Hostname: page-component-77c89778f8-m8s7h Total loading time: 0 Render date: 2024-07-20T00:01:13.472Z Has data issue: false hasContentIssue false

Hydrothermal reactivity of K-smectite at 300°C and 100 bar: dissolution-crystallization process and non-expandable dehydrated smectite formation

Published online by Cambridge University Press:  09 July 2018

R. Mosser-Ruck
Affiliation:
Université Henri Poincaré-UMR 7566, BP239, 54 Vandoeuvre-lès-Nancy Cedex
M. Cathelineau
Affiliation:
Université Henri Poincaré-UMR 7566, BP239, 54 Vandoeuvre-lès-Nancy Cedex
A. Baronnet
Affiliation:
CNRS-CRMC2, Campus Luminy, Case 913, 13298 Marseille, Cedex
A. Trouiller
Affiliation:
ANDRA, Parc de la Croix Blanche, 1-7 rue Jean Monnet, 92298 Chatenay-Malabry Cedex, France

Abstract

The hydrothermal reactivity of a smectite saturated with K was studied experimentally at 300°C and 100 bar in (Na,K) chloride solutions (Na/K = 0, 50 and 100, liquid/solid ratio = 10/1). X-ray diffraction, TEM and microprobe results show: (1) a partial to total dissolution of the initial smectite layers; and (2) the crystallization of newly-formed euhedral I-S. Random I-S is formed after 7 days, but an ordered mixed-layer I-S containing <30% expandable layers formed in the longest runs (112 days). The I-S is characterized by non-expandable layers of two distinct types: dehydrated smectite and illite. The Si content is lower in the I-S than in the initial smectite, thus creating a charge deficit, mostly compensated by the introduction of Na to the interlayer space, and yielding a silica release to the solution and subsequent crystallization of quartz and cristobalite.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 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

Baronnet, A., Amouric, M. & Chabot, B. (1976) Mécanismes de croissance, polytypisme et polymorphisme de la muscovite hydroxylée synthétique. J. Cryst. Growth, 32, 3759.CrossRefGoogle Scholar
Bird, G.W. (1979) Possible buffer materials for use in a nuclear waste vault. Atomic Energy of Canada Limited Technical Record, TR-72.Google Scholar
Bird, G.W. & Cameron, D.J. (1982) Vault sealing research for the Canadian nuclear fuel waste management program. Atomic Energy of Canada Limited Technical Record, TR-145.Google Scholar
Boles, J.R. & Franks, S.G. (1979) Clay diagenesis in Wilcox sandstones of southwest Texas: implication of smectite diagenesis on sandstone cementation. J. Sed. Pet. 49, 5570.Google Scholar
Boutiche, M. (1995) Stabilité physico-chimique des smectites et de I'illite en presence de solutions chargées en electrolytes: étude experiméntale a 150°C. PhD thesis, FNPL Nancy, France.Google Scholar
Cathelineau, M., Ruck, R. & Baronnet, A. (1993) Hydrothermal reactivity of Ca, Na and K-smectite transformation. Proc. 4th Int. Symp. Hydrothermal Reactions, Nancy. Google Scholar
Cuadros, J. & Linares, J. (1995) Experimental kinetic study of the smectite-to-illite transformation. Geochim. Cosmochim. Acta, 60, 3, 439453.Google Scholar
Debrabant, P., Delbart, S. & Lemaguer, D. (1985) Microanalyses géochimiques de minéraux argileux de sediments prélevés en Atlantique Nord (forages du DSDP). Clay Miner. 20, 125145.CrossRefGoogle Scholar
Eberl, D. (1978) Reaction series for dioctahedral smectites. Clays Clay Miner. 26, 327340.CrossRefGoogle Scholar
Eberl, D. & Hower, J. (1976) Kinetics of Mite formation. Geol. Soc. Am. Bull. 87, 13261330.Google Scholar
Eberl, D. & Hower, J. (1977) The hydrothermal transformation of sodium and potassium smectite into mixed-layer clay. Clays Clay Miner. 25, 215227.Google Scholar
Eberl, D. & Środoń, J. (1988) Ostwald ripening and interparticle-diffraction effects for Mite crystals. Am. Miner. 73, 13351345.Google Scholar
Fritz, B. (1981) Etude thermodynamique et modélisation des reactions hydrothermales et diagénétiques. Mém. Sci. Géol. 65.Google Scholar
Hoffmann, J. & Hower, J. (1979) Clay mineral assemblages as low grade metamorphic geothermometers: application to the thrust faulted distributed Belt of Montana, USA. Soc. Econ. Pal. Spec. Pub. 26, 5579.Google Scholar
Howard, J.J. & Roy, D.M. (1985) Development of layer charge and kinetics of experimental smectite alteration. Clays Clay Miner. 33, 8188.Google Scholar
Hower, J., Eslinger, E.V., Hower, M.E. & Perry, E.A. (1976) Mechanism of burial metamorphism of argillaceous sediments: I. Mineralogical and chemical evidence. Geol. Soc. Am. Bull. 87, 725737.Google Scholar
Huang, W.H., Longo, J.M. & Peaver, D.R. (1993) An experimental derived kinetic model for the smectiteto- illite conversion and its use as a geothermometer. Clays Clay Miner. 41, 162177.CrossRefGoogle Scholar
Inoue, A. (1983) Potassium fixation by clay minerals during hydrothermal treatment. Clays Clay Miner. 31, 8191.Google Scholar
Inoue, A., Koyama, N., Kitagawa, R. & Watanabe, T. (1988) Chemical and morphological evidence for the conversion of smectite to Mite. Clays Clay Miner. 35,2, 111120.Google Scholar
Lahann, R.W. & Roberson, H.E. (1980) Dissolution of silica from montmorillonites: effect of solution chemistry. Geochim. Cosmochim. Acta, 44, 19371943.Google Scholar
Meunier, A. & Velde, B. (1989) Solid solution in I/S mixed-layer minerals and Mite. Am. Miner. 74, 11061112.Google Scholar
Nadeau, P.H. & Reynolds, R.C. (1981) Burial and contact metamorphism in the Mancos shale. Clays Clay Miner. 29, 249259.Google Scholar
Perry, E.A. & Hower, J. (1970) Burial diagenesis in Gulf Coast pelitic sediments. Clays Clay Miner. 18, 165177.Google Scholar
Pollastro, R.M. (1985) Mineralogical and morphological evidence for formation of illite at the expense of illite/ smectite. Clays Clay Miner. 33, 265274.Google Scholar
Pusch, R. (1978) Highly compacted Na bentonite as buffer substance. KBS - Teknisk Rapport, 74, Swedish Waste Management Program, Sweden.Google Scholar
Pusch, R. (1980) Swelling pressure of highly compacted bentonite. KBS - Teknisk Rapport, 80-13, Swedish Waste Management Program, Sweden.Google Scholar
Reynolds, R.C. (1980) Interstratified clay minerals. Pp. 249-303 in Crystal Structures of Clay Minerals and their X-ray Identification (Brindley, G.W. & Brown, G., editors). Mineralogical Society, London.Google Scholar
Reynolds, R.C. (1985) NEWMOD®, a computer program for the calculation of basal X-ray diffraction intensities and mixed layered-clays. Reynolds, R.C. Jr, Hanover, New Hampshire.Google Scholar
Ransom, B. & Helgeson, H.C. (1994) A chemical and thermodynamic model of aluminous dioctahedral 2:1 layer clay minerals in diagenetic processes: Regular solution representation of interlayer dehydration in smectite. Am. J. Sci. 294, 449484.Google Scholar
Roberson, H.E. & Lahann, R.W. (1981) Smectite to illite conversion rates: effects of solution chemistry. Clays Clay Miner. 29, 129-135.Google Scholar
Small, J.S. (1994) Fluid composition, mineralogy and morphological changes associated with the smectiteto- illite reaction: an experimental investigation of the effect of organic acid anions. Clay Miner. 29, 539554.CrossRefGoogle Scholar
Weaver, C.E. & Beck, K.C. (1971) Clay-water diagenesis during burial: how mud becomes gneiss. Geol. Soc. Am. Spec. Pap. 134, 96 p.Google Scholar
Whitney, G. (1990) Role of water in the smectite-to-illite reaction. Clays Clay Miner. 38, 4, 343350.Google Scholar
Whitney, G. (1992) Dioctahedral smectite reactions at elevated temperatures: effects of K-availability, Na/ K ratio and ionic strength. Appl. Clay Sci. 7, 97112.CrossRefGoogle Scholar
Whitney, G. & Northrop, H.R. (1988) Experimental investigation of the smectite to illite reaction: dual reaction mechanisms and oxygen-isotope systematics. Am. Miner. 73, 7790.Google Scholar