Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-23T17:08:30.528Z Has data issue: false hasContentIssue false

Relative coherent stacking potential of fundamental particles of illite-smectite and its relationship to geological environment

Published online by Cambridge University Press:  09 July 2018

Il-Mo Kang
Affiliation:
Korea Institute of Geoscience and Mineral Resources, 124 Gwahang-no, Yuseong-gu, Daejeon, 305-350, Korea
S. Hillier
Affiliation:
The James Hutton Institute, Craigiebuckler, Aberdeen AB15 8QH, Scotland, UK
Yungoo Song*
Affiliation:
Department of Earth System Sciences, Yonsei University, 134, Shinchon-dong, Seodaemun-ku, Seoul, 120-749, Korea
In-Joon Kim
Affiliation:
Korea Institute of Geoscience and Mineral Resources, 124 Gwahang-no, Yuseong-gu, Daejeon, 305-350, Korea

Abstract

Interstratified illite-smectite (I-S) occurring authigenically in diverse earth crust environments reacts toward more illite-rich phases as temperature increases. For that reason, I-S is used for geothermometry when prospecting for hydrocarbons or ore mineral deposits. This study develops the mathematical relations for characterizing the coherent stacking potential of fundamental particles (FP) using the expandability ratio K, where K is defined as (%SMAX –; %SXRD)/%SMAX. The ratio can be applied to differentiating I-S samples from shales, bentonites, and hydrothermal alterations. In particular, patterns on a K vs. T diagram, where T is the average thickness of fundamental particles (FPs), appear to be indicative of the geological conditions related to I-S formation. Shale samples plot in the negative K domain of the diagram, possibly due to the intimate mixing of detrital particles. Both bentonitic and hydrothermal samples display trends of increasing K with T, which suggests the coherent stacking potential progressively decreases as FPs increase in thickness. Hydrothermal samples are more extensively distributed on the diagram than samples from bentonites. This result may reflect differences in particle growth conditions (nutrients and space) between bentonites (short supply) and hydrothermal alterations (good supply).

Type
Research Papers
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2012

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

Altaner, S.P. & Ylagan, R.F. (1997) Comparison of structural models of mixed-layer illite/smectite and reaction mechanisms of smectite illitization. Clays and Clay Minerals, 45, 517–533.CrossRefGoogle Scholar
Bove, D.J., Eberl, D.D., McCarty, D.K. & Meeker, G.P. (2002) Characterization and modeling of illite crystal particles and growth mechanisms in a zoned hydrothermal deposit, Lake City, Colorado. American Mineralogist, 87, 1546–1556.CrossRefGoogle Scholar
Cuadros, J. & Altaner, S.P. (1998) Characterization of mixed-layer illite-smectite from bentonites using microscopic, chemical, and X-ray methods; constraints on the smectite-to-illite transformation mechanism. American Mineralogist, 83, 762–774.Google Scholar
Dudek, T. & Środoń, J. (2003) Thickness distribution of illite crystals in shales. II: origin of the distribution and the mechanism of smectite illitization in shales. Clays and Clay Minerals, 51, 529–542.Google Scholar
Dudek, T. Środoń, J., Eberl, D.D., Elsass, F. & Uhlik, P. (2002) Thickness distribution of illite crystals in shales. I: X-ray diffraction vs. high-resolution transmission electron microscopy measurements. Clays and Clay Minerals, 50, 562–577.CrossRefGoogle Scholar
Eberl, D.D. & Środoń, J. (1988) Ostwald ripening and interparticle-diffraction effects for illite crystals. American Mineralogist, 73, 1335–1345.Google Scholar
Eberl, D.D., Środoń, J., Lee, M., Nadeau, P.H. & Northrop, H.R. (1987) Sericite from the Silverton caldera, Colorado: Correlation among structure, composition, origin, and particle thickness. American Mineralogist, 72, 914–935.Google Scholar
Eberl, D.D., Nüesch, R., Sucha, V. & Tsipursky, S. (1998) Measurement of fundamental illite particle thicknesses by X-ray diffraction using PVP-10 intercalation. Clays and Clay Minerals, 46, 89–97.Google Scholar
Honty, M., Uhlík, P., Šucha, V., Čaplovičová, M., Franců, J., Clauer, N. & Biroň, A. (2004) Smectite-to-illite alteration in salt-bearing bentonites (the East Slovak Basin). Clays and Clay Minerals, 52, 533–551.Google Scholar
Inoue, A. & Kitagawa, R. (1994) Morphological characteristics of illitic clay minerals from a hydrothermal system. American Mineralogist, 79, 700–711.Google Scholar
Kang, I.M, Kim, M.H., Song, Y. & Moon, H.S. (2008) Identification of randomly interstratified illite/smectite with basal peak widths. American Mineralogist, 93, 1478–1480.CrossRefGoogle Scholar
Kang, I.M, Kim, M.H. & Moon, H.S. (2009) Finding the layer scattering origin of rectorite for basal peak calculations. American Mineralogist; 94, 1411–1416.Google Scholar
Kasama, T., Murakami, T., Kohyama, N. & Watanabe, T. (2001) Experimental mixtures of smectite and rectorite: re-investigation of “fundamental particles” and “interparticle diffraction.” American Mineralogist, 86, 105–114.Google Scholar
Meunier, A. & Velde, B. (2010) Illite, p. 286. Springer-Verlag, Berlin.Google Scholar
Nadeau, P.H. (1985) The physical dimensions of fundamental clay particles. Clay Minerals, 20, 499–514.Google Scholar
Nadeau, P.H. & Bain, J.M. (1986) Composition of some smectites and diagenetic illitic clays and implications for their origin. Clays and Clay Minerals, 34, 455–464.Google Scholar
Nadeau, P.H., Wilson, M.J., McHardy, W.J. & Tait, J.M. (1984) Interparticle diffraction: A new concept for interstratified clays. Clay Minerals, 19, 757–759.CrossRefGoogle Scholar
Plançon, A. (2004) Consistent modeling of the XRD patterns of mixed-layer phyllosilicates. Clays and Clay Minerals, 52, 47–54.Google Scholar
Pollastro, R.M. (1993) Considerations and applications of the illite/smectite geothermometer in hydrocarbon-bearing rocks of Miocene to Mississippian age. Clays and Clay Minerals, 41, 119–133.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) Monograph 5, Mineralogical Society, London.Google Scholar
Sakharov, B.A., Plançon, A., Lanson, B. & Drits, V.A. (2004) Influence of the outer surface layers of crystals on the X-ray diffraction intensity of basal reflections. Clays and Clay Minerals, 52, 680–692.Google Scholar
Środoń, J. & Elsass, F. (1994) Effect of the shape of fundamental particles on XRD characteristics of illitic minerals. European Journal of Mineralogy, 6, 113–122.Google Scholar
Środoń, J., Elsass, F., McHardy, W.J. & Morgan, D.J. (1992) Chemistry of illite-smectite inferred from TEM measurements of fundamental particles. Clay Minerals, 27, 137–158.CrossRefGoogle Scholar
Środoń, J., Eberl, D.D. & Drits, V.A. (2000) Evolution of fundamental-particle size during illitization of smectite and implications of reaction mechanism. Clays and Clay Minerals, 48, 446–458.Google Scholar
Šucha, V., Kraus, I., Gerthofferova, H., Petes, J. & Serekova, M. (1993) Smectite to illite conversion in bentonites and shales of the East Slovak Basin. Clay Minerals, 28, 243–253.Google Scholar
Šucha, V., Środoń, J., Elsass, F. & McHardy, W.J. (1996) Particle shape versus coherent scattering domain of illite/smectite: evidence from HRTEM of Dolná Ves clays. Clays and Clay Minerals, 44, 665–671.Google Scholar
Ylagan, R.F., Altaner, S.P. & Pozzuoli, A. (2000) Reaction mechanisms of smectite illitization associated with hydrothermal alteration from Ponza Island, Italy. Clays and Clay Minerals, 48, 610–631.Google Scholar