Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T23:08:15.842Z Has data issue: false hasContentIssue false

Ce3+-enriched core–shell ceria nanoparticles for silicate adsorption

Published online by Cambridge University Press:  27 June 2017

Kijung Kim
Affiliation:
Department of Nanoscale Semiconductor Engineering, Hanyang University, Seoul 04763, South Korea
Jihoon Seo
Affiliation:
Department of Energy Engineering, Hanyang University, Seoul 04763, South Korea
Myoungjae Lee
Affiliation:
Department of Nanoscale Semiconductor Engineering, Hanyang University, Seoul 04763, South Korea
Jinok Moon
Affiliation:
Department of Energy Engineering, Hanyang University, Seoul 04763, South Korea; and Clean/CMP Technology Team, Memory, Samsung Electronics, Gyeonggi-Do 16677, South Korea
Kangchun Lee
Affiliation:
Department of Energy Engineering, Hanyang University, Seoul 04763, South Korea
Dong Kee Yi*
Affiliation:
Department of Chemistry, Myongji University, Yongin 18448, South Korea
Ungyu Paik*
Affiliation:
Department of Nanoscale Semiconductor Engineering, Hanyang University, Seoul 04763, South Korea; and Department of Energy Engineering, Hanyang University, Seoul 04763, South Korea
*
a)Address all correspondence to these authors. e-mail: vitalis@mju.ac.kr
Get access

Abstract

Ce3+ ions in ceria nanoparticles (NPs) play a role as reactive sites in the adsorption of silicate anions. However, the limited concentration of Ce3+ ions in ceria NPs remains a major challenge in this regard. Herein, we report a simple strategy to synthesize Ce3+-enriched core–shell ceria NPs for enhanced adsorption of silicate anions. To increase the overall Ce3+ concentration, a shell layer is composed of Ce3+-rich ultrasmall ceria NPs approximately 5 nm in size. The Ce3+ concentration of such core–shell ceria NPs is increased by 12.7–17.1% relative to that of the pristine ceria NPs, resulting in increased adsorption of silicate anions. The Freundlich model fits the observed adsorption isotherm well and the constants of adsorption capacity (KF) and adsorption intensity (1/n) indicate higher adsorption affinity of the core–shell ceria NPs for silicate anions. We attribute these improvements to the increased Ce3+ concentration contributed by the ultrasmall ceria coating. This strategy can be used for enhancing the reactivity of ceria materials.

Type
Articles
Copyright
Copyright © Materials Research Society 2017 

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.)

Footnotes

Contributing Editor: Edson Roberto Leite

References

REFERENCES

Jasinski, P., Suzuki, T., and Anderson, H.U.: Nanocrystalline undoped ceria oxygen sensor. Sens. Actuators, B 95, 73 (2003).CrossRefGoogle Scholar
Hibino, T., Hashimoto, A., Inoue, T., Tokuno, J-i., Yoshida, S-i., and Sano, M.: A low-operating-temperature solid oxide fuel cell in hydrocarbon–air mixtures. Science 288, 2031 (2000).Google Scholar
Kozlov, A.I., Kim, D.H., Yezerets, A., Andersen, P., Kung, H.H., and Kung, M.C.: Effect of preparation method and redox treatment on the reducibility and structure of supported ceria–zirconia mixed oxide. J. Catal. 209, 417 (2002).CrossRefGoogle Scholar
Jiang, T., Wang, Z., Zhang, J., Hao, X., Rooney, D., Liu, Y., Sun, W., Qiao, J., and Sun, K.: Understanding the flash sintering of rare-earth-doped ceria for solid oxide fuel cell. J. Am. Ceram. Soc. 98, 1717 (2015).Google Scholar
Cook, L.M.: Chemical processes in glass polishing. J. Non-Cryst. Solids 120, 152 (1990).Google Scholar
Esch, F., Fabris, S., Zhou, L., Montini, T., Africh, C., Fornasiero, P., Comelli, G., and Rosei, R.: Electron localization determines defect formation on ceria substrates. Science 309, 752 (2005).Google Scholar
Campbell, C.T. and Peden, C.H.: Oxygen vacancies and catalysis on ceria surfaces. Science 309, 713 (2005).Google Scholar
Bevan, D. and Kordis, J.: Mixed oxides of the type MO2 (fluorite)—M2O3—I oxygen dissociation pressures and phase relationships in the system CeO2·Ce2O3 at high temperatures. J. Inorg. Nucl. Chem. 26, 1509 (1964).Google Scholar
Lawrence, N.J., Brewer, J.R., Wang, L., Wu, T-S., Wells-Kingsbury, J., Ihrig, M.M., Wang, G., Soo, Y-L., Mei, W-N., and Cheung, C.L.: Defect engineering in cubic cerium oxide nanostructures for catalytic oxidation. Nano Lett. 11, 2666 (2011).Google Scholar
Dutta, P., Pal, S., Seehra, M., Shi, Y., Eyring, E., and Ernst, R.: Concentration of Ce3+ and oxygen vacancies in cerium oxide nanoparticles. Chem. Mater. 18, 5144 (2006).Google Scholar
Skorodumova, N., Simak, S., Lundqvist, B.I., Abrikosov, I., and Johansson, B.: Quantum origin of the oxygen storage capability of ceria. Phys. Rev. Lett. 89, 166601 (2002).Google Scholar
Seo, J., Moon, J., Kim, J.H., Lee, K., Hwang, J., Yoon, H., Yi, D.K., and Paik, U.: Role of the oxidation state of cerium on the ceria surfaces for silicate adsorption. Appl. Surf. Sci. 389, 311 (2016).Google Scholar
Haron, M.J., Ab Rahim, F., Abdullah, A.H., Hussein, M.Z., and Kassim, A.: Sorption removal of arsenic by cerium-exchanged zeolite P. Mater. Sci. Eng., B 149, 204 (2008).Google Scholar
Song, Z., Liu, W., and Nishiguchi, H.: Quantitative analyses of oxygen release/storage and CO2 adsorption on ceria and Pt–Rh/ceria. Catal. Commun. 8, 725 (2007).Google Scholar
Marrocchelli, D. and Yildiz, B.: First-principles assessment of H2S and H2O reaction mechanisms and the subsequent hydrogen absorption on the CeO2(111) surface. J. Phys. Chem. C 116, 2411 (2012).CrossRefGoogle Scholar
Vanpoucke, D.E., Cottenier, S., Van Speybroeck, V., Van Driessche, I., and Bultinck, P.: Tetravalent doping of CeO2: The impact of valence electron character on group IV dopant influence. J. Am. Ceram. Soc. 97, 258 (2014).Google Scholar
Bueno-López, A., Krishna, K., Makkee, M., and Moulijn, J.: Enhanced soot oxidation by lattice oxygen via La3+-doped CeO2 . J. Catal. 230, 237 (2005).CrossRefGoogle Scholar
Mai, H-X., Sun, L-D., Zhang, Y-W., Si, R., Feng, W., Zhang, H-P., Liu, H-C., and Yan, C-H.: Shape-selective synthesis and oxygen storage behavior of ceria nanopolyhedra, nanorods, and nanocubes. J. Phys. Chem. B 109, 24380 (2005).Google Scholar
Masui, T., Fujiwara, K., Machida, K-i., Adachi, G-y., Sakata, T., and Mori, H.: Characterization of cerium(IV) oxide ultrafine particles prepared using reversed micelles. Chem. Mater. 9, 2197 (1997).Google Scholar
Deshpande, S., Patil, S., Kuchibhatla, S.V., and Seal, S.: Size dependency variation in lattice parameter and valency states in nanocrystalline cerium oxide. Appl. Phys. Lett. 87, 133113 (2005).Google Scholar
Tsunekawa, S., Ishikawa, K., Li, Z-Q., Kawazoe, Y., and Kasuya, A.: Origin of anomalous lattice expansion in oxide nanoparticles. Phys. Rev. Lett. 85, 3440 (2000).Google Scholar
Hailstone, R., DiFrancesco, A., Leong, J., Allston, T., and Reed, K.: A study of lattice expansion in CeO2 nanoparticles by transmission electron microscopy. J. Phys. Chem. C 113, 15155 (2009).Google Scholar
Dandu, P.V., Peethala, B., and Babu, S.: Role of different additives on silicon dioxide film removal rate during chemical mechanical polishing using ceria-based dispersions. J. Electrochem. Soc. 157, H869 (2010).Google Scholar
Huang, P., Chen, G., Jiang, Z., Jin, R., Zhu, Y., and Sun, Y.: Atomically precise Au25 superatoms immobilized on CeO2 nanorods for styrene oxidation. Nanoscale 5, 3668 (2013).Google Scholar
Ou, D.R., Mori, T., Togasaki, H., Takahashi, M., Ye, F., and Drennan, J.: Microstructural and metal–support interactions of the Pt–CeO2/C catalysts for direct methanol fuel cell application. Langmuir 27, 3859 (2011).Google Scholar
Wang, R., Crozier, P.A., Sharma, R., and Adams, J.B.: Measuring the redox activity of individual catalytic nanoparticles in cerium-based oxides. Nano Lett. 8, 962 (2008).Google Scholar
Turner, S., Lazar, S., Freitag, B., Egoavil, R., Verbeeck, J., Put, S., Strauven, Y., and Van Tendeloo, G.: High resolution mapping of surface reduction in ceria nanoparticles. Nanoscale 3, 3385 (2011).Google Scholar
Wu, L., Wiesmann, H., Moodenbaugh, A., Klie, R., Zhu, Y., Welch, D., and Suenaga, M.: Oxidation state and lattice expansion of CeO2−x nanoparticles as a function of particle size. Phys. Rev. B. 69, 125415 (2004).CrossRefGoogle Scholar
Shyu, J., Weber, W., and Gandhi, H.: Surface characterization of alumina-supported ceria. J. Phys. Chem. 92, 4964 (1988).Google Scholar
Senanayake, S.D., Stacchiola, D., Evans, J., Estrella, M., Barrio, L., Pérez, M., Hrbek, J., and Rodriguez, J.A.: Probing the reaction intermediates for the water–gas shift over inverse CeO x /Au(111) catalysts. J. Catal. 271, 392 (2010).Google Scholar
Zhang, C., Yu, Y., Grass, M.E., Dejoie, C., Ding, W., Gaskell, K., Jabeen, N., Hong, Y.P., Shavorskiy, A., and Bluhm, H.: Mechanistic studies of water electrolysis and hydrogen electro-oxidation on high temperature ceria-based solid oxide electrochemical cells. J. Am. Chem. Soc. 135, 11572 (2013).Google Scholar
Armistead, C., Tyler, A., Hambleton, F., Mitchell, S., and Hockey, J.A.: Surface hydroxylation of silica. J. Phys. Chem. 73, 3947 (1969).Google Scholar
Erdem, B., Hunsicker, R.A., Simmons, G.W., Sudol, E.D., Dimonie, V.L., and El-Aasser, M.S.: XPS and FTIR surface characterization of TiO2 particles used in polymer encapsulation. Langmuir 17, 2664 (2001).Google Scholar
Van den Brand, J., Sloof, W., Terryn, H., and De Wit, J.: Correlation between hydroxyl fraction and O/Al atomic ratio as determined from XPS spectra of aluminium oxide layers. Surf. Interface Anal. 36, 81 (2004).Google Scholar
Lee, J-S. and Choi, S-C.: Crystallization behavior of nano-ceria powders by hydrothermal synthesis using a mixture of H2O2 and NH4OH. Mater. Lett. 58, 390 (2004).Google Scholar
Nabavi, M., Spalla, O., and Cabane, B.: Surface chemistry of nanometric ceria particles in aqueous dispersions. J. Colloid Interface Sci. 160, 459 (1993).Google Scholar
Daturi, M., Finocchio, E., Binet, C., Lavalley, J.C., Fally, F., and Perrichon, V.: Study of bulk and surface reduction by hydrogen of Ce x Zr1−x O2 mixed oxides followed by FTIR spectroscopy and magnetic balance. J. Phys. Chem. B 103, 4884 (1999).Google Scholar
Ek, S., Root, A., Peussa, M., and Niinistö, L.: Determination of the hydroxyl group content in silica by thermogravimetry and a comparison with 1H MAS NMR results. Thermochim. Acta 379, 201 (2001).Google Scholar
Kim, J.M., Chang, S.M., Kong, S.M., Kim, K-S., Kim, J., and Kim, W-S.: Control of hydroxyl group content in silica particle synthesized by the sol-precipitation process. Ceram. Int. 35, 1015 (2009).Google Scholar
Mueller, R., Kammler, H.K., Wegner, K., and Pratsinis, S.E.: OH surface density of SiO2 and TiO2 by thermogravimetric analysis. Langmuir 19, 160 (2003).Google Scholar
Asgari, N., Haghighi, M., and Shafiei, S.: Synthesis and physicochemical characterization of nanostructured Pd/ceria-clinoptilolite catalyst used for p-xylene abatement from waste gas streams at low temperature. J. Chem. Technol. Biotechnol. 88, 690 (2013).Google Scholar
Langmuir, I.: The constitution and fundamental properties of solids and liquids. Part I. Solids. J. Am. Chem. Soc. 38, 2221 (1916).Google Scholar
Freundlich, H.: Over the adsorption in solution. J. Phys. Chem. 57, 385 (1906).Google Scholar
Supplementary material: File

Kim supplementary material

Kim supplementary material

Download Kim supplementary material(File)
File 747.7 KB