Hostname: page-component-77c89778f8-fv566 Total loading time: 0 Render date: 2024-07-18T08:52:39.276Z Has data issue: false hasContentIssue false

EXAFS and XANES spectroscopy study of the oxidation and deprotonation of biotite

Published online by Cambridge University Press:  05 July 2018

Bernd Güttler
Affiliation:
Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EQ, U.K.
Wilhelm Niemann
Affiliation:
Haldor Topsøe Research Laboratories, DK-2800 Lyngby, Denmark
Simon A. T. Redfern
Affiliation:
Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EQ, U.K.

Abstract

The coupled thermal oxidation and deprotonation in air of iron-rich biotite (FeO + Fe2O3 = 34%) has been investigated by EXAFS and XANES spectroscopy at the Fe-K edge and by XANES spectroscopy at the Ti-K edge. Samples annealed for 5 h at temperatures between 250° to 600°C have been studied. Distortions mainly of the Fe-Fe correlation within the octahedral layers are reflected in increasing Debye-Waller factors of the Fe-Fe correlation peak proportional to the annealing temperature. Unchanged Fe-O nearest-neighbour and Fe-Fe next-nearest-neighbour coordination numbers show that these distortions, nonetheless, do not change the structural topology of the octahedral layers. A model is introduced to demonstrate that increasing distortions are compatible with the expected heterogenous deprotonation mechanism in biotite. Titanium occurs in octahedral coordination. It was found to be unaffected by the coupled oxidation/deprotonation process. Both the coordination number and the valence state stay constant during the annealing process, in spite of dramatic changes of the Fe2+/Fe3+ ratio. Thermally activated hopping conduction involving Ti according to Fe2+Ti4+ → Fe3+Ti3+ is, therefore, not a significant process during thermal deprotonation and oxidation in biotite.

Type
Silicate Mineralogy
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 1989

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

Addison, C. C. and Sharp, J. H. (1962) Mechanism for the oxidation of ferrous iron in hydroxylated silicates. Clay Mineral. Bull. 5, 73-9.CrossRefGoogle Scholar
Bagin, V. I., Gendler, T. S., Dainyak, L. G. and Kuz'min, R. N. (1980) Mössbauer, thermomagnetic and X-ray study of cation ordering and high-temperature decomposition in biotite. Clays and Clay Minerals 28, 188-96.CrossRefGoogle Scholar
Beran, A. and Bittner, H. (1974) Untersuchungen zur Kristallchemie des Ilvaits. Tscherm. Min. Petr. Mitt. 21, 11-29.CrossRefGoogle Scholar
Bonnin, D., Calas, G., Suquet, H., and Pezerat, H. (1985) Site occupancy of Fe+ in garfield nontronite: a spectroscopic study. Phys. Chem. Minerals 12, 5564.Google Scholar
Brindley, G. W. and Lemaitre, J. (1987) Thermal oxidation and reduction reactions of clay minerals. p. 3170. In Chemistry of Clays and Clay Minerals (Newman, A. C. D., ed.), Mineralogical Society Monograph No. 6, 319-70.Google Scholar
Brindley, G. W. and McKinstry, H. A. (1961) The kaolinite mullite reaction series: IV The coordination of aluminium. J. Am. Ceram. Soc. 44, 506-7.CrossRefGoogle Scholar
Calas, G. and Petiau, J. (1983) Coordination of iron in oxide glasses through high-resolution K-edge spectra: Information from the pre-edge. Solid State Commun. 48, 625-9.CrossRefGoogle Scholar
Brindley, G. W., Basset, W. A., Petiau, J., Steinberg, M., Tchoubar, D. and Zarka, A. (1984) Some mineralogical applications of synchrotron radiation. Phys. Chem. Minerals 11, 7-36.Google Scholar
Davidson, A. T. and Yoffe, D. (1968) Hopping electrical conduction and thermal breakdown in natural and synthetic mica. Phys, Stat. Sol. 30, 741-54.CrossRefGoogle Scholar
Durham, P. J., Pendry, J. B. and Hodges, C. H. (1982) Calculation of X-ray absorption near-edge structure. XANES. Comput. Phys'. Commun. 25, 193-205.CrossRefGoogle Scholar
Farmer, V. C. (1971) The layer silicates. In Infrared spectra of minerals (Farmer, V. C., ed.), Mineralogical Society Monograph No. 4.Google Scholar
Faye, G. H. (1968a) The optical absorption spectra of iron in six-coordinate sites in chlorite, biotite, phlogopite and vivianite. Some aspects of pleochroism in sheet silicates. Can. Mineral. 9, 403-25.Google Scholar
Faye, G. H. (1968b) The optical absorption spectra of certain transition metal ions in muscovite, lepidolite and fuchsite. Can. J. Earth Science 5, 31-8.CrossRefGoogle Scholar
Ferrow, E. (1987) Mössbauer and X-ray studies on the oxidation of annite and ferriannite. Phys. Chem. Minerals 14, 270-5.CrossRefGoogle Scholar
Geiger, C. A., Henry, D. L., Bailey, S. W. and Maj, J. J. (1983) Crystal structure of cronstedtite-2H2 . Clays and Clay Minerals 31, 97-108.CrossRefGoogle Scholar
Greegor, R. B., Lytle, F. W., Sandstrom, D. R., Wong, J. and Schultz, P. (1983) Investigation of TiO2-SiO2 glasses by x-ray absorption spectroscopy. J. Non- Cryst. Solids 55, 27-43.CrossRefGoogle Scholar
Guggenheim, S., Chang, Y. and Koster van Groos, A. F. (1987) Muscovite dehydroxylation: High-temperature studies. Am. Mineral. 72, 537-50.Google Scholar
Hazen, R. M. and Burnham, C. W. (1973) The structure of one-layer phlogopite and annite. Ibid. 58, 88-900.Google Scholar
Heller-Kallai, L. and Rozenson, I. (1980) Dehydroxylation of dioctahedral phyllosilicates. Clays and Clay Minerals 28, 355-68.CrossRefGoogle Scholar
Iwai, S. and Shimamune, T. (1975) X-ray studies of the metakaolin state of dickite. In Contributions to clay mineralogy dedicated to Professor Toshio Sudo (Henmi, K., ed.), 30-3. Tokyo University of Education, Bunkio-Ku, Tokyo.Google Scholar
Manceau, A. and Combes, J. M. (1988) Structure of Mn and Fe oxides and oxyhydroxides: A topological approach by EXAFS. Phys. Chem. Minerals 15, 283-95.CrossRefGoogle Scholar
Meunier, J. F., Currie, M. R., Wertheimer, M. R. and Yelon, A. (1983) Electrical conduction in biotite micas. J. Appl. Phys. 54, 898-905.CrossRefGoogle Scholar
Niemann, W., Malzfeldt, W., Rabe, P. and Haensel, R. (1987) Critical cluster size for mixed valence in small matrix-isolated Sm clusters. Phys. Rev. B 35, 1099-1107.CrossRefGoogle ScholarPubMed
Norman, D., Garg, K. B. and Durham, P. J. (1985) The X-ray absorption near edge structure of transition metal oxides: A one-electron interpretation. Sol. State Cornmun. 56, 895-8.CrossRefGoogle Scholar
Ohta, T., Takeda, H. and Takeuchi, Y. (1982) Mica polytypism: Similarities in the crystal structure of coexisting 1M and 2M1 oxybiotite. Am. Mineral. 67, 298-310.Google Scholar
Otten, M. T. and Buseck, P. R. (1987) The oxidation state of Ti in hornblende and biotite determined by electron-energy-loss spectroscopy, with inferences regarding the Ti substitution. Phys. Chem. Minerals 14, 45-51.CrossRefGoogle Scholar
Pampuch, R. (1971) Le mdcanisme de la déshydroxylation des hydroxydes et des silicates pbylliteux. Bull. Grpe franc. Argiles 23, 107-18.CrossRefGoogle Scholar
Putnis, A. and Giittler, B. (in preparation) TEM studies on the deprotonation and oxidation of biotite.Google Scholar
Robbins, D. W. and Strens, R. G. J. (1968) Polarization dependence and oscillator strength of metal-metal charge-transfer bands in iron (II, III) silicate minerals. Chem. Comm. 508-9.Google Scholar
Rossman, G. R. (1984) Spectroscopy of micas. In: Micas (Baily, S. W., ed.). Reviews in Mineralogy 13, 145-76.CrossRefGoogle Scholar
Rouxhet, P. G., Gillard, J. L. and Fripiat, J. J. (1972) Thermal decomposition of amosite, crocidolite and biotite. Mineral. Mag. 38, 583-92.CrossRefGoogle Scholar
Sanz, J., Gonzales-Carreño, T. and Gancedo, R. (1983) On dehydroxylation mechanisms of a biotite in vacuo and in oxygen. Phys. Chem. Minerals 9, 14-8.CrossRefGoogle Scholar
Sherman, D. M. (1987) Molecular orbital (SCF-Xα- SW) theory of metal-metal charge transfer processes in minerals. I. Application to Fe2+-Ti 4+ charge transfer and ‘electron delocalization’ in mixed valence oxides and silicates. Ibid. 14, 364-7.Google Scholar
Smith, G., Howes, B. and Hasan, Z. (1980) Mössbauer and optical spectra of biotite: a case of Fe2+-Fe3+ interactions. Phys. Stat. Sol. (a) 57, K18792.CrossRefGoogle Scholar
Takeda, H. and Ross, M. (1975) Mica polytypism: Dissimilarities in the crystal structure of coexisting 1M and 2M1 oxybiotite. Am. Mineral. 60, 1030-40.Google Scholar
Thompson, J. B. (1973) In: Hazen, R. M. and Burnham, C. W. (1973) as cited above.Google Scholar
Tossel, J. A., Vaughan, D. J. and Johnson, K. H. (1974) The electronic structure of rutile, wiistite, and hematite from molecular orbital calculations. Am. Mineral. 59, 319-34.Google Scholar
Vedder, W. and Wilkins, R. W. T. (1969) Dehydroxylation and rehydroxylation, oxidation and reduction of micas. Ibid. 54, 482-509.Google Scholar
Waychunas, G. A. (1987) Synchrotron radiation XANES spectroscopy of Ti in minerals: Effects of bonding distances, Ti valence, and site geometry on absorption edge structure. Ibid. 72, 89-101.Google Scholar
Waychunas, G. A. and Rossman, G. R. (1983) Spectroscopic standard for tetrahedrally coordinated ferric iron: LiA102: Fe3+ . Phys. Chem. Minerals 9, 212-5.CrossRefGoogle Scholar
Waychunas, G. A. Apted, M. J. and Brown, G. E. (1983) X-ray edge absorption spectra of Fe minerals and model compounds: near-edge structure. Ibid. 10, 1-9.CrossRefGoogle Scholar
Wirth, R. (1985) Dehydration of mica (phengite) by electron bombardment in a transmission electron microscope (TEM). J. Mat. Sci. Lett. 4, 327-30.CrossRefGoogle Scholar