Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-07-04T14:22:42.218Z Has data issue: false hasContentIssue false

Reactivity and Stability of Superionics MSnF4 at High Temperature in Various Media

Published online by Cambridge University Press:  16 February 2011

Georges Dénès
Affiliation:
Concordia University, Department of Chemistry and Biochemistry, Laboratory of Solid State Chemistry and Mössbauer Spectroscopy, and Laboratories for Inorganic Materials, Montréal, Québec, Canada, gdenes@vax2.concordia.ca
M. Cecilia Madamba
Affiliation:
Concordia University, Department of Chemistry and Biochemistry, Laboratory of Solid State Chemistry and Mössbauer Spectroscopy, and Laboratories for Inorganic Materials, Montréal, Québec, Canada, gdenes@vax2.concordia.ca
Abdualhafeed Muntasar
Affiliation:
Concordia University, Department of Chemistry and Biochemistry, Laboratory of Solid State Chemistry and Mössbauer Spectroscopy, and Laboratories for Inorganic Materials, Montréal, Québec, Canada, gdenes@vax2.concordia.ca
Get access

Abstract

Superionic MSnF4 are the highest performance fluoride ion conductors, with PbSnF4 being the best. Prototypes of devices using PbSnF4 have been constructed and tested. Since the fluoride ion mobility is thermally activated, some devices might be used more efficiently above ambient temperature. Therefore, it is of prime importance that the stability of these materials be tested under potential conditions of use, since thermal degradation and phase transitions are likely to alter the conducting properties. Tetragonal SrSnF4, α-PbSnF4 and BaSnF4, and orthorhombic o-PbSnF4, are stable at ambient conditions, even in air. However, all undergo significant deterioration when heated in air: the color changes from white to yellowish, and tin hydrolysis and oxidation takes place. They are much more stable under inert conditions (nitrogen or argon). However, PbSnF4 undergoes several phase transitions at high temperatures: o to α starting at ca. 100°C, α to β (reversible, but the reverse reaction is very sluggish) starting at 250°C, β to γ (reversible) at 390°C. β-PbSnF4 can be quenched to ambient temperature and is metastable for long times, however, eventually it starts changing to stable α-PbSnF4 and this change is uncontrollable and is faster above ambient temperature. Microcrystalline μγ-PbSnF4, obtained by ball milling any other phase of PbSnF4, gives rapidly a-PbSnF4 at 200°C. In addition, for all MSnF4, hydrolysis of the Sn-F bonds to Sn-O occurs with traces of moisture.

Type
Research Article
Copyright
Copyright © Materials Research Society 1999

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

1. Donaldson, J. D. and Senior, B. J., J. Chem. Soc. (A), 1821 (1967).Google Scholar
2. Dénès, G., Pannetier, J. and Lucas, J., C. R. Acad. Sc. Paris, C 280, 831 (1975).Google Scholar
3. Dénès, G., Madamba, M. C. and Parris, J. M. in Solid State Ionics IV, edited by Nazri, G. A., Tarascon, J. M. and Schrieber, M. (Mater. Res. Soc. Proc. 369, Pittsburgh, PA 1995), p. 463468.Google Scholar
4. Calandrino, R., Collin, A., Dénès, G., Madamba, M. C. and Parris, J. M. in Solid State Chemistry of Inorganic Materials, edited by Davies, P. K, Jacobson, A. J., Torardi, C. C. and Vanderah, T. A. (Mater. Res. Soc. Proc. 453, Pittsburgh, PA 1997), p. 585590.Google Scholar
5. Pannetier, J., Dénès, G. and Lucas, J., Mat. Res. Bull. 14, p. 626 (1979).Google Scholar
6. Dénès, G. and Madamba, M. C., Materials Structure 3, 227 (1996).Google Scholar
7. Birchall, T., Dénès, G., Ruebenbauer, K. and Pannetier, J., Hyperf. Interact. 29, 1331 (1986).Google Scholar
8. Dénès, G., Yu, Y. H., Tyliszczak, T. and Hitchcock, A. P, J. Solid State Chem. 91, p. 1 (1991).Google Scholar
9. Collin, A., Dénès, G., Le Roux, D., Madamba, M. C., Parris, J. M. and Salafin, A., J. Inorg. Mater., submitted.Google Scholar
10. Dénès, G. and Laou, E., Hyperf. Interact. 92, 1013 (1994).Google Scholar
11. Dénès, G. and Laou, E. in Structure and Properties of Interfaces in Ceramics, edited by Bonnel, D., ROhle, M. and Chowdhry, U. (Mater. Res. Soc. Proc. 357, Pittsburgh, PA 1995), p. 109114.Google Scholar
12. Dénès, G., Gueune, A., Laou, E. and Le Hudrou, S., in High Temperature Electrochemistry: Ceramics and Metals, edited by Poulsen, F. W., Bonanos, N., Linderoth, S., Mogensen, M. and Zachau-Christiansen, B., Proc. 17th Risø Intern. Symp. Mater. Sc., p. 223228 (1996).Google Scholar
13. Wakagi, A. and Kuwano, J., J. Mater. Chem. 41, 973 (1994).Google Scholar
14. Wakagi, A., Kuwano, J., Kato, M. and Hanomoto, H., Solid State Ionics 70/71, 601 (1994).Google Scholar
15. Dénès, G., J. Solid State Chem. 77, 54 (1988).Google Scholar
16. Dénès, G., Milova, G. and Antonov, B. D. in Solid State Ionics V, edited by Nazri, G. A., Julien, C. and Rougier, A. (Mater. Res. Soc. Proc. xxx, Pittsburgh, PA 1999), submitted.Google Scholar
17. Dénès, G., Madamba, M. C. and Muntasar, A. in Solid State Chemistry of Inorganic Materials II, edited by Zur Loye, H. C., Kauzlarich, S. M., Sleight, A. W. and McCarron, E. M. III (Mater. Res. Soc. Proc., Pittsburgh, PA 1999), submitted.Google Scholar