Hostname: page-component-7c8c6479df-r7xzm Total loading time: 0 Render date: 2024-03-28T23:40:15.754Z Has data issue: false hasContentIssue false

Apophyllite group: effects of chemical substitutions on dehydration behaviour, recrystallization products and cell parameters

Published online by Cambridge University Press:  05 July 2018

Giselle F. Marriner
Affiliation:
Department of Geology, Royal Holloway and Bedford New College, Egham Hill, Egham, Surrey TW20 0EX
John Tarney
Affiliation:
Department of Geology, University of Leicester, Leicester LE1 7RH
J. Ian Langford
Affiliation:
School of Physics and Space Research, University of Birmingham, Birmingham B15 2TT

Abstract

The effects of (OH)/F substitutions on thermally induced reactions in the apophyllite group have been investigated using TGA and DTA methods, supplemented by X-ray diffraction, infra-red and other techniques to identify the products of reaction. The samples studied cover the range between fluorapophyllite and hydroxyapophyllite, and include some unusual ammonium varieties. The presence or absence of fluorine exerts a major control on thermally induced reactions, not only during low-temperature dehydration, but also on high temperature recrystallization reactions up to 900°C+. All apophyllites dehydrate in two stages, the first being loss of a proportion of the water with only minor distortion of the crystal lattice, whereas the second results in total collapse into an amorphous material. Higher fluorine concentrations stabilize the structure, shifting dehydration reactions to higher temperatures, and permitting more water molecules to be lost during the first stage without destruction of the structure. Fluorine is not lost during the second-stage dehydration, as previously believed, but most is retained and influences subsequent recrystallization reactions.

Dehydrated hydroxyapophyllites remain amorphous until ca. 900°C, when high-temperature wollastonite crystallizes. Fluorapophyllites initially form a metastable CaF2.SiO2 phase by 700°C, low-temperature wollastonite by 800°C, both disappearing above 900°C when fluorine loss occurs and high-temperature wollastonite is produced. TGA and DTA provide simple methods of distinguishing between hydroxy- and fluorapophyllites.

Ammonian apophyllites, with up to 25% of the stoichiometric K+ replaced by NH4+ (in the samples studied), can be recognized by their DTA patterns, the endothermic peak for the second stage reaction being significantly reduced, and that for the first, dehydration, stage being correspondingly enhanced. Such samples have weight losses well above the calculated water content, the excess being due to evolution of ammonia. Ammonian varieties also have an extra infra-red absorption band at 1460 cm−1.

Type
Mineralogy and Petrology
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 1990

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

Aumento, F. (1965) Thermal transformations of selected zeoIites and related hydrated silicates'. PhD Thesis, Dalhousie University, Halifax, Canada.Google Scholar
Bartl, H. and Pfeifer, G. (1976) Neutronen beigungs-analyse des Apophyllite KCa4(Si4O10)2(F/OH).8H2O. Neues Jahrb. Mineral. Mh., 58-65.Google Scholar
Chao, G. Y. (1971) Refinement of the crystal structure of apophyllite. II. Determination of the hydrogen positions by X-ray diffraction. Am. Mineral. 58, 1234-42.Google Scholar
Chukrov, P. V., Yermilova, L. P. and Rudnitskaya, Y. S. (1971) Nature of water in apophyllite. Nouv. Denn. Mineral. SSSR., Akad. Nauk SSSR Mineral. Muz. 20, 221-5.Google Scholar
Colville, A. H., Anderson, C. P. and Black, P. M. (1971) Refinement of the crystal structure of apophyllite. I. X-ray diffraction and physical properties. Am. Mineral. 53, 1222-33.Google Scholar
Duit, W., Jansen, J. B. H., Van Breemen, A. and Bos, A. (1986) Ammonium micas in metamorphic rocks as exemplified by Dome de L'Agout (France). Am. J. Sci. 286, 702-32.CrossRefGoogle Scholar
Dunn, P. J., Rouse, R. C. and Norberg, J. A. (1978) Hydroxyapophyllite, a new mineral and a redefinition of the apophylfite group, I. Description and nomenclature. Am. Mineral. 63, 196-9.Google Scholar
Lacy, E. D. (1965) Factors in the study of metamorphic reaction rates. In Controls of Metamorphism (Pitcher, W. S. and Flinn, G. W., eds.) 140-54. Oliver and Boyd, London.Google Scholar
Langford, J. I. (1971) Powder pattern programs. J. Appl. Crystallogr. 4, 259-60.CrossRefGoogle Scholar
Mann, L. T. (1963) Spectrophotometric determination of nitrogen in total micro-kjeldahl digest. Anal. Chem. 35, 2179-82.CrossRefGoogle Scholar
Marriner, G. F., Langford, J. I. and Tarney, J. (1979) Crystal data for calcium fluoride silicon oxide (CaF2. SiO2), a dehydration product of fluorapophyllite. J. Appl. Crystallogr. 12, 131-2.CrossRefGoogle Scholar
Matsueda, H., Miura, Y. & Rucklidge, J. (1981) Natroapophyllite, a new orthorhombic sodium analog of apophyllite. I. Description, occurrence and nomenclature. Am. Mineral. 66, 410-15.Google Scholar
Miura, Y., Kato, T., Rucklidge, J. and Matsueda, H. (1981) Natroapophyllite, a new orthorhombic sodium analog of apophyllite. II. Crystal structure. Ibid. 66, 416-23.Google Scholar
Peng, C. J. (1955) Thermal analysis study of the natrolite group. Ibid. 40, 834-56.Google Scholar
Prince, E. (1971) Refinement of the crystal structure of apophyllite. III. Determination of the hydrogen positions by neutron diffraction. Ibid. 56, 1243-51.Google Scholar
Reed, S. J. B. (1976) The electron microprobe. Cambridge University Press.Google Scholar
Rouse, R. C., Peacor, D. R. and Dunn, P. J. (1978) Hydroxyapophyllite, a new mineral and a redefinition of the apophyllite group. II. Crystal structure. Am. Mineral. 63, 199-202.Google Scholar
Sarig, S. (1973) Simultaneous TG, DTG and DTA analysis in controlled inert gas flow for the determination of desquification steps of polyhydrated sulphates. Thermochim. Acta 297-301.Google Scholar
Solomon, G. C. and Grossman, G. R. (1988) NH4 + in pegmatitic feldspars from the Southern Black Hills, South Dakota. Am. Mineral. 73, 818-21.Google Scholar
Taylor, H. F. W. and Náray-Szabò, St. (1931) The structure of apophyllite. Zeits. Kristallogr. 77, 148-56.Google Scholar
Vorma, A. (1961) A new apophyllite occurrence in Viipurl Rapkivi area. Bull. Comm. Geol. Finlande, 196, 399-404.Google Scholar
Walsh, N. (1979). The simultaneous determination of major, minor and trace constituents of silicate rocks using inductively coupled plasma. Spectrochim. Acta, 35, 107-11.CrossRefGoogle Scholar
Walsh, N., Howie, R. A. (1980) The evaluation of the performance of ICP for the determination of major and trace constituents of silicate rocks. Mineral. Mag. 43, 96-74.CrossRefGoogle Scholar
Yamamoto, T. and Nakahira, M. (1966) Ammonium ions in sericites. Am. Mineral. 51, 1775-8.Google Scholar