Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-25T10:26:48.331Z Has data issue: false hasContentIssue false

Cation reducibility of LaNi0.5Ti0.5O3, LaNi0.5Ti0.45Co0.05O3, and LaNi0.45Co0.05Ti0.5O3 perovskites from X-ray powder diffraction data using the Rietveld method

Published online by Cambridge University Press:  11 May 2022

Mayra Guamán-Ayala
Affiliation:
Departamento de Energía y Mecánica, Carrera de Petroquímica, Universidad de las Fuerzas Armadas – ESPE, EC170501 Sangolquí, Ecuador
Pablo V. Tuza*
Affiliation:
Departamento de Energía y Mecánica, Carrera de Petroquímica, Universidad de las Fuerzas Armadas – ESPE, EC170501 Sangolquí, Ecuador
Mariana M. V. M. Souza
Affiliation:
Escola de Química, Universidade Federal do Rio de Janeiro (UFRJ), Centro de Tecnologia, Bloco E, Sala 206, CEP 21941-909, Rio de Janeiro/RJ, Brazil
*
a)Author to whom correspondence should be addressed. Electronic mail: pvtuza@espe.edu.ec

Abstract

In the present work, LaNi0.5Ti0.5O3, LaNi0.5Ti0.45Co0.05O3, and LaNi0.45Co0.05Ti0.5O3 perovskites were synthesized using the modified Pechini method. After reduction, the studied perovskites changed crystal structure from the perovskite crystal structure to a cubic symmetry, with space group $Pm\bar{3}m$. The reduction partially decomposed the samples to Ni0 (Co free perovskite), Ni0–Co0, La2O3, La2TiO5, and non-stoichiometric La2NiO4, depending on H2 content of the reductive gases. The degree of reduction of nickel from LaNi0.5Ti0.5O3 reduced with 1.8% H2/Ar and 10% H2/Ar was equal to 36.5% and 95.3%, respectively, while that from LaNi0.5Ti0.45Co0.05O3 or LaNi0.45Co0.05Ti0.5O3, including cobalt, reduced with 10% H2/Ar, was equal to 71.9% and 93.9%, respectively. LaNi0.5Ti0.45Co0.05O3 showed Ni3+ and Co3+ amounts higher than the other perovskites. By increasing H2 content in the reductive mixture from 1.8% to 10%, sintering of metallic nickel was not observed. Moreover, Ni0 displayed weaker metal–support interaction than that observed for Co0, where the support was composed of La containing oxides. LaNi0.5Ti0.5O3 perovskite was used as a catalyst for steam reforming of methane. Syngas production was attributed to the number of Ni sites determined using Rietveld Refinement of X-ray diffraction pattern of this catalyst after the reaction.

Type
Technical Article
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on behalf of International Centre for Diffraction Data

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Arrivé, C., Delahaye, T., Joubert, O., and Gauthier, G. (2013). “Exsolution of nickel nanoparticles at the surface of a conducting titanate as potential hydrogen electrode material for solid oxide electrochemical cells,” J. Power Sources 223, 341348.CrossRefGoogle Scholar
Atkins, P., Overton, T., Rourke, J., Weller, M., Armstrong, F., and Hagerman, M. (2010). Shriver & Atkins’ Inorganic Chemistry (Oxford University Press, Milano), 5th ed., pp. 792796.Google Scholar
Attfield, M., Barnes, P., Cockcroft, J. K., and Driessen, H. (2004). Advanced Certificate in Powder Diffraction (School of Crystallography, Birkbeck College, University of London). Available at: http://pd.chem.ucl.ac.uk/%0Apdnn/chapter.htm.Google Scholar
Bahout, M., Managutti, P. B., Dorcet, V., Le Gal La Salle, A., Paofai, S., and Hansen, T. C. (2020). “In situ exsolution of Ni particles on the PrBaMn2O5 SOFC electrode material monitored by high temperature neutron powder diffraction under hydrogen,” J. Mater. Chem. A 8(7), 35903597.CrossRefGoogle Scholar
Benito, P., Labajos, F. M., and Rives, V. (2007). “Microwave-assisted synthesis of layered double hydroxides,” in Solid State Chemistry Research Trends, edited by Buckley, R. W. (Nova Science Publishers, New York), pp. 173225.Google Scholar
Cordero, B., Gómez, V., Platero-Prats, A. E., Revés, M., Echeverría, J., Cremades, E., Barragán, F., and Alvarez, S. (2008). “Covalent radii revisited,” Dalton Trans. 21, 28322838.CrossRefGoogle Scholar
Decorse-Pascanut, C., Berthon, J., Pinsard-Gaudart, L., Dragoe, N., and Berthet, P. (2009). “Magnetic properties of low-doped bulk Sr(Ti, Co)O3−δ perovskites,” J. Magn. Magn. Mater. 321(20), 35263531.CrossRefGoogle Scholar
Gallezot, P. (2003). “Preparation of metal clusters in zeolites,” in Post-Synthesis Modification I, Vol. 3, edited by Karge, H. G. and Weitkamp, J. (Springer, New York), pp. 257306.CrossRefGoogle Scholar
Ghogia, A. C., Cayez, S., Machado, B. F., Nzihou, A., Serp, P., Soulantica, K., and Pham Minh, D. (2020). “Hydrogen spillover in the Fischer-Tropsch synthesis on carbon-supported cobalt catalysts,” ChemCatChem 12(4), 11171128.CrossRefGoogle Scholar
Gómez-Cuaspud, J. A., Vera-López, E., Carda-Castelló, J. B., and Barrachina-Albert, E. (2017). “One-step hydrothermal synthesis of LaFeO3 perovskite for methane steam reforming,” React. Kinet. Mech. Catal. 120, 167179.CrossRefGoogle Scholar
Huang, L., Bassir, M., and Kaliaguine, S. (2005). “Reducibility of Co3+ in perovskite-type LaCoO3 and promotion of copper on the reduction of Co3+ in perovskite-type oxides,” Appl. Surf. Sci. 243, 360375.CrossRefGoogle Scholar
ICSD. (2017). “Inorganic Crystal Structure Database.” Available at: https://bdec.dotlib.com.br/.Google Scholar
Kapteijn, F., Moulijn, J. A., and Tarfaoul, A. (1993). “Temperature programmed reduction and sulphiding,” in Catalysis: An Integrated Approach to Homogeneous, Heterogeneous and Industrial Catalysis, Vol. 79, edited by Moulijn, J. A., van Leeuwen, P. W. N. M. and van Santen, R. A. (Elsevier, Amsterdam), pp. 401418.Google Scholar
Kim, C. Y., Sekino, T., and Niihara, K. (2010). “Optical, mechanical, and dielectric properties of Bi1/2Na1/2TiO3 thin film synthesized by sol-gel method,” J. Sol-Gel Sci. Technol. 55(3), 306310.CrossRefGoogle Scholar
King, G., and Woodward, P. M. (2010). “Cation ordering in perovskites,” J. Mater. Chem. 20(28), 57855796.CrossRefGoogle Scholar
Kobayashi, K. I., Kimura, T., Sawada, H., Terakura, K., and Tokura, Y. (1998). “Room-temperature magnetoresistance in an oxide material with an ordered double-perovskite structure,” Nature 395(6703), 677680.CrossRefGoogle Scholar
Mosayebi, A., and Nasabi, M. (2020). “Steam methane reforming on LaNiO3 perovskite-type oxide for syngas production, activity tests, and kinetic modeling,” Int. J. Energy Res. 44(7), 55005515.CrossRefGoogle Scholar
Panfilov, A. S., Lyogenkaya, A. A., Grechnev, G. E., Pashchenko, V. A., Vasylechko, L. O., Hreb, V. M., and Kovalevsky, A. V. (2020). “Effects of temperature and pressure on the magnetic properties of La1–xPrxCoO3,” Phys. Status Solidi B 257, 2000085.CrossRefGoogle Scholar
Paz, S. P. A., Kahn, H., and Angélica, R. S. (2018). “A proposal for bauxite quality control using the combined Rietveld – Le Bail – internal standard PXRD method – part 1: hkl model developed for kaolinite,” Miner. Eng. 118, 5261.CrossRefGoogle Scholar
Périllat-Merceroz, C., Gauthier, G., Roussel, P., Huve, M., Gelin, P., and Vannier, R. N. (2011). “Synthesis and study of a Ce-doped La/Sr titanate for solid oxide fuel cell anode operating directly on methane,” Chem. Mater. 23, 15391550.CrossRefGoogle Scholar
Rauch, R., Kiennemann, A., and Sauciuc, A. (2013). “Fischer-Tropsch synthesis to biofuels (BtL process),” in The Role of Catalysis for the Sustainable Production of Bio-Fuels and Bio-Chemicals, edited by Triantafyllidis, K., Lappas, A. and Stöcker, M. (Elsevier, Amsterdam), pp. 398444.Google Scholar
Rayner, M. K., Billing, D. G., and Coville, N. J. (2014). “In-situ X-ray diffraction activation study on an Fe/TiO2 pre-catalyst,” Acta Crystallogr. 70(3), 498509.Google Scholar
Rodríguez, E., Álvarez, I., López, M. L., Veiga, M. L., and Pico, C. (1999). “Structural, electronic, and magnetic characterization of the perovskite LaNi1−xTixO3 (0 ≤ x ≤ 1/2),” J. Solid State Chem. 148(2), 479486.CrossRefGoogle Scholar
Rodríguez-Carvajal, J. (1993). “Recent advances in magnetic structure determination by neutron powder diffraction,” Phys. B 192(1–2), 5569.CrossRefGoogle Scholar
Rodríguez-Carvajal, J. (2001). “Fullprof Manual.” Available at: http://www.ill.eu/sites/fullprof/.Google Scholar
Satterfield, C. N. (1991). Heterogeneous Catalysis in Industrial Practice (MacGraw Hill, Malabar), 2nd ed., p. 103.Google Scholar
Schimpf, S., and Muhler, M. (2009). “Methanol catalysis,” in Synthesis of Solid Catalysis, edited by de Jong, K. P. (Wiley-VCH, Weinheim), pp. 329352.CrossRefGoogle Scholar
Schmal, M. (2016). Heterogeneous Catalysis and Its Industrial Applications (Springer, Lavergne), 1st ed., pp. 135, 137 and 141.CrossRefGoogle Scholar
Smyth, D. M. (2009). “Perovskite: the electrically active structure,” Ferroelectrics 380, 113.CrossRefGoogle Scholar
Souza, M. M. V. M., Maza, A., and Tuza, P. V. (2021). “X-ray powder diffraction data of LaNi0.5Ti0.45Co0.05O3, LaNi0.45Co0.05Ti0.5O3, and LaNi0.5Ti0.5O3 perovskites,” Powder Diffr. 36(1), 2934.CrossRefGoogle Scholar
Sumathi, R., Johnson, K., Viswanathan, B., and Varadarajan, T. K. (1998). “Selective oxidation and dehydrogenation of benzyl alcohol on ABB′O3 (A = Ba, B = Pb, Ce, Ti and B′ = Bi, Cu, Sb)-type perovskite oxides-temperature programmed reduction studies,” Appl. Catal., A 172(1), 1522.CrossRefGoogle Scholar
Tuza, P. V., and Souza, M. M. V. M. (2016). “Steam reforming of methane over catalyst derived from ordered double perovskite: effect of crystalline phase transformation,” Catal. Lett. 146(1), 4753.CrossRefGoogle Scholar
Tuza, P. V., and Souza, M. M. V. M. (2017). “B-cation partial substitution of double perovskite La2NiTiO6 by Co2+: effect on crystal structure, reduction behavior and catalytic activity,” Catal. Commun. 97, 9397.CrossRefGoogle Scholar
Ufer, K., Stanjek, H., Roth, G., Dohrmann, R., Kleeberg, R., and Kaufhold, S. (2008). “Quantitative phase analysis of bentonites by the Rietveld method,” Clays Clay Miner. 56(2), 272282.CrossRefGoogle Scholar
Wang, Y., Xu, Y., Luo, L., Ding, Y., and Liu, X. (2010). “Preparation of perovskite-type composite oxide LaNi0.5Ti0.5O3-NiFe2O4 and its application in glucose biosensor,” J. Electroanal. Chem. 642(1), 3540.CrossRefGoogle Scholar
Yang, M., Huo, L., Zhao, H., Gao, S., and Rong, Z. (2009). “Electrical properties and acetone-sensing characteristics of LaNi1−xTixO3 perovskite system prepared by amorphous citrate decomposition,” Sens. Actuators, B 143(1), 111118.CrossRefGoogle Scholar
Yang, W. Z., Liu, X. Q., Lin, Y. Q., and Chen, X. M. (2012). “Structure, magnetic, and dielectric properties of La2Ni(Mn1−xTix)O6 ceramics,” J. Appl. Phys. 111, 084106.CrossRefGoogle Scholar
Yao, W. F., Xu, X. H., Wang, H., Zhou, J. T., Yang, X. N., Zhang, Y., Shang, S. X., and Huang, B. B. (2004). “Photocatalytic property of perovskite bismuth titanate,” Appl. Catal., B 52(2), 109116.CrossRefGoogle Scholar
Zhang, M., Ma, H., and Gao, Z. (2015). “Phase composition of Ni/Mg1−xNixO as a catalyst prepared for selective methanation of CO in H2-rich gas,” J. Nanomater. 2015, 110.Google Scholar
Supplementary material: File

Guamán-Ayala et al. supplementary material

Guamán-Ayala et al. supplementary material

Download Guamán-Ayala et al. supplementary material(File)
File 192.2 KB