Hostname: page-component-76dd75c94c-h9cmj Total loading time: 0 Render date: 2024-04-30T08:58:36.356Z Has data issue: false hasContentIssue false

Celadonite and its Transformation to Smectite in an Entisol at Red Rock Canyon, Kern County, California

Published online by Cambridge University Press:  02 April 2024

D. A. Reid
Affiliation:
Department of Soil and Environmental Sciences, University of California, Riverside, California 92521
R. C. Graham
Affiliation:
Department of Soil and Environmental Sciences, University of California, Riverside, California 92521
S. B. Edinger
Affiliation:
Department of Soil and Environmental Sciences, University of California, Riverside, California 92521
L. H. Bowen
Affiliation:
Department of Chemistry, North Carolina State University, Raleigh, North Carolina 27695
J. O. Ervin
Affiliation:
Department of Soil and Environmental Sciences, University of California, Riverside, California 92521

Abstract

A green, Lithic Torriorthent soil derived from a celadonite-rich, hydrothermally altered basalt immediately north of the Mojave Desert region in southern California was studied to investigate the fate of the celadonite in a pedogenic weathering environment. Celadonite was found to be disseminated in the highly altered rock matrix with cristobalite, chalcedony, and stilbite. X-ray powder diffraction (XRD) showed the soil material to contain celadonite having a d(060) value of 1.510 Å, indicative of its dioctahedral nature. Very little smectite was detected in the parent material, whereas Fe-rich smectite was found to be abundant in the soil. The Fe-smectite and celadonite were identified as the sole components of the green-colored clay fraction (<2 µm) of all soil horizons. The soil clay showed a single d(060) value of 1.507 Å, indicating that the smectite was also dioctahedral and that its b-dimension was the same as that of the celadonite. Mössbauer spectroscopy showed that the chemical environments of Fe in the rock-matrix celadonite and in the smectite-rich soil clay were also nearly identical. These data strongly suggest a simple transformation of the celadonite to an Fe-rich smectite during soil formation.

Supporting evidence for this transformation was obtained by artificial weathering of celadonite, using sodium tetraphenyl boron to extract interlayer K. The intensity of the 001 XRD peak (at 10.1 Å) of celadonite was greatly reduced after the treatment and a peak at 14.4 Å, absent in the pattern of the untreated material, appeared. On glycolation of the sample, this peak expanded to 17.4 Å, similar to the behavior of the soil smectite. The alteration of celadonite to smectite is a simple transformation requiring only the loss of interlayer K. The transformation is apparently possible under present-day conditions, inasmuch as the erosional landscape position, shallow depth, and lack of significant horizonation indicate that the soil is very young.

Type
Research Article
Copyright
Copyright © 1988, The Clay Minerals Society

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

Abudelgawad, G., Page, A. L. and Lund, L. J., 1975 Chemical weathering of glauconite Soil Sci. Soc. Amer. Proc. 39 567571.CrossRefGoogle Scholar
Bailey, S. W., Brindley, G. W., Kodama, H. and Martin, R. T., 1979 Report of the Clay Minerals Society Nomenclature Committee for 1977 and 1978 Clays & Clay Minerals 27 238239.CrossRefGoogle Scholar
Beutelspacher, H. and Van der Marel, H. W., 1968 Atlas of Electron Microscopy of Clay Minerals and Their Admixtures Amsterdam Elsevier.CrossRefGoogle Scholar
Coey, J. M. D., 1980 Clay minerals and their transformations studied with nuclear techniques: The contribution of Mössbauer spectroscopy At. Energy Rev. 18 73124.Google Scholar
Daynyak, L. G. and Drits, V. A., 1987 Interpretation of Mössbauer spectra of nontronite, celadonite, and glauconite Clays & Clay Minerals 35 363372.CrossRefGoogle Scholar
Fanning, D. S., Keramidas, V. Z., Dixon, J. B. and Weed, S. B., 1977 Micas Minerals in Soil Environments Wisconsin Soil Sci. Soc. Amer., Madison 195258.Google Scholar
Heller-Kallai, L., 1982 Mössbauer studies of synthetic and natural micas on the polylithionite-siderophyllite join: Comments Phys. Chem. Minerals 8 98.CrossRefGoogle Scholar
Heller-Kallai, L. and Rozenson, J., 1980 Dehydroxylation of dioctahedral phyllosilicates Clays & Clay Minerals 28 355368.CrossRefGoogle Scholar
Heller-Kallai, L. and Rozenson, J., 1981 The use of Mössbauer spectroscopy of iron in clay mineralogy Phys. Chem. Minerals 7 223238.CrossRefGoogle Scholar
Hendricks, S. B. and Ross, C. S., 1941 Chemical composition and genesis of glauconite and celadonite Amer. Mineral. 26 683709.Google Scholar
Jackson, M. L., 1979 Soil Chemical Analysis—Advanced Course 2nd ed. Wisconsin Dept. Soil Science, Univ. Wisconsin, Madison.Google Scholar
Kimbara, K. and Shimoda, S., 1973 A ferric celadonite in amygdules ofdolerite at Taiheican, Akita Prefecture, Japan Clay Sci. 4 143150.Google Scholar
Kittrick, J. A. and Hope, E. W., 1963 A procedure for the particle-size separation of soils for X-ray diffraction analysis Soil Sci. 96 318325.CrossRefGoogle Scholar
Levillain, C., Marel, P. and Menil, F., 1982 Mössbauer studies of synthetic and natural micas on the polylithionite-siderophyllitejoin. Reply Phys. Chem. Minerals 8 99100.CrossRefGoogle Scholar
Loomis, D. P., 1984 Miocene stratigraphy and tectonic evolution of the El Paso Basin, California North Carolina Univ. North Carolina, Chapel Hill.Google Scholar
Munsell Color Company, 1970 Munsell Book of Color Maryland Munsell Color Company, Inc., Baltimore.Google Scholar
Odom, E. and Bailey, S. W., 1984 Glauconite and celadonite minerals Reviews in Mineralogy, Micas, Vol. 13 Washington D.C. Mineralogical Society of America 545572.Google Scholar
Pemberton, H. E., 1983 Minerals of California New York van Nostrand Reinhold Company 430.Google Scholar
Robert, M., 1973 The experimental transformation of mica toward smectite; relative importance of total charge and tetrahedral substitution Clays & Clay Minerals 21 167174.CrossRefGoogle Scholar
Scott, A. D., 1968 Effect of particle size on interlayer potassium exchange in mica Trans. 9th Int. Congr. Soil Sci., Adelaide, Vol. 2, 1968 Adelaide Int. Soc. Soil Sci. and Angus and Robertson 649660.Google Scholar
Scott, A. D., Hunziker, R. R. and Hanway, J. T., 1960 Chemical extraction of potassium from soils and micaceous minerals with solutions containing sodium tetraphenyl boron. I. Preliminary experiments Soil Sci. Soc. Amer. Proc. 24 191194.CrossRefGoogle Scholar
Scott, A. D. and Reed, M. G., 1962 Chemical extractions of potassium from soils and micaceous minerals with solutions containing sodium tetraphenol boron. II. Biotite Soil Sci. Soc. Amer. Proc. 26 4145.CrossRefGoogle Scholar
Scott, A. D., Reed, M. G., Bradley, W. F. and Bailey, S. W., 1966 Expansion of potassium-depleted muscovite Clays and Clay Minerals, Proc. 13th Natl. Conf, Madison, Wisconsin, 1964 New York Pergamon Press 247261.Google Scholar
Scott, A. D. and Smith, S. J., 1966 Susceptibility of interlayer potassium in micas to exchange with sodium Clays and Clay Minerals, Proc. 14th Natl. Conf, Berkeley, California, 1965 New York Pergamon Press 6981.Google Scholar
Soil Conservation Service, 1984 Procedures for collecting soil samples and methods of analysis for soil survey Soil Survey Investigations Report No. 1 Washington D.C. U.S. Dept. Agriculture.Google Scholar
Theisen, A. A. and Harward, M. E., 1962 A paste method for preparation of slides for clay mineral identification by X-ray diffraction Soil Sci. Soc. Amer. Proc. 26 9091.CrossRefGoogle Scholar
Wise, W. S. and Eugster, H. D., 1964 Celadonite: Synthesis, thermal stability, and occurrence Amer. Mineral. 49 10311083.Google Scholar