Hostname: page-component-5c6d5d7d68-vt8vv Total loading time: 0.001 Render date: 2024-08-18T14:39:09.912Z Has data issue: false hasContentIssue false

Relation of Porosity and Permeability to the Origin of Diaspore Clay

Published online by Cambridge University Press:  01 January 2024

Victor T. Allen*
Affiliation:
St. Louis University, USA

Abstract

The apparent porosity of flint, boehmite and diaspore clays bears no direct, consistent relation to the permeability of the same type of clay or to the alumina content of each type. The coefficient of permeability varies with the type and the structure of each flint clay. Diaspore clays have high permeability compared to flint clays, and the boehmite clays measured generally have a permeability intermediate between that of diaspore and that of flint clays. The apparent increase in permeability with increase in alumina content suggests that permeability has increased as silica was removed from flint clay to form boehmite and diaspore clays.

A vertical pipe of diaspore clay 1½ inches in diameter was collected by Fred Mertens of the Laclede Christy Company at the Shockley-Thompson diaspore pit at Belle, Missouri. The pipe contains 69.16 percent alumina compared with 55.30 percent alumina for the average of the walls a few inches from the pipe. Diaspore in the pipe has vertical banding indicating it was deposited along a vertical crack. The formation of cracks in brittle flint clay is an important process in increasing the permeability and providing openings along which ground water can migrate.

A restudy of boehmite and diaspore clays emphasizes not only a replacement of the colloform and oolitic structures of flint clays by diaspore and boehmite, a gradual alteration of flint clay to diaspore clay along fractures, and second-generation diaspore cutting early diaspore, but also the importance of secondary processes in the origin of diaspore and boehmite clays.

Type
Article
Copyright
Copyright © The Clay Minerals Society 1954

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

Allen, Victor T. (1935) Mineral composition and origin of Missouri flint and diaspore clays: Mo. Geol. Survey, 58th Bienn. Rept., App. IV, pp, 1-24.Google Scholar
Bretz, J. Harlan (1950) Origin of filled sink-structures and circle deposits of Missouri: Geol. Soc. America Bull., vol, 61, pp. 789834.CrossRefGoogle Scholar
Bridge, Josiah (1952) Correlation of aluminum hydroxide and age of bauxite deposits: Problems of Clay and Laterite Genesis, Symposium, Am. Inst. Min. Met. Eng., pp. 212213.Google Scholar
Clark, F. W. (1924) Artificial coals: Data of Geochemistry, U.S. Geol. Survey Bull. 770, pp. 778781.Google Scholar
Clark, F. W. (1924) Analyses of river waters: U.S. Geol. Survey Bull. 770, pp. 7296.Google Scholar
Gilboy, Glennon (1933) Soil mechanics research: Paper No. 1832, Am. Soc. Civil Eng. Trans., vol. 98, pp. 223226.Google Scholar
Keller, W. D. (1952) Observations on the origin of Missouri high-alumina clays: Problems of Clay and Laterite Genesis, Symposium, Am. Inst. Min. Met. Eng., pp. 115134.Google Scholar
Laubengayer, A. W., and Weisz, R. S. (1943) A hydrothermal study of equilibria in the system alumina-water: Am. Chem, Soc. Jour., vol. 65, pp. 247250.CrossRefGoogle Scholar
Ries, H. (1927) Clays: Their occurrence, properties and uses, with special reference to those of the United States and Canada: 3d ed., New York, John Wiley & Sons, Inc., p. 90.Google Scholar
Roy, Rustum, and Osborn, E. F. (1952) Studies in the system alumina-silica-water: Problems of Clay and Laterite Genesis, Symposium, Am. Inst. Min. Met. Eng., pp. 7680.Google Scholar
Zapffe, Carl (1944) Memorandum report on iron ores of the Cle Elum district: Washington, Wash. Dept. Cons. Dev., Div. Mines Min., Invest. No. 15, pp. 127.Google Scholar