, Fongarland P, Khodakov A. Cobalt species in promoted cobalt alumina-supported Fischer-Tropsch catalysts. J Catal. 2007;252(2):215–30.10.1016/j.jcat.2007.09.018Search in Google Scholar

[34] Jiang BQ, Liu Y, Wu ZB. Low-temperature selective catalytic reduction of NO on MnOx/TiO2 prepared by different methods. J Hazard Mater. 2009;162:1249–54.10.1016/j.jhazmat.2008.06.013Search in Google Scholar PubMed

Received: 2020-06-27
Revised: 2020-08-06
Accepted: 2020-08-09
Published Online: 2020-12-02

© 2020 Ran Zhao et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Obituary for Prof. Dr. Jun-ichi Yoshida
  2. Regular Articles
  3. Optimization of microwave-assisted manganese leaching from electrolyte manganese residue
  4. Crustacean shell bio-refining to chitin by natural deep eutectic solvents
  5. The kinetics of the extraction of caffeine from guarana seed under the action of ultrasonic field with simultaneous cooling
  6. Biocomposite scaffold preparation from hydroxyapatite extracted from waste bovine bone
  7. A simple room temperature-static bioreactor for effective synthesis of hexyl acetate
  8. Biofabrication of zinc oxide nanoparticles, characterization and cytotoxicity against pediatric leukemia cell lines
  9. Efficient synthesis of palladium nanoparticles using guar gum as stabilizer and their applications as catalyst in reduction reactions and degradation of azo dyes
  10. Isolation of biosurfactant producing bacteria from Potwar oil fields: Effect of non-fossil fuel based carbon sources
  11. Green synthesis, characterization and photocatalytic applications of silver nanoparticles using Diospyros lotus
  12. Dielectric properties and microwave heating behavior of neutral leaching residues from zinc metallurgy in the microwave field
  13. Green synthesis and stabilization of silver nanoparticles using Lysimachia foenum-graecum Hance extract and their antibacterial activity
  14. Microwave-induced heating behavior of Y-TZP ceramics under multiphysics system
  15. Synthesis and catalytic properties of nickel salts of Keggin-type heteropolyacids embedded metal-organic framework hybrid nanocatalyst
  16. Preparation and properties of hydrogel based on sawdust cellulose for environmentally friendly slow release fertilizers
  17. Structural characterization, antioxidant and cytotoxic effects of iron nanoparticles synthesized using Asphodelus aestivus Brot. aqueous extract
  18. Phase transformation involved in the reduction process of magnesium oxide in calcined dolomite by ferrosilicon with additive of aluminum
  19. Green synthesis of TiO2 nanoparticles from Syzygium cumini extract for photo-catalytic removal of lead (Pb) in explosive industrial wastewater
  20. The study on the influence of oxidation degree and temperature on the viscosity of biodiesel
  21. Prepare a catalyst consist of rare earth minerals to denitrate via NH3-SCR
  22. Bacterial nanobiotic potential
  23. Green synthesis and characterization of carboxymethyl guar gum: Application in textile printing technology
  24. Potential of adsorbents from agricultural wastes as alternative fillers in mixed matrix membrane for gas separation: A review
  25. Bactericidal and cytotoxic properties of green synthesized nanosilver using Rosmarinus officinalis leaves
  26. Synthesis of biomass-supported CuNi zero-valent nanoparticles through wetness co-impregnation method for the removal of carcinogenic dyes and nitroarene
  27. Synthesis of 2,2′-dibenzoylaminodiphenyl disulfide based on Aspen Plus simulation and the development of green synthesis processes
  28. Catalytic performance of the biosynthesized AgNps from Bistorta amplexicaule: antifungal, bactericidal, and reduction of carcinogenic 4-nitrophenol
  29. Optical and antimicrobial properties of silver nanoparticles synthesized via green route using honey
  30. Adsorption of l-α-glycerophosphocholine on ion-exchange resin: Equilibrium, kinetic, and thermodynamic studies
  31. Microwave-assisted green synthesis of silver nanoparticles using dried extracts of Chlorella vulgaris and antibacterial activity studies
  32. Preparation of graphene oxide/chitosan complex and its adsorption properties for heavy metal ions
  33. Green synthesis of metal and metal oxide nanoparticles from plant leaf extracts and their applications: A review
  34. Synthesis, characterization, and electrochemical properties of carbon nanotubes used as cathode materials for Al–air batteries from a renewable source of water hyacinth
  35. Optimization of medium–low-grade phosphorus rock carbothermal reduction process by response surface methodology
  36. The study of rod-shaped TiO2 composite material in the protection of stone cultural relics
  37. Eco-friendly synthesis of AuNPs for cutaneous wound-healing applications in nursing care after surgery
  38. Green approach in fabrication of photocatalytic, antimicrobial, and antioxidant zinc oxide nanoparticles – hydrothermal synthesis using clove hydroalcoholic extract and optimization of the process
  39. Green synthesis: Proposed mechanism and factors influencing the synthesis of platinum nanoparticles
  40. Green synthesis of 3-(1-naphthyl), 4-methyl-3-(1-naphthyl) coumarins and 3-phenylcoumarins using dual-frequency ultrasonication
  41. Optimization for removal efficiency of fluoride using La(iii)–Al(iii)-activated carbon modified by chemical route
  42. In vitro biological activity of Hydroclathrus clathratus and its use as an extracellular bioreductant for silver nanoparticle formation
  43. Evaluation of saponin-rich/poor leaf extract-mediated silver nanoparticles and their antifungal capacity
  44. Propylene carbonate synthesis from propylene oxide and CO2 over Ga-Silicate-1 catalyst
  45. Environmentally benevolent synthesis and characterization of silver nanoparticles using Olea ferruginea Royle for antibacterial and antioxidant activities
  46. Eco-synthesis and characterization of titanium nanoparticles: Testing its cytotoxicity and antibacterial effects
  47. A novel biofabrication of gold nanoparticles using Erythrina senegalensis leaf extract and their ameliorative effect on mycoplasmal pneumonia for treating lung infection in nursing care
  48. Phytosynthesis of selenium nanoparticles using the costus extract for bactericidal application against foodborne pathogens
  49. Temperature effects on electrospun chitosan nanofibers
  50. An electrochemical method to investigate the effects of compound composition on gold dissolution in thiosulfate solution
  51. Trillium govanianum Wall. Ex. Royle rhizomes extract-medicated silver nanoparticles and their antimicrobial activity
  52. In vitro bactericidal, antidiabetic, cytotoxic, anticoagulant, and hemolytic effect of green-synthesized silver nanoparticles using Allium sativum clove extract incubated at various temperatures
  53. The green synthesis of N-hydroxyethyl-substituted 1,2,3,4-tetrahydroquinolines with acidic ionic liquid as catalyst
  54. Effect of KMnO4 on catalytic combustion performance of semi-coke
  55. Removal of Congo red and malachite green from aqueous solution using heterogeneous Ag/ZnCo-ZIF catalyst in the presence of hydrogen peroxide
  56. Nucleotide-based green synthesis of lanthanide coordination polymers for tunable white-light emission
  57. Determination of life cycle GHG emission factor for paper products of Vietnam
  58. Parabolic trough solar collectors: A general overview of technology, industrial applications, energy market, modeling, and standards
  59. Structural characteristics of plant cell wall elucidated by solution-state 2D NMR spectroscopy with an optimized procedure
  60. Sustainable utilization of a converter slagging agent prepared by converter precipitator dust and oxide scale
  61. Efficacy of chitosan silver nanoparticles from shrimp-shell wastes against major mosquito vectors of public health importance
  62. Effectiveness of six different methods in green synthesis of selenium nanoparticles using propolis extract: Screening and characterization
  63. Characterizations and analysis of the antioxidant, antimicrobial, and dye reduction ability of green synthesized silver nanoparticles
  64. Foliar applications of bio-fabricated selenium nanoparticles to improve the growth of wheat plants under drought stress
  65. Green synthesis of silver nanoparticles from Valeriana jatamansi shoots extract and its antimicrobial activity
  66. Characterization and biological activities of synthesized zinc oxide nanoparticles using the extract of Acantholimon serotinum
  67. Effect of calcination temperature on rare earth tailing catalysts for catalytic methane combustion
  68. Enhanced diuretic action of furosemide by complexation with β-cyclodextrin in the presence of sodium lauryl sulfate
  69. Development of chitosan/agar-silver nanoparticles-coated paper for antibacterial application
  70. Preparation, characterization, and catalytic performance of Pd–Ni/AC bimetallic nano-catalysts
  71. Acid red G dye removal from aqueous solutions by porous ceramsite produced from solid wastes: Batch and fixed-bed studies
  72. Review Articles
  73. Recent advances in the catalytic applications of GO/rGO for green organic synthesis
Downloaded on 1.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2020-0053/html
Scroll to top button