Next Article in Journal
Effect of PET Size, Content and Mixing Process on the Rheological Characteristics of Flexible Pavement
Next Article in Special Issue
The Visible-Light-Driven Activity of Biochar-Doped TiO2 Photocatalysts in β-Blockers Removal from Water
Previous Article in Journal
An Optimalization Study on the Surface Texture and Machining Parameters of 60CrMoV18-5 Steel by EDM
Previous Article in Special Issue
Outperformance in Acrylation: Supported D-Glucose-Based Ionic Liquid Phase on MWCNTs for Immobilized Lipase B from Candida antarctica as Catalytic System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modification of CeNi0.9Zr0.1O3 Perovskite Catalyst by Partially Substituting Yttrium with Zirconia in Dry Reforming of Methane

by
Mahmud S. Lanre
1,
Ahmed E. Abasaeed
1,*,
Anis H. Fakeeha
1,
Ahmed A. Ibrahim
1,*,
Abdullah A. Alquraini
1,
Salwa B. AlReshaidan
2 and
Ahmed S. Al-Fatesh
1,*
1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
2
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(10), 3564; https://doi.org/10.3390/ma15103564
Submission received: 22 March 2022 / Revised: 16 April 2022 / Accepted: 21 April 2022 / Published: 16 May 2022

Abstract

:
Methane Dry Reforming is one of the means of producing syngas. CeNi0.9Zr0.1O3 catalyst and its modification with yttrium were investigated for CO2 reforming of methane. The experiment was performed at 800 °C to examine the effect of yttrium loading on catalyst activity, stability, and H2/CO ratio. The catalyst activity increased with an increase in yttrium loading with CeNi0.9Zr0.01Y0.09O3 catalyst demonstrating the best activity with CH4 conversion >85% and CO2 conversion >90% while the stability increased with increases in zirconium loading. The specific surface area of samples ranged from 1–9 m2/g with a pore size of 12–29 nm. The samples all showed type IV isotherms. The XRD peaks confirmed the formation of a monoclinic phase of zirconium and the well-crystallized structure of the perovskite catalyst. The Temperature Program Reduction analysis (TPR) showed a peak at low-temperature region for the yttrium doped catalyst while the un-modified perovskite catalyst (CeNi0.9Zr0.1O3) showed a slight shift to a moderate temperature region in the TPR profile. The Thermogravimetric analysis (TGA) curve showed a weight loss step in the range of 500–700 °C, with CeNi0.9Zr0.1O3 having the least carbon with a weight loss of 20%.

1. Introduction

Fossil fuels such as natural gas, coal, and crude oil serve as the backbone for the 21st century energy, industrial, and transportation sectors of the economy [1]. Their usage liberates CO2 that accumulates over time in the atmosphere depleting the ozone layer shield of the earth, leading to the rising temperature of the earth causing global warming. The impact of this global warming on the environment includes ecosystem collapse, desertification, etc., and can have direct effects on humans such as extremely hot weather conditions, or indirect effects such as crop failure and farmland loss causing a food shortage. In order to save the earth, the emissions of the greenhouse gases must be drastically reduced; hence, this necessitates dry reforming of methane (DRM) reaction which employs methane and carbon dioxide greenhouse gases to produce syngas (H2/CO), thereby controlling the mitigation of these two global warming gases [2]. DRM can also utilize biogas, composed mainly of CO2 and CH4. This reaction is strongly endothermic.
The dry reforming reaction comprises:
Methane Dry Reforming: CO2 + CH4 ↔ 2CO + 2H2
Reverse Water-Gas Shift Reaction: CO2 + H2 ↔ CO + H2O
Methane Cracking Reaction: CH4 ↔ C + H2
Disproportionation Reaction: 2CO ↔ C + CO2
Carbon Gasification: C + H2O ↔ CO + H2
Synthesis gas (syngas) obtained from DRM reaction can be converted to methanol and synthetic fuels by Shell middle distillate synthesis (SMDS) or Fischer–Tropsch syntheses [3]. Hydrogen production is being increased during the conversion of CO with steam in water gas shift reaction to CO2 [1]. Suitable catalysts for the DRM have been investigated using transition elements like Ni, Co, Pd, Ir, and perovskite-type oxides, which are very active [4]. Nickel catalysts are majorly considered for dry reforming reactions [5,6,7,8] based on its low price. However, coking and sintering cause the catalyst to deactivate [9,10]. The two major characteristics of the catalyst that defines coking are surface properties and acidity [11]. Carbon deposition depends on factors such as the nature of the hydrocarbon and the catalyst and reaction operating conditions [11]. Coking can be mitigated by supporting the active metal on a metal oxide with strong basicity [11]. Methane adsorption on nickel requires the bond breaking of C-H. The CH4 is converted thus: CH4→CH3*→CH2*→CH*→C* [11]. Ni catalysts are being bolstered so as to hamper carbon deposits and achieve good stability leading to the development of catalysts with an abundance of oxygen for coke gasification such as perovskite catalyst [12].
The structural stability of perovskite is altered by incorporating an atom into its structure resulting in changes in oxygen mobility [13]. It was reported that lower valent cations addition into perovskite catalyst structure resulted in oxygen ions adsorption onto the surfaces and changes to lattice oxygen. Such changes to the surface lattice oxygen ions would lead to various types of adsorption onto the surface of the perovskite catalyst structure and different mass transfer rates of oxygen [14]. Investigating the influence of cerium on nickel-based catalysts for hydrogen production and reported that the addition of the right amount of cerium can increase oxygen vacancies formation, which can activate oxygen-containing compounds to react with carbon species as soon as it forms [15]. Basically, zirconia is known to be a better support for active metal, due to heat stability and explicit characteristics such as acid-base and reduction-oxidation properties [16,17,18,19,20]. It has been proven that ZrO2 is a proper support for Ni, as it gives restricted coke formation with a small carbon combustion temperature [21]. Yttrium oxide (Y2O3) can either be acting as a promoter or support as a result of its unique chemical and thermal properties. Y2O3 supported Ni catalyst allows the ease reduction, better activity, stability, and limits the reverse of water gas shift (RWGS) reaction [22]. Yttria-zirconia as support for DRM suggests that the supported Ni catalysts will gain from redox properties of the material which is exceptional to limit the carbonaceous deposit formation thereby increasing the lifetime of the catalyst [23,24,25]. Promoters like alkali earth and alkali with rare earth metals have been used for the activity and stability enhancement of nickel-based catalysts [26,27,28]. Doping of trace amounts of noble metals to Ni catalyst leads to direct improvements of the catalytic reaction features of DRM [29,30,31].
The aim of this work is to study the effect of CeNi0.9Zr0.1O3 perovskite catalyst modified with yttrium on DRM. The incorporation of yttrium into the perovskite structure is evaluated in terms of catalyst activity, stability, amount, and nature of carbon formed.

2. Materials and Methods

The perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) were prepared by the sol-gel method with propionic acid acting as a solvent, to dissolve nitrates of each metal. In the preparation, Ni (NO3)2·6H2O (Sigma, St. Louis, MO, USA), Y(NO3)3·6H2O (Sigma), Ce (NO3)3·6H2O (Sigma), Zr (NO3)4·6H2O (Sigma), and propionic acid (C3H6O2, Sigma) were used. The nitrates were separately dissolved in propionic acid, stirred, and heated at T = 90 °C with oil as a heating medium. Afterward, the solutions were continuously stirred for about 2 h at T = 130 °C. Thereafter, the propionic acid was evaporated with a rotary evaporator at T = 70–80 °C until a gel was formed. The gel obtained was dried at T = 90 °C overnight, and calcined at 725 °C for 4 h. The catalysts formed after calcination were ground into powder and used for the DRM reaction.

2.1. Catalytic Testing

The catalysts were tested for DRM at 800 °C reaction temperature under atmospheric pressure. A packed bed reactor stainless steel reactor (0.0091 m internal diameter; 0.3 m height) was used to perform the experiment. An amount of 0.10 g of catalyst was placed in the reactor on top of glass wool. Stainless steel, sheathed thermocouple K-type, axially positioned close to the catalyst bed was used to determine the temperature during the reaction. Preceding the reaction, activation of the perovskite catalysts was done at 700 °C with H2. This lasted for 60 min and the remnant H2 was purged with N2. During the dry reforming reaction, the feed volume ratio was kept at 3:3:1 for CH4, CO2, and N2 gases, respectively, with a space velocity of 42 L/h./gcat. The outlet gas from the reactor was connected to an online Gas Chromatography (GC) with a thermal conductivity detector to analyze its composition. The CH4, CO2 conversion, and H2/CO (syngas ratio) were calculated using Equations (4)–(6):
Methane   conversion   ( % ) = CH 4 , in CH 4 , out CH 4 , in 100
Carbon   dioxide   conversion   = CO 2 , in CO 2 , out CO 2 , in 100
Syngas   Ratio = mole   of   H 2   produced mole   of   CO   produced

2.2. Catalyst Physicochemical Properties Determination

2.2.1. Nitrogen Physisorption

The perovskite catalysts surface area, as well the pore size distribution, was measured by N2 adsorption–desorption at −196 °C using a Micromeritics Tristar II 3020 for porosity and surface area analyzer.

2.2.2. Hydrogen Temperature Programmed Reduction Analysis

A total of 70 mg of the sample was loaded inside the TPR sample holder of a Micromeritics apparatus. Thereafter, TPR measurements were performed at 150 °C using Ar gas for 30 min and then cooled to ambient temperature. Thereafter, the sample was heated in a furnace up to 800 °C ramping at 10 °C min−1, in the atmosphere of H2/Ar mixture (1:9 vol. %) at 40 mL/min. The thermal conductivity detector recorded H2 consumption during the operation.

2.2.3. Thermo-Gravimetric (TGA) Analysis

Quantification of carbon deposits on the used catalysts was determined by TGA analysis. 10–15 mg of the used catalysts was filled in a platinum pan. Heating was performed at ambient temperature up to 1000 °C at 20 °C min−1 temperature ramp. Loss in mass was constantly monitored as the heating progressed.

2.2.4. X-ray Diffraction (XRD) Analysis

The X-ray Diffraction patterns of the perovskite catalysts were recorded on a Miniflex Rigaku diffractometer that was equipped with Cu Kα X-ray radiation. The device was run at 40 mA and 40 kV.

2.2.5. Transmission Electron Microscopy (TEM)

Transmission electron microscopy (JEOL JEM-2100F) with high resolution to give larger magnification was used to carry out the TEM measurement of both the fresh and used catalyst. The electron microscope operated at 200 kV produces the active metal nickel particle sizes and depicts the morphology of carbon deposit on the used catalyst. Before the TEM measurement, the catalysts were first dispersed ultrasonically in ethanol at room temperature. Thereafter, the drop from the suspension was placed in a lacey carbon-coated Copper grid to produce the images.

2.2.6. Laser Raman (NMR-4500) Spectrometer

Laser Raman (NMR-4500) Spectrometer (JASCO, Tokyo, Japan) was used to record Raman spectra of the spent catalyst samples. The wavelength of the excitation beam was set to 532 nm, and an objective lens of 100× magnification was used for the measurement. The laser intensity was adjusted to 1.6 mW. Each spectrum was received by averaging 3 exposures on 10 s. Spectra were recorded in the range 1200–3000 cm−1 (Raman shift) and were processed by using Spectra Manager Ver.2 software (JASCO, Tokyo, Japan).

3. Results

3.1. BET (Brunauer–Emmett–Teller) Analysis

The surface area of the catalysts does not vary largely with one another with pore volume less than one. The isotherm curves as shown in Figure 1 suggest that the materials are mesoporous in nature with pore diameter less than 50 nm. The isotherm linear plot of all the samples represents Type IV isotherm. This occurs due to capillary condensation of gases in the tiny pores of solid at pressures below the gas saturation pressure. The textural properties of the fresh catalyst are provided in Table 1.

3.2. Temperature-Programmed Reduction (TPR)

The active sites for the catalyst samples are Ni and Zr, the samples are activated by H2 reduction precedent to reaction. The reducibility of the catalyst samples was performed as shown in Figure 2. Peaks at the lower region are almost the same but there is a shoulder peak for un-modified sample in the moderate temperature region in the TPR profile. It has been reported that the reduction becomes easier while the energy of the metal–oxygen bond decreases [32]. The peaks at moderate temperature regions can be attributed to the reduction of Ni3+ to Ni2+.

3.3. X-ray Diffraction (XRD) Analysis

The XRD peaks of the perovskite catalysts are depicted in Figure 3 and Figure 4. At Y = 0.03–0.05, no XRD peaks of ZrO2 phases are observed. In other samples, cubic zirconia oxide (JCPDS card reference number 00-020-0684) or tetragonal zirconium oxide (JCPDS card reference number 01-081-1547) or both phases are found. The XRD peaks confirm the formation of tetragonal and cubic zirconia [33,34]. The peaks from the XRD include cubic CeO2 (JCPDS card reference number 01-002-1306), cubic NiO (JCPDS card reference number 00-002-1216), and cubic yttrium cerium oxide (JCPDS card reference number 01-075-0177). For the CeNi0.9Zr0.1O3 catalyst, the intensity of the peaks at 2θ = 28.5° and 48.2° are higher, suggesting a discrete crystalline phase of CeO2 formation. The Ni in the Ce-Ni systems exists as NiO on the ceria surface and Ni2+ ions in the CeO2 lattice [35,36]. The catalyst CeNi0.9Zr0.03Y0.07O3 had an additional rhombohedral nickel yttrium phase (JCPDS card reference number 00-020-0646).

3.4. Transmission Electron Microscope (TEM) Analysis

The TEM analysis for both the fresh and spent catalysts are illustrated in Figure 5 and Figure 6 at 200 nm magnification. The used CeNi0.9Zr0.01Y0.09O3 catalyst particles are shown with the carbon spreading across the surface. The nickel particle size distribution expressed in the nanometer is plotted for each of the TEM images with that of the spent catalyst higher than the fresh catalyst samples and the result tabulated in Table 2.

3.5. Catalyst Activity

A varying amount of yttrium was used to modify the catalyst CeNi0.9Zr0.1−xYxO3 with x = 0, 0.03, 0.05, 0.07, and 0.09) and tested at 800 °C for DRM reaction. The activity of the yttrium modified catalyst CeNi0.9Zr0.07Y0.03O3 was lower than the un-modified catalyst CeNi0.9Zr0.1O3 and had the lowest activity for all the tested catalysts. The activity of the CeNi0.9Zr0.05Y0.05O3 catalyst is slightly higher than the un-modified CeNi0.9Zr0.1O3 catalyst, with the latter having 87% CO2 and 78% CH4 conversion, as shown in Figure 7 and Figure 8. The activity of yttrium modified catalyst CeNi0.9Zr0.1−xYxO3 when x = 0.07 and 0.09 is higher than the un-modified CeNi0.9Zr0.1O3 catalyst. The former has 87% CH4 and 90% CO2 for x = 0.07, and 90% CH4 and 91% CO2 for x = 0.09. Yttrium results in oxygen vacancies formations that are believed to contribute to improved catalytic activity [37,38]. The catalyst activity increases with Y loading increase. Figure 9 displays the syngas ratios. The H2/CO ratio values for CeNi0.9Zr0.1O3 and CeNi0.9Zr0.07Y0.03O3 catalyst are less than one which suggests that the predominant side reaction is RWGS while CeNi0.9Zr0.05Y0.05O3, CeNi0.9Zr0.03Y0.07O3, CeNi0.9Zr0.01Y0.09O3 catalyst have H2/CO ratio greater than one which suggests that the Boudouard reaction is the predominant side reaction. CeO2 impregnated with yttrium increases ionic conductivity with Ce and Y bonded together, suggesting the possible presence of the CeO2-Y2O3 phase, which favors NiO dispersion and limits the sintering [28]. Y2O3 as a basic carrier allows Ni catalyst to be easily reduced and have a better activity [29]. Yttrium enhances smaller nickel crystallites, thereby improving the dispersion of active sites. The promotion with 0.09Y led to better activity which was linked to the formation of a solid solution of ZrO2-Y2O3, thereby enhancing the reduction of bulk NiO [29]. In addition, a number of oxygen vacancies and mobility can be enhanced by doping cerium with yttrium [39]. Ions of Ce3+ (0.97 Å) can be substituted by Y3+ (1.04 Å) and form a solid solution due to their similar ionic radii [39].

3.6. Thermogravimetric Analysis of the Used Catalyst (TGA)

The TGA curve of the spent catalysts was plotted in Figure 10 showing the weight loss of each catalyst expressed in percentage. The weight loss of the spent catalyst is shown to be 20%, 30%, 40%, 50%, and 55% for CeNi0.9Zr0.1O3, CeNi0.9Zr0.07Y0.03O3, CeNi0.9Zr0.05Y0.05O3, CeNi0.9Zr0.03Y0.07O3, and CeNi0.9Zr0.01Y0.09O3, respectively. The CeNi0.9Zr0.1O3 catalyst has the lowest amount of carbon compared to modified yttrium catalysts due to the former having a lesser amount of zirconium. The amount of carbon deposited on the catalyst increases with decreases in zirconium amount.

3.7. RAMAN Analysis

Raman spectra of the used catalysts (Figure 11) depict two bands with Raman shifts in the range of 1574 ± 5 cm−1 and 2650 ± 10 cm−1, corresponding to the D and G bands respectively. The D band is related to coke deposits with imperfect, that is, disordered structure (amorphous carbon), while the G band is related to well-ordered structure (graphitic carbon). The ID/IG ratio is less than one since ID is less than IG for used catalyst CeNi0.9Zr0.1−xYxO3 when x = 0, 0.03 and 0.05; hence, it contains imperfect (amorphous carbon) on the catalyst surface after reaction while on the other hand for CeNi0.9Zr0.1−xYxO3 catalysts where x = 0.07 and 0.09 have more graphitic carbon as the ID/IG ratio is greater than one for these catalysts.

4. Conclusions

The methane conversion activities of the best-modified catalyst (CeNi0.9Zr0.01Y0.09O3) is significantly higher than the base catalyst (CeNi0.9Zr0.1O3). The amount of the metal ion affects activity as CeNi0.9Zr0.1O3 modified with 0.09Y shows an increase in activity compared to CeNi0.9Zr0.1O3 modified with 0.03Y. The catalyst activity increases with an increase in yttrium loading with CeNi0.9Zr0.01Y0.09O3 catalyst having the best activity with CH4 conversion >85% and CO2 conversion >90%. The H2/CO ratios for CeNi0.9Zr0.1O3 and CeNi0.9Zr0.07Y0.03O3 catalysts are less than one, which suggests that the predominant side reaction is RWGS while CeNi0.9Zr0.05Y0.05O3, CeNi0.9Zr0.03Y0.07O3, and CeNi0.9Zr0.01Y0.09O3 catalysts having H2/CO ratios greater than one suggests that the Boudouard reaction is the predominant side reaction. The amount of carbon deposited on the catalyst decreases with the increase in zirconium amount hence zirconium helps to improve the catalyst stability being thermally stable.

Author Contributions

Experiment, M.S.L.; writing—original draft, M.S.L. and A.S.A.-F.; preparation of catalyst, M.S.L., A.A.A., S.B.A. and A.S.A.-F.; characterization of catalyst, M.S.L. and A.A.I.; writing—review and editing, A.H.F., A.E.A., A.A.A. and A.A.I. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to sincerely thank Researchers Supporting Project number (RSP-2021/368), King Saud University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Steinhauer, B.; Kasireddy, M.R.; Radnik, J.; Martin, A. Development of Ni-Pd bimetallic catalysts for the utilization of carbon dioxide and methane by dry reforming. Appl. Catal. A Gen. 2009, 366, 333–341. [Google Scholar] [CrossRef]
  2. Hanley, E.S.; Deane, J.P.; Gallachóir, B.P.Ó. The role of hydrogen in low carbon energy futures—A review of existing perspectives. Renew. Sustain. Energy Rev. 2018, 82, 3027–3045. [Google Scholar] [CrossRef]
  3. Eilers, J.; Posthuma, S.A.; Sie, S.T. The shell middle distillate synthesis process (SMDS). Catal. Lett. 1991, 7, 253–269. [Google Scholar] [CrossRef]
  4. Valderrama, G.; Goldwasser, M.R.; de Navarro, C.U.; Tatibouët, J.M.; Barrault, J.; Batiot-Dupeyrat, C.; Martínez, F. Dry reforming of methane over Ni perovskite type oxides. Catal. Today 2005, 107–108, 785–791. [Google Scholar] [CrossRef]
  5. Ewbank, J.L.; Kovarik, L.; Diallo, F.Z.; Sievers, C. Effect of metal–support interactions in Ni/Al2O3 catalysts with low metal loading for methane dry reforming. Appl. Catal. A Gen. 2015, 494, 57–67. [Google Scholar] [CrossRef] [Green Version]
  6. Oemar, U.; Ang, M.L.; Hidajat, K.; Kawi, S. Enhancing performance of Ni/La2O3 catalyst by Sr-modification for steam reforming of toluene as model compound of biomass tar. RSC Adv. 2015, 5, 17834–17842. [Google Scholar] [CrossRef]
  7. Zou, X.; Wang, X.; Li, L.; Shen, K.; Lu, X.; Ding, W. Development of highly effective supported nickel catalysts for pre-reforming of liquefied petroleum gas under low steam to carbon molar ratios. Int. J. Hydrogen Energy 2010, 35, 12191–12200. [Google Scholar] [CrossRef]
  8. Shen, K.; Wang, X.; Zou, X.; Wang, X.; Lu, X.; Ding, W. Pre-reforming of liquefied petroleum gas over nickel catalysts sup-ported on magnesium aluminum mixed oxides. Int. J. Hydrogen Energy 2011, 36, 4908–4916. [Google Scholar] [CrossRef]
  9. Raberg, L.; Jensen, M.; Olsbye, U.; Daniel, C.; Haag, S.; Mirodatos, C.; Sjastad, A. Propane dry reforming to synthesis gas over Ni-based catalysts: Influence of support and operating parameters on catalyst activity and stability. J. Catal. 2007, 249, 250–260. [Google Scholar] [CrossRef]
  10. Liu, D.; Quek, X.; Cheo, W.N.E.; Lau, R.; Borgna, A.; Yang, Y. MCM-41 supported nickel-based bimetallic catalysts with superior stability during carbon dioxide reforming of methane: Effect of strong metal–support interaction. J. Catal. 2009, 266, 380–390. [Google Scholar] [CrossRef]
  11. Fakeeha, A.H.; Bagabas, A.A.; Lanre, M.S.; Osman, A.; Kasim, S.O.; Ibrahim, A.A.; Arasheed, R.; Alkhalifa, A.; Elnour, A.Y.; Abasaeed, A.E.; et al. Catalytic Performance of Metal Oxides Promoted Nickel Catalysts Supported on Mesoporous γ-Alumina in Dry Reforming of Methane. Process 2020, 8, 522. [Google Scholar] [CrossRef]
  12. Lu, H.; Yang, X.; Gao, G.; Wang, J.; Han, C.; Liang, X.; Li, C.; Li, Y.; Zhang, W.; Chen, X. Metal (Fe, Co, Ce or La) doped nickel catalyst supported on ZrO2 modified mesoporous clays for CO and CO2 methanation. Fuel 2016, 183, 335–344. [Google Scholar] [CrossRef]
  13. Zhi, G.; Guo, X.; Wang, Y.; Jin, G.; Guo, X. Effect of La2O3 modification on the catalytic performance of Ni/SiC for methanation of carbon dioxide. Catal. Commun. 2011, 16, 56–59. [Google Scholar] [CrossRef]
  14. De la Cruz, R.G.; Falcón, H.; Peña, M.A.; Fierro, J. Role of bulk and surface structures of La1−xSrxNiO3 perovskite-type oxides in methane combustion. Appl. Catal. B Environ. 2001, 33, 45–55. [Google Scholar] [CrossRef]
  15. De Lima, S.M.; da Silva, A.M.; da Costa, L.O.; Assaf, J.M.; Mattos, L.V.; Sarkari, R.; Venugopal, A.; Noronha, F.B. Hydrogen production through oxidative steam reforming of ethanol over Ni-based catalysts derived from La1−xCexNiO3 perovskite-type oxides. Appl. Catal. B Environ. 2012, 121–122, 1–9. [Google Scholar] [CrossRef]
  16. Stagg-Williams, S.M.; Noronha, F.B.; Fendley, G.; Resasco, D.E. CO2 Reforming of CH4 over Pt/ZrO2 Catalysts Promoted with La and Ce Oxides. J. Catal. 2000, 194, 240–249. [Google Scholar] [CrossRef] [Green Version]
  17. Reddy, G.K.; Loridant, S.; Takahashi, A.; Delichère, P.; Reddy, B.M. Reforming of methane with carbon dioxide over Pt/ZrO2/SiO2 catalysts—Effect of zirconia to silica ratio. Appl. Catal. A Gen. 2010, 389, 92–100. [Google Scholar] [CrossRef]
  18. Mustu, H.; Yaşyerli, S.; Yasyerli, N.; Dogu, G.; Doğu, T.; Djinović, P.; Pintar, A. Effect of synthesis route of mesoporous zirconia based Ni catalysts on coke minimization in conversion of biogas to synthesis gas. Int. J. Hydrogen Energy 2015, 40, 3217–3228. [Google Scholar] [CrossRef]
  19. Zhang, X.; Zhang, M.; Zhang, J.; Zhang, Q.; Tsubaki, N.; Tan, Y.; Han, Y. Methane decomposition and carbon deposition over Ni/ZrO2 catalysts: Comparison of amorphous, tetragonal, and monoclinic zirconia phase. Int. J. Hydrogen Energy 2019, 44, 17887–17899. [Google Scholar] [CrossRef]
  20. Gong, D.; Li, S.; Guo, S.; Tang, H.; Wang, H.; Liu, Y. Lanthanum and cerium co-modified Ni/SiO2 catalyst for CO methanation from syngas. Appl. Surf. Sci. 2018, 434, 351–364. [Google Scholar] [CrossRef]
  21. Santamaria, L.; Arregi, A.; Alvarez, J.; Artetxe, M.; Amutio, M.; Lopez, G.; Bilbao, J.; Olazar, M. Performance of a Ni/ZrO2 catalyst in the steam reforming of the volatiles derived from biomass pyrolysis. J. Anal. Appl. Pyrolysis 2018, 136, 222–231. [Google Scholar] [CrossRef]
  22. Damyanova, S.; Shtereva, I.; Pawelec, B.; Mihaylov, L.; Fierro, J.L.G. Characterization of none and yttrium-modified Ni-based catalysts for dry reforming of methane. Appl. Catal. B Environ. 2020, 278, 119335. [Google Scholar] [CrossRef]
  23. Muñoz, M.; Calvino, J.J.; Rodríguez-Izquierdo, J.; Blanco, G.; Arias, D.; Omil, J.A.P.; Hernandez-Garrido, J.C.; González-Leal, J.; Cauqui, M.; Yeste, M. Highly stable ceria-zirconia-yttria supported Ni catalysts for syngas production by CO2 reforming of methane. Appl. Surf. Sci. 2017, 426, 864–873. [Google Scholar] [CrossRef]
  24. Bellido, J.D.A.; Assaf, E.M. Effect of the Y2O3–ZrO2 support composition on nickel catalyst evaluated in dry reforming of methane. Appl. Catal. A Gen. 2009, 352, 179–187. [Google Scholar] [CrossRef]
  25. Wang, Y.; Li, L.; Wang, Y.; Da Costa, P.; Hu, C. Highly Carbon-Resistant Y Doped NiO–ZrOm Catalysts for Dry Reforming of Methane. Catalyst 2019, 9, 1055. [Google Scholar] [CrossRef] [Green Version]
  26. Liu, H.R.; Świrk, K.; Galvez, M.E.; da Costa, P. Nickel Supported Modified Ceria Zirconia Lanthanum/Praseodymium/Yttrium Oxides Catalysts for Syngas Production through Dry Methane Reforming. Mater. Sci. Forum 2018, 941, 2214–2219. [Google Scholar] [CrossRef]
  27. Burbano, M.; Norberg, S.T.; Hull, S.; Eriksson, S.G.; Marrocchelli, D.; Madden, P.A.; Watson, G.W. Oxygen Vacancy Ordering and the Conductivity Maximum in Y2O3-Doped CeO2. Chem. Mater. 2011, 24, 222–229. [Google Scholar] [CrossRef]
  28. Munteanu, G.; Petrova, P.; Ivanov, I.; Liotta, L.F.; Kaszkur, Z.; Tabakova, T.; Ilieva, L. Temperature-programmed reduction of lightly yttrium-doped Au/CeO2 catalysts: Correlation between oxygen mobility and WGS activity. J. Therm. Anal. 2017, 131, 145–154. [Google Scholar] [CrossRef]
  29. Fakeeha, A.H.; Kasim, S.O.; Ibrahim, A.A.; Abasaeed, A.E.; Al-Fatesh, A.S. Influence of Nature Support on Methane and CO2 Conversion in a Dry Reforming Reaction over Nickel-Supported Catalysts. Materials 2019, 12, 1777. [Google Scholar] [CrossRef] [Green Version]
  30. Verykios, X. Catalytic dry reforming of natural gas for the production of chemicals and hydrogen. Int. J. Hydrogen Energy 2003, 28, 1045–1063. [Google Scholar] [CrossRef]
  31. Lima, S.; Assaf, J.; Peña, M.; Fierro, J. Structural features of La1−xCexNiO3 mixed oxides and performance for the dry reforming of methane. Appl. Catal. A Gen. 2006, 311, 94–104. [Google Scholar] [CrossRef]
  32. Pyatnitskii, Y.I.; Bostan, A.I.; Raevskaya, L.N.; Nedil’Ko, S.A.; Dzyaz’Ko, A.G.; Zen’Kovich, E.G. Effect of the Composition of Mn-, Co- and Ni-Containing Perovskites on their Catalytic Properties in the Oxidative Coupling of Methane. Theor. Exp. Chem. 2005, 41, 117–121. [Google Scholar] [CrossRef]
  33. Peña, M.A.; Fierro, J.L.G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 2001, 101, 1981–2017. [Google Scholar] [CrossRef] [PubMed]
  34. Ferri, D. Methane combustion on some perovskite-like mixed oxides. Appl. Catal. B Environ. 1998, 16, 119–126. [Google Scholar] [CrossRef]
  35. Goula, M.A.; Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I.V. Syn-gas production via the biogas dry reforming reaction over Ni supported on zirconia modified with CeO2 or La2O3 catalysts. Int. J. Hydrogen Energy 2017, 42, 13724–13740. [Google Scholar] [CrossRef]
  36. Charisiou, N.; Siakavelas, G.; Tzounis, L.; Sebastian, V.; Monzon, A.; Baker, M.; Hinder, S.; Polychronopoulou, K.; Yentekakis, I.; Goula, M. An in depth investigation of deactivation through carbon formation during the biogas dry reforming reaction for Ni supported on modified with CeO2 and La2O3 zirconia catalysts. Int. J. Hydrogen Energy 2018, 43, 18955–18976. [Google Scholar] [CrossRef]
  37. Świrk, K.; Rønning, M.; Motak, M.; Grzybek, T.; Da Costa, P. Synthesis strategies of Zr- and Y-promoted mixed oxides derived from double-layered hydroxides for syngas production via dry reforming of methane. Int. J. Hydrogen Energy 2021, 46, 12128–12144. [Google Scholar] [CrossRef]
  38. Fernández, J.M.; Barriga, C.; Ulibarri, M.A.; Labajos, F.M.; Rives, V. New Hydrotalcite-like Compounds Containing Yttrium. Chem. Mater. 1997, 9, 312–318. [Google Scholar] [CrossRef]
  39. Wang, J.B.; Tai, Y.-L.; Dow, W.-P.; Huang, T.-J. Study of ceria-supported nickel catalyst and effect of yttria doping on carbon dioxide reforming of methane. Appl. Catal. A Gen. 2001, 218, 69–79. [Google Scholar] [CrossRef]
Figure 1. Nitrogen physisorption isotherms of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09).
Figure 1. Nitrogen physisorption isotherms of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09).
Materials 15 03564 g001
Figure 2. H2-TPR profiles of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0, 0.07 and 0.09).
Figure 2. H2-TPR profiles of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0, 0.07 and 0.09).
Materials 15 03564 g002
Figure 3. XRD pattern of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.00, 0.07 and 0.09).
Figure 3. XRD pattern of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.00, 0.07 and 0.09).
Materials 15 03564 g003
Figure 4. XRD pattern of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.03 and 0.05).
Figure 4. XRD pattern of perovskite catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.03 and 0.05).
Materials 15 03564 g004
Figure 5. TEM micrographs and matching particle size distribution for fresh CeNi0.9Zr0.1O3. (A) and used CeNi0.9Zr0.1O3. (B) catalyst.
Figure 5. TEM micrographs and matching particle size distribution for fresh CeNi0.9Zr0.1O3. (A) and used CeNi0.9Zr0.1O3. (B) catalyst.
Materials 15 03564 g005
Figure 6. TEM micrographs and matching particle size distribution for fresh CeNi0.9Zr0.01Y0.09O3. (A) and used CeNi0.9Zr0.01Y0.09O3. (B) catalyst.
Figure 6. TEM micrographs and matching particle size distribution for fresh CeNi0.9Zr0.01Y0.09O3. (A) and used CeNi0.9Zr0.01Y0.09O3. (B) catalyst.
Materials 15 03564 g006
Figure 7. CO2 conversion of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Figure 7. CO2 conversion of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Materials 15 03564 g007
Figure 8. CH4 conversion of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Figure 8. CH4 conversion of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Materials 15 03564 g008
Figure 9. H2/CO ratio of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Figure 9. H2/CO ratio of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C, 1 atmosphere and GHSV = 42 L/(h.gcat).
Materials 15 03564 g009
Figure 10. TGA Curves of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C.
Figure 10. TGA Curves of perovskite catalysts CeNi0.9Zr0.1−xYxO3 (x = 0, 0.03, 0.05, 0.07, and 0.09) at 800 °C.
Materials 15 03564 g010
Figure 11. Raman Spectra of Used Perovskite Catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.03, 0.05, 0.07, and 0.09).
Figure 11. Raman Spectra of Used Perovskite Catalyst CeNi0.9Zr0.1−xYxO3 (x = 0.03, 0.05, 0.07, and 0.09).
Materials 15 03564 g011
Table 1. Tabular presentation of Specific surface area (SSA), Pore volume (Pv), and Pore diameter (Pd) of the fresh catalyst samples.
Table 1. Tabular presentation of Specific surface area (SSA), Pore volume (Pv), and Pore diameter (Pd) of the fresh catalyst samples.
SamplesSSA (m2/g)Pv (cm3/g)Pd (nm)
CeNi0.9Zr0.1O39.350.03112.12
CeNi0.9Zr0.07Y0.03O31.270.00315.44
CeNi0.9Zr0.05Y0.05O31.620.00314.43
CeNi0.9Zr0.03Y0.07O31.850.00222.21
CeNi0.9Zr0.01Y0.09O32.660.01129.33
Table 2. Summary of Nickel Particle Size Derived from TEM Analysis.
Table 2. Summary of Nickel Particle Size Derived from TEM Analysis.
NameNi Particle Size
Fresh CeNi0.9Zr0.1O39.44 nm
Used CeNi0.9Zr0.1O311.68 nm
Fresh CeNi0.9Zr0.01Y0.09O36.66 nm
Used CeNi0.9Zr0.01Y0.09O37.10 nm
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lanre, M.S.; Abasaeed, A.E.; Fakeeha, A.H.; Ibrahim, A.A.; Alquraini, A.A.; AlReshaidan, S.B.; Al-Fatesh, A.S. Modification of CeNi0.9Zr0.1O3 Perovskite Catalyst by Partially Substituting Yttrium with Zirconia in Dry Reforming of Methane. Materials 2022, 15, 3564. https://doi.org/10.3390/ma15103564

AMA Style

Lanre MS, Abasaeed AE, Fakeeha AH, Ibrahim AA, Alquraini AA, AlReshaidan SB, Al-Fatesh AS. Modification of CeNi0.9Zr0.1O3 Perovskite Catalyst by Partially Substituting Yttrium with Zirconia in Dry Reforming of Methane. Materials. 2022; 15(10):3564. https://doi.org/10.3390/ma15103564

Chicago/Turabian Style

Lanre, Mahmud S., Ahmed E. Abasaeed, Anis H. Fakeeha, Ahmed A. Ibrahim, Abdullah A. Alquraini, Salwa B. AlReshaidan, and Ahmed S. Al-Fatesh. 2022. "Modification of CeNi0.9Zr0.1O3 Perovskite Catalyst by Partially Substituting Yttrium with Zirconia in Dry Reforming of Methane" Materials 15, no. 10: 3564. https://doi.org/10.3390/ma15103564

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