RESEARCH ARTICLE

Catalytic combustion of methane over a highly active and stable NiO/CeO2 catalyst

  • Xiuhui Huang , 1,2 ,
  • Junfeng Li 3 ,
  • Jun Wang 1,2 ,
  • Zeqiu Li 1,2 ,
  • Jiayin Xu , 1,2
Expand
  • 1. School of Energy and Power Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
  • 2. Shanghai Key Laboratory of Multiphase Flow and Heat Transfer in Power Engineering, Shanghai 200093, China
  • 3. Shanghai MCC20 Construction Co. Ltd., Shanghai 201999, China

Received date: 28 Sep 2018

Accepted date: 05 Feb 2019

Published date: 15 Aug 2020

Copyright

2020 Higher Education Press

Abstract

In the last decades, many reports dealing with technology for the catalytic combustion of methane (CH4) have been published. Recently, attention has increasingly focused on the synthesis and catalytic activity of nickel oxides. In this paper, a NiO/CeO2 catalyst with high catalytic performance in methane combustion was synthesized via a facile impregnation method, and its catalytic activity, stability, and water-resistance during CH4 combustion were investigated. X-ray diffraction, low-temperature N2 adsorption, thermogravimetric analysis, Fourier transform infrared spectroscopy, hydrogen temperature programmed reduction, methane temperature programmed surface reaction, Raman spectroscopy, electron paramagnetic resonance, and transmission electron microscope characterization of the catalyst were conducted to determine the origin of its high catalytic activity and stability in detail. The incorporation of NiO was found to enhance the concentration of oxygen vacancies, as well as the activity and amount of surface oxygen. As a result, the mobility of bulk oxygen in CeO2 was increased. The presence of CeO2 prevented the aggregation of NiO, enhanced reduction by NiO, and provided more oxygen species for the combustion of CH4. The results of a kinetics study indicated that the reaction order was about 1.07 for CH4 and about 0.10 for O2 over the NiO/CeO2 catalyst.

Cite this article

Xiuhui Huang , Junfeng Li , Jun Wang , Zeqiu Li , Jiayin Xu . Catalytic combustion of methane over a highly active and stable NiO/CeO2 catalyst[J]. Frontiers of Chemical Science and Engineering, 2020 , 14(4) : 534 -545 . DOI: 10.1007/s11705-019-1821-4

Introduction

In recent decades, a great deal of research into technology for the catalytic combustion of methane (CH4) has been conducted because of the high efficiency, safety, and low emissions (NOx, CO, etc.) of this technology [13]. CH4 is the most inert hydrocarbon, and the synthesis of highly active and stable catalysts for the catalytic combustion of CH4 remains challenging. Noble metal-based catalysts, especially supported Pt and Pd catalysts, have been extensively investigated for CH4 combustion [46]. For instance, 100% conversion of CH4 over Pd-NiCo2O4/SiO2 was achieved at 378°C [7]. However, the use of Pt and Pd catalysts in industrial applications is limited by their high cost, low resistance against poisoning, and decreased catalytic activity at high temperature. Transition metal-based catalysts (hexaaluminates, perovskites, and transition metal composite oxides) have also been widely investigated owing to their fairly good stability and low cost [810]. Hexaaluminate catalysts are very stable during the combustion of methane, but their catalytic activity requires improvement. Similarly, perovskites exhibit desirable catalytic performance, but their Brunner-Emmett-Teller (BET) surface area decreases greatly at high temperatures, which negatively affects their activity [11]. Cobalt oxide (generally Co3O4) catalysts have shown high catalytic activity in CH4 combustion as cobalt is the one of the most active transition metal elements [1214]. High conversion (90%) [12] can be achieved at a low temperature of 345°C; this performance is similar to that of noble metal catalysts. Unfortunately, cobalt oxide catalysts are sensitive to moisture and easily deactivated. Compared to cobalt oxides [15,16], CuO-based catalysts exhibit higher stability and better water resistance in CH4 combustion, but their catalytic performance is much lower. Therefore, it is necessary to develop a new, thermally stable low-cost catalyst for the combustion of methane.
Recently, attention has increasingly focused on the synthesis and catalytic activity of nickel oxides [1719]. For example, H2 has been produced by ethanol steam reforming over NiO/CeO2 with a selectivity value of 61.5 mol-% and an ethanol conversion of 95.0 mol-% [18]. However, NiO-based catalysts show low catalytic activity in the reforming of methane [2022]. To enhance its catalytic activity in CH4 combustion, NiO must be loaded onto an appropriate support. CeO2 exhibits unique redox properties and a high oxygen storage capacity (OSC) [23,24]. Thus, CeO2-supported NiO catalysts have been considered for CH4 combustion [2527]. The addition of CeO2 to NiO has been observed to notably enhance its N2O decomposition activity [26]. However, to the best of our knowledge, almost no research into the thermal stability and water resistance of NiO-based catalysts has been reported. The mechanism for the improvement performance of NiO-based catalysts was also unclear. In this study, an effective NiO/CeO2 catalyst was prepared using a facile impregnation method, and the catalytic activity, stability, and water resistance of the catalyst during CH4 combustion were investigated. X-ray diffraction (XRD), low-temperature N2 adsorption, thermogravimetric analysis (TG), Fourier transform infrared spectroscopy (FT-IR), hydrogen temperature programmed reduction (H2-TPR), methane temperature programmed surface reaction (CH4-TPSR), Raman spectroscopy, electron paramagnetic resonance (EPR), and transmission electron microscopy (TEM) characterization were carried out in order to explain the increased catalytic activity and stability in detail.

Experimental

Catalyst preparation

Typically, the CeO2 support was obtained by the thermal decomposition method using Ce(NO3)3·6H2O (Sinopharm Chemical Reagent Co., Ltd. (SCRC), 99.0%) as a precursor at 450°C for 4 h in an air atmosphere [28]. The obtained CeO2 support was then impregnated with Ni(NO3)2 (SCRC, 99.0%) solutions of various concentrations to achieve NiO wt-% loadings from 1 to 20 wt-%. The Ni/CeO2 products were calcined at 450°C for an additional 4 h in an air atmosphere to fabricate NiO/CeO2 catalysts with 1, 3, 5, 7, 10, 15 and 20 wt-% NiO.
NiO/Al2O3 (10 wt-%) was synthesized using the same method as the 10 wt-% NiO/CeO2 catalyst (Al2O3 was purchased from BASF Co., Ltd.). The preparation methods for the SiO2 and KIT-6 supports were described in previous reports [29,30]. 10 wt-% NiO/SiO2 and 10 wt-% NiO/KIT-6 were obtained using the same impregnation method as for the NiO/CeO2 catalysts.

Catalyst characterization

XRD patterns of all samples were recorded using a Bruker D8 Focus diffractometer with CuKα radiation (l = 1.541 Å, operated at 40 kV and 40 mA). The N2 adsorption/desorption isotherms were measured at ‒196°C using a NOVA 4200e Surface Area and Pore Size Analyzer. Before the measurements, the samples were outgassed at 180°C under a vacuum for 4 h. The BET surface areas of the samples were calculated using the BET method, and the pore size distributions were calculated using the Barrett-Joyner-Halenda method from the adsorption branches of the isotherms. The total pore volumes (Vp) were estimated at a relative pressure of 0.985 (P/P0). The combined thermogravimetric analysis and differential thermal analysis (TG-DTA) runs were performed using a PerkinElmer Pyris Diamond with a WCT-2 thermal analyzer at a heating rate of 10°C·min1 from room temperature to 800°C. The Raman spectra were obtained using a Renishaw in Viat+ Reflex spectrometer equipped with a Charge Coupled Device detector at ambient temperature under moisture-free conditions. The emission line at 514.5 nm was produced by an Ar+ ion laser (Spectra Physics) and was focused on an analysis spot of about 1 mm on the sample under the microscope. FT-IR absorption spectra were recorded using a Nicolet NEXUS 670 FT-IR spectrometer, with 32 scans at an effective resolution of 4 cm1. The sample to be measured was ground with KBr and pressed into a thin wafer for analysis. H2-TPR was conducted using a thermal conductivity detector (TCD). The sample (50 mg) was loaded into a quartz tube reactor. A 5 vol-% H2/N2 gas mixture was used with a flow rate of 40 mL·min1, and the reactor was heated from room temperature to 850°C at a heating rate of 10°C·min1. CH4-TPSR of the samples was performed in a quartz micro-reactor. The catalyst (200 mg) was pretreated in the reactor at 500°C for 30 min under a 50 mL·min-1 flow of a 20 vol-% O2/He gas mixture. After the sample was cooled, it was then heated to 800°C in 50 mL·min1 of 1 vol-% CH4/He at a heating rate of 10°C·min-1. The outlet gas was analyzed using a quadrupole mass spectrometer (INFICON Transpecter 2). The signals of CH4, H2O, CO, O2, CO2 were recorded at mass-to-charge ratio (m/z) = 15, 18, 28, 32, and 44, respectively.

Catalytic activity tests and kinetic studies

The catalytic performance of the NiO/CeO2 catalysts for CH4 combustion was tested in a quartz tube reactor (ϕ 6 mm) at atmospheric pressure. The reactant gas mixture, which consisted of 1 vol-% CH4 + 4 vol-% O2 in Ar at a total flow rate of 50 mL·min1, was passed through the catalyst (200 mg) bed. The reactants and products were analyzed online using a gas chromatograph equipped with TCD. The activities of the catalysts were measured at programmed temperatures from 200°C to 700°C at a heating rate of 5°C·min1. The catalyst activity was characterized using the T10, T50, and T90 values, which represent the temperatures at which CH4 conversions of 10%, 50% and 90% were obtained, respectively.
In the measurement of kinetic parameters, the flow rate of the reactant gas was over 50 mL·min1, eliminating external diffusion. Internal diffusion was eliminated by pelleting the catalyst to 20‒40 mesh. During the measurement of the reaction orders, the reaction temperature was 320°C and the reactant gas consisted of 0.5‒4 vol-% CH4 and 2‒10 vol-% O2, balanced with Ar. During the measurement of the activation energy, the reactant gas consisted of 1 vol-% CH4 and 4 vol-% O2, balanced with Ar, and the weight hourly space velocity (WHSV) ranged from 15000 to 60000 mL·g1·h1. The conversion of CH4 was adjusted to below 15%. The reaction reached a steady state after 30 min, and the reaction rates were obtained.

Results and discussion

Effect of the support

In order to investigate the influence of the catalytic support on the activity of the NiO catalyst in CH4 combustion, SiO2, Al2O3 and KIT-6 supports with BET surface areas of 387, 208 and 708 m2·g1, respectively, were compared with the CeO2 support. Figure 1 shows the catalytic activities of the four catalysts, NiO/CeO2, NiO/SiO2, NiO/Al2O3 and NiO/KIT-6. All the catalysts contained 10 wt-% NiO. Although SiO2, Al2O3, and KIT-6 had much higher BET surface areas than CeO2, they exhibited lower catalytic performance. The T10 values of 10 wt-% NiO/SiO2, 10 wt-% NiO/Al2O3, and 10 wt-% NiO/KIT-6 were 407°C, 439°C and 389°C, respectively. That is, they were 74°C, 106°C and 56°C higher than that of the 10 wt-% NiO/CeO2 catalyst.
Fig.1 Effect of the supports on the activity of the supported NiO catalysts (gas space velocity= 60000 mL·g1·h1).

Full size|PPT slide

The XRD patterns of the four samples are displayed in Fig. 2(a). Diffraction peaks corresponding to cubic phase NiO were observed for all samples except the NiO/Al2O3 catalyst. Due to the high surface area of the supports, the NiO nanoparticles were better dispersed on the SiO2 and KIT-6 supports than on the CeO2 support. Only diffraction peaks corresponding to Al2O3 were observed on the NiO/Al2O3 catalyst due to the strong interaction between NiO and Al2O3. The reduction behaviors of the 10 wt-% NiO/CeO2, 10 wt-% NiO/SiO2, 10 wt-% NiO/Al2O3, and 10 wt-% NiO/KIT-6 samples are presented in Fig. 2(b). For the NiO/CeO2 sample, the peak centered at ~330°C was assigned to the reduction of NiO. The intensity of the reduction peak at 330°C was very low for the NiO/Al2O3 catalyst, and an additional strong reduction peak at 545°C was observed. The higher temperature reduction peak was assigned to the reduction of NiO due to its strong interaction with Al2O3 [31,32]. The reduction temperature of NiO in NiO/SiO2 was ~350°C. Two reduction peaks at 355°C and 500°C were observed on NiO/KIT-6, which corresponded to the reduction of NiO via weak and strong interactions with KIT-6, respectively. These results demonstrated that the intensity of the interaction between NiO and the support decreased in the order Al2O3>CeO2>KIT-6 ≈ SiO2. The different reduction properties of NiO depended on the intensity of the interaction between NiO and the support rather than the BET surface area or the NiO dispersion. The high activity of the NiO/CeO2 catalyst in CH4 combustion resulted from the appropriate intensity of the interaction between NiO and CeO2, which allowed for higher oxygen mobility and OSC than SiO2, Al2O3, or KIT-6 [33].
Fig.2 (a) XRD patterns and (b) H2-TPR profiles of the different NiO catalysts.

Full size|PPT slide

Effect of the NiO loading

The light-off curves of CH4 conversion over the CeO2-supported NiO samples with different NiO loadings are shown in Fig. 3.
Fig.3 Light-off curves of the catalytic combustion of methane over the NiO/CeO2 catalysts (gas space velocity= 60000 mL·g1·h1).

Full size|PPT slide

During the CH4 combustion reaction, CH4 was completely oxidized to CO2 and H2O (as determined by analysis using the INFICON Transpecter 2 quadruple mass spectrometer). Pure CeO2 had a low catalytic activity with high T10, T50, and T90 values of 456°C, 540°C and 615°C, respectively. However, the catalytic activity of NiO/CeO2 improved with increasing NiO content. The 10 wt-% NiO/CeO2 showed the highest catalytic activity, with low T10, T50, and T90 values of 333°C, 415°C, and 467°C, respectively. No improvement in the catalytic activity was observed when the NiO loading was further increased to 15% and 20%, as the NiO nanoparticles became aggregated and the number of reactive sites did not increase. As shown in Table 1, the increase in the reaction rate of CH4 with increasing NiO loading (from 1‒10 wt-%) was due to the increase in the number of NiO reactive sites. The CH4 reaction rate did not increase further for NiO loadings of over 10%, which was in agreement with the results of the light-off curves.
Tab.1 CH4 reaction rate over NiO/CeO2 catalysts at 360°C
Samples Cat./mg Flow rate/(mL·min‒1) Concentration of CH4 /vol-% Conversion of
CH4 /%
CH4 reaction rate
/(×10‒4 mmol·g‒1·s‒1)
1 wt-% NiO/CeO2 200 50 1 6.4 1.09
3 wt-% NiO/CeO2 200 50 1 10.4 1.77
5 wt-% NiO/CeO2 200 50 1 14.6 2.49
7 wt-% NiO/CeO2 200 100 1 9.0 3.07
10 wt-% NiO/CeO2 200 100 1 10.7 3.63
15 wt-% NiO/CeO2 200 100 1 10.3 3.51
20 wt-% NiO/CeO2 200 100 1 10.9 3.70

Effect of water vapor in the feed gas

In order to investigate the effect of water vapor on the catalytic performance of the synthesized NiO/CeO2 during CH4 combustion, 3.1% water vapor (25°C) was added to the feed gas; the catalytic performance of 10 wt-% NiO/CeO2 in the absence and presence of the water vapor is shown in Fig. 4. When 3.1% water vapor was added to the reaction stream, the catalytic performance of the 10 wt-% NiO/CeO2 catalyst declined slightly; T10, T50, and T90 increased to 355°C, 425°C and 480°C, respectively. It was concluded that the addition of water resulted in competition between water and CH4 molecules for adsorption on the active sites of the NiO/CeO2 catalyst, resulting in some of the active oxygen sites on the catalyst being blocked by water. The result also showed that CH4 was still totally oxidized to CO2 and H2O over the NiO/CeO2 catalyst without steam reforming reactions or partial oxidation to generate CO and H2, even in the presence of 3.1% water vapor. This finding was not in agreement with the previous work [34]. Therefore, the 10 wt-% NiO/CeO2 catalyst exhibited a degree of water resistance during CH4 combustion.
Fig.4 Effect of the addition of water vapor in CH4 combustion over 10 wt-% NiO/CeO2: (a) without H2O and (b) with 3.1% water vapor (25°C; gas space velocity= 60000 mL·g1·h1).

Full size|PPT slide

Determination of the kinetic parameters of CH4 oxidation

The empirical kinetic rate equation of CH4 combustion was as follows:
r=k PCH4αP O 2βP CO2γ PH2Oδ.
When the conversion of CH4 at 320°C was less than 10%,P CO2γand P H2O δwere assumed to be constant, allowing Eq. (1) to be simplified as follows:
r=k PCH4αPO2β.
By taking the logarithm of Eq. (2), Eq. (3) can be obtained:
lnr=αlnPC H4+β ln PO2+lnk.
Under the conditions of 0.5‒4 vol-% CH4 and 4% O2 at a WHSV of 30000 mL·g1·h1 and 320°C, Eq. (3) was simplified tolnr =α ln PCH4+C. A plot of ln r against lnP CH4for the 10 wt-% NiO/CeO2 catalyst at 320°C is shown in Fig. 5(a). The reaction order of CH4 (a) was equal to the slope of the resulting straight line, which was about 1.07. The reaction order of O2 (b) was determined to be about 0.10 by calculation. Thus, the reaction rates for CH4 and O2 were nearly first order and zero order, respectively (Figs. 5(a,b)). These findings were in agreement with those reported for a NiO/Ce0.75Zr0.25O2 catalyst and Pd-based catalyst [25,35].
Fig.5 Dependence of the reaction rates on the CH4 and O2 concentration over the 10 wt-% NiO/CeO2 catalyst at 320°C: (a) 0.5‒4 vol-% CH4, 4 vol-% O2, in Ar; (b) 2‒10 vol-% O2, 1 vol-% CH4, in Ar. (P CH4and PO2, partial pressures; gas space velocity= 30000 mL·g1·h1).

Full size|PPT slide

The Arrhenius equation is given below as Eq. (4):
k=Aexp( Ea RT).
Taking the logarithm of Eq. (4), Eq. (5) can be obtained:
lnr ln PCH4αln PO2β=lnA EaRT.
The components of the reactant feed gas did not change during the kinetic tests, and the conversion of CH4 at 320°C was<15%. Thus, ln PCH 4αand ln PO2β were assumed to be approximately constant, and Eq. (5) was simplified to lnr= Ea/(RT)+C. The plots of ln r against ln PCH4over the 10 wt-% NiO/CeO2 and CeO2 catalysts are shown in Fig. 6. The activation energy for 10 wt-% NiO/CeO2 was determined to be 61.7 kJ∙mol1, which was much lower than that of CeO2 (114.6 kJ∙mol1). Additionally, the activation energy for 10 wt-% NiO/CeO2 fell within the range of values reported for other transition oxides [36,37].
Fig.6 Arrhenius plots of the reaction rate of CH4 combustion (activation energy, Ea) over (a) pure CeO2 and 10 wt-% (b) NiO/CeO2 (gas space velocity= 30000 mL·g1·h1).

Full size|PPT slide

Catalyst characterization

XRD

The XRD patterns of the CeO2-supported NiO catalysts are shown in Fig. 7. The CeO2 fluorite structure (JCPDS file No. 34-0394) was observed in all the samples. Weak diffraction peaks of cubic phase NiO (JCPDS file No. 65-2901) were observed for the NiO/CeO2 catalysts with loadings of 7 wt-% or greater. The NiO particles were highly dispersed on the support surface, and thus NiO peaks could not be observed at the lower loading amounts (NiO<7 wt-%). When the concentration of NiO was greater than 7 wt-%, the NiO nanoparticles tended to aggregate. However, the diffraction peaks of the NiO crystal phase were still weak even in the XRD pattern of the 20 wt-% NiO/CeO2 catalyst, which indicated that CeO2 improved the dispersion of NiO. For all the NiO/CeO2 samples, there was no evident shift in diffraction peaks with CeO2, which revealed that most of the Ni2+ ions formed NiO nanoparticles dispersed on the CeO2 surface. Only a small amount of Ni2+ ions was incorporated into the CeO2 lattice by the impregnation method.
Fig.7 XRD patterns of the NiO/CeO2 samples with various NiO loadings.

Full size|PPT slide

N2 adsorption/desorption

The textural properties and catalytic activity of all the samples are shown in Table 2. No obvious changes in the crystallite size or pore volume of NiO/CeO2 were observed with different NiO loadings. CeO2 had a BET surface area of 76.4 m2·g1, and the BET surface area decreased with increasing NiO wt-%. The BET surface area of the 10 wt-% NiO/CeO2 catalyst was 58.1 m2·g1, which was lower than that of CeO2. Unexpectedly, the catalytic performance of NiO/CeO2 was much higher than that of CeO2. This suggested that the BET surface area was not the key factor in the catalytic performance in CH4 combustion.
Tab.2 Textural properties and catalytic activities of CeO2 and NiO/CeO2 samples for CH4 combustion
Samples Crystallite
size /(na·ma)
Surface area
/(m2·g‒1)
Pore volume
/(cm3·g‒1)
Catalytic activity /°C
T10 T50 T90
CeO2 8 76.4 0.206 456 540 615
1 wt-% NiO/CeO2 8.1 70.4 0.197 381 471 575
3 wt-% NiO/CeO2 8.3 68.8 0.195 362 442 516
5 wt-% NiO/CeO2 8.5 65.3 0.185 352 430 493
7 wt-% NiO/CeO2 8.8 61.1 0.183 349 424 487
10 wt-% NiO/CeO2 8.9 58.1 0.173 333 415 467
15 wt-% NiO/CeO2 8.9 57.2 0.171 350 416 472
20 wt-% NiO/CeO2 9 55.5 0.169 355 410 467
10 wt-% NiO/CeO2
(used)b
10.6 45.5 0.154

a) Crystallite sizes were calculated from the line broadenings of the (111) plane of CeO2 using the Scherrer equation; b) 10% NiO/CeO2 catalyst collected after reaction at 600°C for 150 h.

H2-TPR

Figure 8 shows the H2-TPR profiles of CeO2, NiO, and all the NiO/CeO2 catalysts. A weak reduction peak was observed at ~780°C for CeO2 and all the NiO/CeO2 samples, which was attributed to the reduction of bulk oxygen. Three hydrogen consumption peaks, a1, a2, and b, were observed for the NiO/CeO2 catalysts. When Ni2+ is incorporated into the lattice of CeO2 to replace some Ce4+ cations, charge unbalance and lattice distortion occur within the structure of CeO2 due to the smaller radius of Ni2+. Oxygen vacancies were then generated, which could easily adsorb oxygen to form oxygen species with high activity and reducibility [21]. Thus, the a1 and a2 peaks were ascribed to the reduction of surface-adsorbed oxygen [17]. As the NiO loading was increased from 3 wt-% to 20 wt-%, the areas of the a1 and a2 peaks did not change remarkably, but the reduction temperatures of a1 and a2 shifted to higher temperature.
Fig.8 H2-TPR profiles of the CeO2, NiO, and NiO/CeO2 samples.

Full size|PPT slide

The reduction peak centered at ~330°C (b peak) was assigned to the reduction of NiO. The temperature for the reduction of NiO in NiO/CeO2 was much lower than that of pure NiO (~416°C) due to the interaction between CeO2 and NiO, which facilitated the reduction of Ni2+. For the 1 wt-% NiO/CeO2 sample, only one main reduction peak at 248°C was observed; the reduction peak at ~330°C was absent. NiO was easily reduced in this sample because it was highly dispersed on the surface of CeO2.

Raman spectroscopy

The Raman spectra of CeO2 and the CeO2-supported NiO samples are shown in Fig. 9. The spectra were obtained using a 514.5 nm laser as the excitation source. As shown in Fig. 9, a strong band was observed at ~462 cm1 in the CeO2 spectrum; this band was assigned to the F2g Raman active mode of the fluorite structure of CeO2 [38]. No evident shifts of the band at 462 cm1 were observed for any of the NiO/CeO2 samples, which indicated that the NiO/CeO2 samples did not experience any evident lattice expansion or contraction in comparison with CeO2. All samples also exhibited another broad weak band at ~575 cm1 that was attributed to the oxygen vacancy [39,40] mentioned in the discussion of the H2-TPR results. The bands at ~620 cm-1 in the NiO/CeO2 samples were ascribed to NiO species [41], which suggested that highly dispersed NiO was present on the surface of the samples. The areas of the bands at 462, 575, and 620 cm1 were denoted as A462, A575, and A620, respectively. The A575/A462 ratio in Raman spectra of CeO2-based catalysts is typically used to determine the concentration of oxygen vacancies, while A620/A462 represents the concentration of highly dispersed NiO [42]. Increasing the oxygen mobility of CeO2-based catalysts by introducing defective sites is known to be effective in promoting hydrocarbon oxidation reactions [43]. Moreover, highly dispersed NiO exhibits high catalytic activity in CH4 combustion due to its good redox properties [17]. To investigate the influence of the concentrations of oxygen vacancies and highly dispersed NiO on the catalytic activity in greater detail, the correlations between the special activity (CH4 conversion per unit surface area) at 460°C and the (A575+A620)/A462 ratio are shown in Fig. 10. With increasing (A575+A620)/A462 ratio, the concentration of oxygen vacancies and highly dispersed NiO increased, and the special activity of the catalyst also increased.
Fig.9 Raman spectra of NiO/CeO2 with NiO loadings of (a) 0, (b) 1 wt-%, (c) 3 wt-%, (d) 5 wt-%, (e) 7 wt-%, (f) 10 wt-%, (g) 15 wt-%, and (h) 20 wt-%. Inset: Raman spectrum of 10 wt-% NiO/CeO2.

Full size|PPT slide

Fig.10 Special conversion of CH4 per unit surface area at 460°C vs. (A575+A620)/A462 in the Raman spectra of CeO2-supported NiO with NiO loadings of 0, 1, 3, 5, 7, and 10 wt-% (gas space velocity= 60000 mL·g1·h1).

Full size|PPT slide

CH4-TPSR

To determine the origin of the high performance of the NiO-based catalysts, CH4-TPSR was also performed, and the results are shown in Fig. 11 and Fig. 12. Two CO2 generation peaks were observed in the CH4-TPSR profile of pure CeO2 (Fig. 11), which corresponded to the oxidation of methane by surface oxygen and bulk oxygen [44]. The CH4-TPSR profiles of the NiO/CeO2 samples were similar to that of CeO2; CO2 desorption was observed as a weak broad plateau at 320°C‒520°C in the former and as a sharp peak at 580°C‒680°C in the latter. As the NiO loading was increased from 1 wt-% to 10 wt-%, the intensity of the latter peak increased. This indicated that the activity of bulk oxygen increased and the amount of bulk oxygen was enhanced, and thus, that the incorporation of NiO improved the bulk oxygen mobility of CeO2 [45]. Generally, oxygen transport is a key step in CH4 combustion over NiO-based catalysts [46]. Therefore, the incorporation of NiO was an effective way to improve the catalytic performance in CH4 combustion.
Fig.11 MS signals of CO2 (m/z = 44) from CH4-TPSR on NiO/CeO2 with various NiO loadings.

Full size|PPT slide

Fig.12 MS signals from CH4-TPSR on CeO2 and 10 wt-% NiO/CeO2 (H2 (m/z = 2), CH4 (m/z = 16), CO (m/z = 28), and CO2 (m/z = 44)).

Full size|PPT slide

Figure 12 shows the details of the reaction during the CH4-TPSR of the CeO2 and NiO/CeO2 samples. CH4 was completely oxidized to CO2 and H2O over pure CeO2 when the temperature reached 410°C, which was very similar to the T10 of CeO2. However, H2 was generated when the temperature was above 496°C, revealing that partial oxidation of CH4 occurred because the amount of surface oxygen in CeO2 was insufficient to achieve the complete oxidization of CH4. When NiO was incorporated onto CeO2, both the activity and the amount of surface oxygen in CeO2 were improved. As a result, CH4 was completely oxidized to CO2 and H2O by the surface oxygen at 320°C‒520°C.

Electron paramagnetic resonance (EPR)

EPR is a sensitive and efficient method for the direct investigation and detection of paramagnetic materials, but it can only characterize magnetically active materials or substances with a single electron. According to the principles of EPR detection, the unpaired 4f electrons of Ce3+ should produce a paramagnetic signal under the external magnetic field. However, as shown in Fig. 13, the paramagnetic signal of pure CeO2 was weak and it was difficult to distinguish the signals of the unpaired electrons. This occurred because the Ce3+ signal is broadened by rapid spin-lattice relaxation at room temperature, resulting in undetectable signals [46]. It is believed that the paramagnetic signal of Ce3+ can only be detected at low temperature (close to or even below the liquid helium temperature), as the rapid spin-lattice relaxation effect is avoided at such low temperatures. After the CeO2 was loaded with NiO, a weak paramagnetic signal appeared at g = 2.009, corresponding to the oxygen radical (O); the paramagnetic signal appearing near g = 2.029 corresponded to the Ni2+ ion or to oxygen ions (O2) absorbed on the surface of CeO2. It was difficult to observe the paramagnetic signal of Ce3+ on pure CeO2. The signal of Ce3+ was still difficult to observe after NiO loading. The 10 wt-% NiO/CeO2 sample exhibited broad paramagnetic signals at g = 2.384 and g = 2.474, which were signals of the unpaired electrons in the Ni+ ion, indicating that Ce promoted the reduction of Ni.
Fig.13 EPR signals of (a) CeO2 and (b) 10 wt-% NiO/CeO2 (inset: magnification of the EPR spectrum).

Full size|PPT slide

TEM characterization

Figure 14 shows the change in the particle size before and after loading NiO on the surface of CeO2. Since the combustion of the precursor in the direct thermal decomposition process was intense, and no agents such as surfactants were added to control the morphology, the prepared CeO2 (D-CeO2) had no regular morphology, and its particle size was not uniform. After loading NiO, namely, in the 10 wt-% NiO/CeO2 sample, most of the NiO was well-dispersed, but a few large NiO flakes were present. These results indicated that some of the NiO aggregated, which was consistent with the XRD results.
Fig.14 TEM images of (a) CeO2 and (b) 10 wt-% NiO/C.

Full size|PPT slide

Long term stability of the NiO/CeO2 catalyst

The long-term stability of the NiO/CeO2 catalyst was also investigated. The stability test of the NiO/CeO2 catalyst was conducted at 600°C for 150 h with a WHSV of 15000 mL·g1·h1. As shown in Fig. 15, the CH4 conversion of the 10 wt-% NiO/CeO2 catalyst remained stable at 100% without any decrease, which suggested that the synthesized NiO/CeO2 had good stability in CH4 combustion.
Fig.15 Methane conversion over 10 wt-% NiO/CeO2 vs. reaction time at 600°C (gas space velocity= 15000 mL·g1·h1).

Full size|PPT slide

The XRD patterns of fresh and used 10 wt-% NiO/CeO2 catalysts are presented in Fig. 16(a). For the fresh sample, the diffraction peaks of the cubic CeO2 fluorite structure and a weak diffraction peak of the cubic NiO phase were observed. After 150 h of reaction, no distinct changes to the diffraction peaks of CeO2 and NiO were observed. This indicated that the NiO nanoparticles were still highly dispersed on the CeO2. Generally speaking, pure CeO2 and NiO can be sintered easily at high temperature [25], which would result in deactivation during CH4 combustion. The sintering between CeO2 and NiO at high temperature was partly prevented in the 10 wt-% NiO/CeO2 catalyst by the appropriate interactions between NiO and CeO2. Thus, the presence of NiO enhanced the stability of CeO2 in CH4 combustion. H2-TPR profiles of fresh and used samples are shown in Fig. 16(b). After 150 h of reaction, the reduction temperature and the intensity of the b peak (corresponding to the reduction of NiO) of the used catalyst were similar to those of the fresh sample, which indicated that NiO was very stable when supported on CeO2. Evident changes in the a1 and a2 peaks (corresponding to reduction of adsorbed oxygen) in the used catalyst were compared to those of the fresh sample, revealing that the amount of adsorbed oxygen decreased due to the consumption of oxygen during long-term reaction. In addition, the decrease in the amount of adsorbed oxygen was also caused by the decrease in the BET surface area (shown in Table 2) of NiO/CeO2 after reaction at 600°C for 150 h.
Fig.16 (a) XRD patterns of CeO2, fresh 10 wt-% NiO/CeO2, and used 10 wt-% NiO/CeO2; (b) H2-TPR profiles of fresh and used 10 wt-% NiO/CeO2 catalyst.

Full size|PPT slide

As shown in Fig. 17(a), a weak adsorption band was observed in the FT-IR spectra of fresh and used 10 wt-% NiO/CeO2 at ~1640 cm1, which corresponded to the O–H vibration. Adsorption bands at 1065, 1350, and 1510 cm1 were also observed in both the fresh and used samples, and were ascribed to unidentate carbonates [47]. Due to surface reconstruction or the basicity of the surface, the formation of carbonates on the surface of solids containing CeO2 is favorable, and can lead to a decline in catalytic activity [48]. The intensity of these peaks did not change remarkably after the reaction, which indicated that new carbonates were not formed. The TG-DTA curves of fresh and used 10 wt-% NiO/CeO2 catalyst are presented in Fig. 17(b). No evident weight loss was observed above 200°C (<2%). This indicated that no coke was deposited on the surface of the used catalyst.
Fig.17 (a) FT-IR spectra of fresh and used 10 wt-% NiO/CeO2; (b) TG-DTA curves of used NiO/CeO2 catalyst.

Full size|PPT slide

Conclusions

In summary, a series of NiO/CeO2 catalysts were prepared via a facile impregnation method. When the NiO loading amount was 10 wt-%, the NiO/CeO2 exhibited high catalytic performance in CH4 combustion, with low T10, T50, and T90 values of 333°C, 415°C and 467°C, respectively. Moreover, the 10 wt-% NiO/CeO2 catalyst also demonstrated high stability and good water resistance during CH4 combustion, as evidenced by the fact that it maintained a CH4 conversion of 100% after 150 h of reaction with a 600°C stream.
The interaction between CeO2 and NiO was also investigated. The oxygen vacancy concentration, activity, and amount of surface oxygen were found to be enhanced in the NiO/CeO2 catalyst due to the incorporation of Ni2+ into the CeO2 lattice. As a result, the mobility of bulk oxygen in CeO2 was increased, and the activation energy of CH4 combustion decreased. NiO was well-dispersed on the surface of CeO2, which prevented the aggregation of NiO. In this way, CeO2 improved the reduction properties of NiO and provided a much greater amount of oxygen species for CH4 combustion. Therefore, the synthesized 10 wt-% NiO/CeO2 catalyst is an excellent catalyst for CH4 combustion.
Further work will be focused on the catalytic mechanism of CH4 combustion over NiO/CeO2. Additionally, the development of a more active NiO/CeO2 catalyst should be feasible through optimization of the catalyst parameters.

Acknowledgements

This work was supported financially by Shanghai Sailing Program (17YF1413100) and Shanghai Rising-Star Program (19QB1401700).
1
Yoshida H, Nakajima T, Yazawa Y, Hattori T. Support effect on methane combustion over palladium catalysts. Applied Catalysis B: Environmental, 2007, 71(1): 70–79

DOI

2
Zhou F, Xia T, Wang X, Zhang Y, Sun Y, Liu J. Recent developments in coal mine methane extraction and utilization in China: A review. Journal of Natural Gas Science and Engineering, 2016, 31: 437–458

DOI

3
Shao C, Li W, Lin Q, Huang Q, Pi D. Low temperature complete combustion of lean methane over cobalt-nickel mixed-oxide catalysts. Energy Technology (Weinheim), 2016, 5(4): 604–610

DOI

4
Ramírez-López R, Elizalde-Martinez I, Balderas-Tapia L. Complete catalytic oxidation of methane over Pd/CeO2-Al2O3: The influence of different ceria loading. Catalysis Today, 2010, 150(3-4): 358–362

DOI

5
Fouladvand S, Schernich S, Libuda J, Grönbeck H, Pingel T, Olsson E, Skoglundh M, Carlsson P A. Methane oxidation over Pd supported on ceria-alumina under rich/lean cycling conditions. Topics in Catalysis, 2013, 56(1-8): 410–415

6
Lei Y, Li W, Liu Q, Lin Q, Zheng X, Huang Q, Guan S, Wang X, Wang C, Li F. Typical crystal face effects of different morphology ceria on the activity of Pd/CeO2 catalysts for lean methane combustion. Fuel, 2018, 233: 10–20

DOI

7
Huang Q, Li W, Lin Q, Zheng X, Pan H, Pi D, Shao C, Hu C, Zhang H. Catalytic performance of Pd-NiCo2O4/SiO2 in lean methane combustion at low temperature. Journal of the Energy Institute, 2018, 91(5): 733–742

DOI

8
Svensson E E, Boutonnet M, Järås S G. Stability of hexaaluminate-based catalysts for high-temperature catalytic combustion of methane. Applied Catalysis B: Environmental, 2008, 84(1): 241–250

DOI

9
Ordóñez S, Paredes J R, Díez F V. Sulphur poisoning of transition metal oxides used as catalysts for methane combustion. Applied Catalysis A, General, 2008, 341(1-2): 174–180

DOI

10
Zhang Y, Qin Z, Wang G, Zhu H, Dong M, Li S, Wu Z, Li Z, Wu Z, Zhang J, Catalytic performance of MnOx-NiO composite oxide in lean methane combustion at low temperature. Applied Catalysis B: Environmental, 2013, 129(2): 172–181

DOI

11
Pena M A, Fierro J L G. ChemInform abstract: Chemical structures and performance of perovskite oxides. Chemical Reviews, 2001, 101(7): 1981–2017

DOI

12
Hu L, Peng Q, Li Y. Selective synthesis of Co3O4 nanocrystal with different shape and crystal plane effect on catalytic property for methane combustion. Journal of the American Chemical Society, 2008, 130(48): 16136–16137

DOI

13
Corbella B M, Palacios J M. Titania-supported iron oxide as oxygen carrier for chemical-looping combustion of methane. Fuel, 2007, 86(1-2): 113–122

DOI

14
Teng F, Chen M, Li G, Teng Y, Xu T, Hang Y, Yao W, Santhanagopalan S, Meng D D, Zhu Y. High combustion activity of CH4 and catalluminescence properties of CO oxidation over porous Co3O4 nanorods. Applied Catalysis B: Environmental, 2011, 110: 133–140

DOI

15
Qiao D, Lu G, Mao D, Liu X, Li H, Guo Y, Guo Y. Effect of Ca doping on the catalytic performance of CuO-CeO2 catalysts for methane combustion. Catalysis Communications, 2010, 11(9): 858–861

DOI

16
Ashour S S. Structural, textural and catalytic properties of pure and Li-doped NiO/Al2O3 and CuO/Al2O3 catalysts. Journal of Saudi Chemical Society, 2014, 18(1): 69–76

DOI

17
Shan W, Luo M, Ying P, Shen W, Li C. Reduction property and catalytic activity of Ce1−xNixO2 mixed oxide catalysts for CH4 oxidation. Applied Catalysis A, General, 2003, 246(1): 1–9

DOI

18
Sun J, Wang Y, Li J, Xiao G, Zhang L, Li H, Cheng Y, Sun C, Cheng Z, Dong Z, H2 production from stable ethanol steam reforming over catalyst of NiO based on flowerlike CeO2 microspheres. International Journal of Hydrogen Energy, 2010, 35(7): 3087–3091

DOI

19
Fan Y, Xu X, Peng H, Yu H, Dai Y, Liu W, Ying J, Qi S, Xiang W. Porous NiO nano-sheet as an active and stable catalyst for CH4 deep oxidation. Applied Catalysis A, General, 2015, 507: 109–118

DOI

20
Kobayashi Y, Horiguchi J, Kobayashi S, Yamazaki Y, Omata K, Nagao D, Konno M, Yamada M. Effect of NiO content in mesoporous NiO-Al2O3 catalysts for high pressure partial oxidation of methane to syngas. Applied Catalysis A, General, 2011, 395(1): 129–137

DOI

21
Xu S, Yan X, Wang X. Catalytic performances of NiO-CeO2 for the reforming of methane with CO2 and O2. Fuel, 2006, 85(14-15): 2243–2247

DOI

22
Ding C, Liu W, Wang J, Liu P, Zhang K, Gao X, Ding G, Liu S, Han Y, Ma X. One step synthesis of mesoporous NiO-Al2O3 catalyst for partial oxidation of methane to syngas: The role of calcination temperature. Fuel, 2015, 162(9): 148–154

DOI

23
Sun C, Sun J, Xiao G, Zhang H, Qiu X, Li H, Chen L. Mesoscale organization of nearly monodisperse flowerlike ceria microspheres. Journal of Physical Chemistry B, 2006, 110(27): 13445–13452

DOI

24
Moretti E, Lenarda M, Riello P, Storaro L, Talon A, Frattini R, Reyes-Carmona A, Jiménez-López A, Rodríguez-Castellón E. Influence of synthesis parameters on the performance of CeO2-CuO and CeO2-ZrO2-CuO systems in the catalytic oxidation of CO in excess of hydrogen. Applied Catalysis B: Environmental, 2013, 129(3): 556–565

DOI

25
Thaicharoensutcharittham S, Meeyoo V, Kitiyanan B, Rangsunvigit P, Rirksomboon T. Catalytic combustion of methane over NiO/Ce0.252 catalyst. Catalysis Communications, 2009, 10(5): 673–677

DOI

26
Liu Z, Zhou Z, He F, Chen B, Zhao Y, Xu Q. Catalytic decomposition of N2O over NiO-CeO2 mixed oxide catalyst. Catalysis Today, 2017, 293: 56–60

DOI

27
Rombi E, Cutrufello M G, Atzori L, Monaci R, Ardu A, Gazzoli D, Deiana P, Ferino I. CO methanation on Ni-Ce mixed oxides prepared by hard template method. Applied Catalysis A, General, 2016, 515: 144–153

DOI

28
Dai Q, Wang X, Lu G. Low-temperature catalytic combustion of trichloroethylene over cerium oxide and catalyst deactivation. Applied Catalysis B: Environmental, 2008, 81(3): 192–202

DOI

29
Miao Y, Lu G, Liu X, Guo Y, Wang Y, Guo Y. Effects of preparation procedure in sol-gel method on performance of MoO3/SiO2 catalyst for liquid phase epoxidation of propylene with cumene hydroperoxide. Journal of Molecular Catalysis A: Chemical, 2009, 306(1-2): 17–22

DOI

30
Kleitz F, Choi S H, Ryoo R. Cubic Ia3d large mesoporous silica: Synthesis and replication to platinum nanowires, carbon nanorods and carbon nanotubes. Chemical Communications, 2003, 9(17): 2136–2137

DOI

31
Roh H S, Jun K W, Dong W S, Chang J S, Park S E, Joe Y I. Highly active and stable Ni/Ce-ZrO2 catalyst for H2 production from methane. Journal of Molecular Catalysis A Chemical, 2002, 181(1-2): 137–142

DOI

32
Gonzalez-Delacruz V, Holgado J, Pereniguez R, Caballero A. Morphology changes induced by strong metal-support interaction on a Ni-ceria catalytic system. Journal of Catalysis, 2008, 257(2): 307–314

DOI

33
Royer S, Duprez D, Kaliaguine S. Oxygen mobility in LaCoO3 perovskites. Catalysis Today, 2006, 112(1): 99–102

DOI

34
Laosiripojana N, Assabumrungrat S. Methane steam reforming over Ni/Ce-ZrO2 catalyst: Influences of Ce-ZrO2 support on reactivity, resistance toward carbon formation, and intrinsic reaction kinetics. Applied Catalysis A, General, 2005, 290(1-2): 200–211

DOI

35
Giezen J C V, Berg F R V D, Kleinen J L, Dillen A J V, Geus J W. The effect of water on the activity of supported palladium catalysts in the catalytic combustion of methane. Catalysis Today, 1999, 47(1-4): 287–293

DOI

36
Ciuparu D, Lyubovsky M R, Altman E, Pfefferle L D, Datye A. Catalytic combustion of methane over palladium-based catalysts. Catalysis Reviews, 2002, 44(4): 593–649

DOI

37
Bahlawane N. Kinetics of methane combustion over CVD-made cobalt oxide catalysts. Applied Catalysis B: Environmental, 2006, 67(3): 168–176

DOI

38
Vovchok D, Guild C J, Llorca J, Palomino R M, Waluyo I, Rodriguez J A, Suib S L, Senanayake S D. Structural and chemical state of doped and impregnated mesoporous Ni/CeO2 catalysts for the water-gas shift. Applied Catalysis A, General, 2018, 567: 1–11

DOI

39
Vidal H, Kašpar J, Pijolat M, Colon G, Bernal S, Cordón A, Perrichon V, Fally F. Redox behavior of CeO2-Zr2 mixed oxides. Applied Catalysis B: Environmental, 2000, 27(1): 49–63

DOI

40
Bueno-López A, Such-Basáñez I, Lecea S M D. Stabilization of active Rh2O3 species for catalytic decomposition of N2O on La-, Pr-doped CeO2. Journal of Catalysis, 2006, 244(1): 102–112

DOI

41
Chan S S, Wachs I E. In situ laser Raman spectroscopy of nickel oxide supported on g-Al2O3. Journal of Catalysis, 1987, 103(1): 224–227

DOI

42
Pu Z Y, Lu J Q, Luo M F, Xie Y L. Study of oxygen vacancies in Ce0.9Pr0.1O2−d solid solution by in situ X-ray diffraction and in situ Raman spectroscopy. Journal of Physical Chemistry C, 2007, 111(50): 18695–18702

DOI

43
Wolf D. Microkinetic analysis of the oxidative conversion of methane. Dependence of rate constants on the electrical properties of (CaO)x (CeO2)1−x catalysts. Catalysis Letters, 1994, 27(1-2): 207–220

DOI

44
Zenboury L, Azambre B, Weber J V. Transient TPSR, DRIFTS-MS and TGA studies of a Pd/ceria-zirconia catalyst in CH4 and NO2 atmospheres. Catalysis Today, 2008, 137(2-4): 167–173

DOI

45
Li Y, Zhang B, Tang X, Xu Y, Shen W. Hydrogen production from methane decomposition over Ni/CeO2 catalysts. Catalysis Communications, 2006, 7(6): 380–386

DOI

46
Readman J E, Olafsen A, And J B S, Blom R. Chemical looping combustion using NiO/NiAl2O4: Mechanisms and kinetics of reduction-oxidation (red-ox) reactions from in situ powder x-ray diffraction and thermogravimetry experiments. Annals of the New York Academy of Sciences, 2006, 20(4): 1382–1387

47
Fernández-García M, MartíNez-Arias A, Guerrero-Ruiz A, Conesa J C, Soria J. Ce-Zr-Ca ternary mixed oxides: Structural characteristics and oxygen handling properties. Journal of Catalysis, 2002, 211(2): 326–334

DOI

48
Djinović P, Levec J, Pintar A. Effect of structural and acidity/basicity changes of CuO-CeO2 catalysts on their activity for water-gas shift reaction. Catalysis Today, 2008, 138(3-4): 222–227

DOI

Outlines

/