Hostname: page-component-848d4c4894-cjp7w Total loading time: 0 Render date: 2024-07-04T21:38:27.189Z Has data issue: false hasContentIssue false

Kaolinite, Smectite, and K-Rectorite in Bentonites: Relation to Coal Rank at Tulameen, British Columbia

Published online by Cambridge University Press:  01 July 2024

D. R. Pevear
Affiliation:
Department of Geology, Western Washington University, Bellingham, Washington, 98225
V. E. Williams
Affiliation:
Department of Geology, University of British Columbia, Vancouver, British Columbia V6T 1W5, Canada
G. E. Mustoe
Affiliation:
Department of Geology, Western Washington University, Bellingham, Washington, 98225

Abstract

The Tulameen coal field is part of an Eocene nonmarine basin which received extensive volcaniclastic sediments due to its location within an active magmatic arc. Bentonite partings in the coal originally consisted of glassy rhyolitic tephra with phenocrysts of sanidine, biotite, and quartz. During the initial alteration, which took place within the swamp or shortly after burial, glass was transformed to either smectite-cristobalite-clinoptilolite or to smectite-kaolinite. The formation of kaolinite depended on the degree of leaching of silica and alkalies in the swamp environment. Some beds are nearly 100% kaolinite and can be designated as tonsteins. The smectite shows no evidence of interlayering; the kaolinite is well ordered. During alteration, sodium, originally a component of the glass, was lost from the system.

A later thermal event, which affected only the southern part of the basin, metamorphosed the smectite to a regularly interstratified illite/smectite with 55% illite layers and rectorite-type superlattice (IS-type). The source of potassium was dissolution of sanidine. Vitrinite reflectance measurements of the coal suggest that the smectite was stable to 145–160°C, at which temperature it transformed to K-rectorite.

The absence of randomly interstratified intermediates, even in beds rich in potassium, suggests that the transformation of smectite to K-rectorite was controlled by a steep thermal gradient possibly resulting from local magmatism or circulating geothermal fluids.

Резюме

Резюме

Туламинский угольный бассейн является частью эоценового континентального бассейна, в который поступило большое количество вулканокластических отложений из-за его расположения в активном магматическом поясе. Бентонитовые прослои в угле первоначально состояли из стекловидной виолитовой тефры с фенокристаллами санидина, биотита, и кварца. Во время первоначального изменения, которое происходило в болотных условиях или вскоре после захоронения, стекло преобразовалось в смектит-кристобалит-клиноптилолит или в смектит-као-линит. Образование каолинита зависело от степени выщелачивания кремнезема и шелочей в болотной среде. Некоторые пласты состоят почти на 100% из каолинита и могут быть определены как тонштейн. Смектит является бейделлитовым и не проявляет признаков переслаивания; каолинит хорошо упорядочен. Во время изменения натрий, который первоначально входил в состав стекла, был удален из системы.

Позднее какое-то термическое событие, которое повлияло только на южную часть бассейна, превратило смектит в равномерно переслаивающийся иллит-смектит с 55% иллитовых слоев и сверхструктурой ректоритового типа (тип 18). Источником калия служило растворение санидина. Измерения витринитовых отражений угля показывают, что смектит был стойким до температур 145-160°С, при которых он превращался в К-ректорит. Отсутствие беспорядочно переслаивающихся промежуточных звеньев даже в богатых калием пластах указывает на то, что преобразование смектита в К-ректорит было обусловлено значительным термальным градиентом, возможно происшедшим из-за местного магматизма или циркуляции геотермальных растворов. [N.R.]

Resümee

Resümee

Das Kohlengebiet von Tulameen ist Teil eines eozänen, nichtmarinen Beckens, das aufgrund seiner Lage in einem aktiven Magmenbogen sehr viele vulkanoklastische Sedimente enthält. Die Bentonitanteile in der Kohle waren ursprünglich glasige, rhyolithische Tephra mit Einsprenglingen von Sanidin, Biotit, und Quarz. Zu Beginn der Umwandlung, die im Schlamm oder kurz nach der Überdeckung stattfand, wurde das Glas entweder in Smektit-Cristobalit-Klinoptilolit oder in Smektit-Kaolinit umgewandelt. Die Bildung von Kaolinit hängt vom Auslaugungsgrad des SiO2 und der Alkalien im Schlamm ab. Einige Lagen bestehen aus nahezu 100% Kaolinit und können als Tonstein bezeichnet werden. Der Smektit ist beidellitisch und zeigt keine Anzeichen von Zwischenlagen; der Kaolinit ist gut geordnet. Während der Umwandlung ging Natrium, das ursprünglich eine Komponente des Glases war, aus dem System verloren.

Ein späteres thermales Ereignis, das nur den südlichen Teil des Beckens betraf, wandelte den Smektit in eine reguläre Wechsellagerung Illit-Smektit um, mit 55% Illitlagen und einem Gitter vom Rektorittyp (IS-Typ). Das Kalium kam von der Auflösung des Sanidins. Messungen des Reflexionsvermögens am Vitrinit der Kohle deuten darauf hin, daß der Smektit bis zur Temperatur zu 145–160°C stabil war, bei der er in K-Rektorit umgewandelt wurde. Das Fehlen von Übergangsphasen mit unregelmäßigen Wechsellagerungsstrukturen, selbst in den Kalium-reichen Lagen, deutet darauf hin, daß die Umwandlung von Smektit in K-Rectorit durch einen steilen thermischen Gradienten bestimmt wurde, der möglicherweise mit einem lokalen Magnetismus oder mit zirkulierenden geothermalen Lösungen zusammenhängt. [U.W.]

Résumé

Résumé

Le champ charbonnier Tulameen fait partie d'un bassin non marin éocien qui recevait des sédiments volcaniclastiques extensifs dûs à son emplacement duns un arc magmatique actif. Des délitations de bentonite dans le charbon consistaient originalement de tephra vitreux rhyolitiques avec des phénocrystes de sanidine, de biotite, et de quartz. Pendant la période initiale d'altération, qui s'est deroulée dans le marais ou peu après enterrement, le verre a été transformé soit en de la smectite-cristobalite-clinoptilolite, soit en de la smectite-kaolinite. La formation de kaolinite dépendait du degré de lessivage de la silice et des alcalins dan l'environement marécageux. Certains lits sont presque 100% kaolinite et pourraient être designés tonstein. La smectite est beidellitique et ne montre aucune évidence d'interstratification, la kaolinite est bien ordonnée. Pendant l'altération, le système a perdu le sodium, originalement un composé du verre.

Un évenement thermique subséquent, qui n'a affecté que la partie sud du bassin, a metamorphosé la smectite en une illite/smectite régulièrement interstratifiée avec 55% de couches d'illite et un super-réseau du type rectorite (type IS). La source du potassium était la dissolution de sanidine. Des mesures de réflectance de vitrinite du charbon suggèrent que la smectite était stable jusqu’à 145°–160°C, température à laquelle elle a été transformée en rectorite-K. L'absence d'intermédiaires interstratifiés au hasard suggère que la transformation de smectite en rectorite-K était controllée par un gradient thermique résultant possiblement d'un magmatisme local ou de fluides géothermiques circulants. [D.J.]

Type
Research Article
Copyright
Copyright © Clay Minerals Society 1980

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

Bohor, B. F. Pollastro, R. M. and Phillips, R. E., (1978) Mineralogical evidence for the volcanic origin of kaolinitic partings (tonstein) in Upper Cretaceous and Tertiary coals of the Rocky Mountain Region Program and Abstracts, 15th Annual Meeting, Clay Minerals Soc Indiana Bloomington 47.Google Scholar
Bostick, N. H., (1979) Microscopic measurement of the level of catagenesis of solid organic matter in sedimentary rocks to aid exploration for petroleum and to determine former burial temperatures—A review Aspects of Diagenesis 26 1743.CrossRefGoogle Scholar
Bramlette, M. N. and Posnjak, E., (1933) Zeolite alteration of pyroclastics Amer. Mineral. 18 492493.Google Scholar
Castano, J. R. and Sparks, D. M., (1974) Interpretation of vitrinite reflectance measurements in sedimentary rocks and determination of burial history using reflectance and authigenic minerals Carbonaceous Materials as Indicators of Metamorphism 153 3153.CrossRefGoogle Scholar
Davies, D. Almon, W. R. Bonis, S. B. and Hunter, B. E., (1979) Deposition and diagenesis of Tertiary-Holocene vol-caniclastics, Guatemala Aspects of Diagenesis 26 281306.CrossRefGoogle Scholar
Davis, G. A., (1977) Tectonic evolution of the Pacific Northwest from Precambrian to present WNP-I/4, PSAR, Amendment 23, Subappendix 2RC .Google Scholar
Deffeyes, K. S., (1959) Zeolites in sedimentary rocks J. Sediment. Petrology 29 602609.Google Scholar
Dickinson, W. R., Armentrout, J. M. Cole, M. R. and TerBest, H Jr., (1979) Cenozoic plate tectonic setting of the Cordilleran region in the United States Cenozoic Paleogeography of the Western United States 113.Google Scholar
Divis, A. F. and McKenzie, J. A., (1975) Experimental au-thigenesis of phyllosilicates from feldspathic sands Sedi-mentology 22 147155.Google Scholar
Dunoyer de Segonzac, G., (1970) The transformation of clay minerals during diagenesis and low grade metamorphism: A review Sedimentology 15 281346.CrossRefGoogle Scholar
Eberl, D. D., (1978) Reaction series for dioctahedral smectites Clays & Clay Minerals 26 327340.CrossRefGoogle Scholar
Eberl, D. D., (1978) The reaction of montmorillonite to mixed-layer clay. The effect of interlay er alkali and alkaline earth cations Geochim. Cosmochim. Acta 42 17.CrossRefGoogle Scholar
Hay, R. L., Sand, L. B. and Mumpton, F. A., (1978) Geologic occurrence of zeolites Natural Zeolites: Occurrence, Properties, Use Elmsford, N.Y. Pergamon Press 135145.Google Scholar
Hein, J. R. Scholl, D. W. Barron, J. A. Jones, M. G. and Miller, J., (1978) Diagenesis of late Cenozoic diatomaceous deposits and formation of the bottom simulating reflector in the southern Bering Sea Sedimentology 25 155181.CrossRefGoogle Scholar
Helgeson, H. C. Brown, T. H. and Leeper, R. H., (1969) Handbook of Theoretical Activity Diagrams Depicting Chemical Equilibria in Geologic Systems Involving an Aqueous Phase at One Atm. and 0° to 300°C Calif Freeman, Cooper and Company, San Francisco.Google Scholar
Henderson, J. H. Jackson, M. L. Syers, J. K. Clayton, R. N. and Rex, R. W., (1971) Cristobalite authigenic origin in relation to montmorillonite and quartz origin in bentonites Clays & Clay Minerals 19 229238.CrossRefGoogle Scholar
Hills, L. V. and Baadsgaard, H., (1967) Potassium-argon dating of some lower Tertiary strata in British Columbia Bull. Can. Petrol. Geol. 15 138149.Google Scholar
Hoffman, J. and Hower, J., (1979) Clay mineral assemblages as low grade metamorphic geothermometers: Application to the thrust faulted disturbed belt of Montana, U.S.A. Aspects of Diagenesis 26 5579.CrossRefGoogle Scholar
Horton, D. G., (1978) Clay Mineralogy and Origin of the Huntingdon Fire Clays on Canadian Sumas Mountain, Southwest British Columbia Bellingham, Washington M.Sc. thesis, Western Washington University.Google Scholar
Hower, J. Eslinger, E. V. Hower, M. E. and Perry, E. A., (1976) Mechanism of burial metamorphism of argillaceous sediment: 1. Mineralogical and chemical evidence Geol. Soc. Amer. Bull. 87 725737.2.0.CO;2>CrossRefGoogle Scholar
Iijima, A., Sand, L. B. and Mumpton, F. A., (1978) Geological occurrences of zeolite in marine environments Natural Zeolites: Occurrence, Properties, Use Elmsford, N.Y. Pergamon Press 175198.Google Scholar
Jackson, M. L., (1974) Soil Chemical Analysis—Advanced Course. 2nd Univ. of Wisconsin, Madison, Wise 9th printing: Published by the author, Dept. of Soil Science.Google Scholar
Jones, J. B. and Segnit, E. R., (1971) The nature of opal I. Nomenclature and constituent phases J. Geol. Soc. Aust. 18 5767.CrossRefGoogle Scholar
Keith, T. E. White, D. E. and Beeson, M. H., (1978) Hydrothermal alteration and self-sealing in Y-7 and Y-8 drill holes in northern part of Upper Geyser Basin, Yellowstone NationalPark, Wyoming U.S. Geol. Surv. Prof. Pap. 1054 126.Google Scholar
Kisch, H. J., Schenk, P. A. and Havenaar, I., (1968) Coal-rank and burial-metamorphic mineral facies Advances in Organic Geochemistry Elmsford, N.Y. Pergamon Press 407425.Google Scholar
Loughnan, F. C., (1978) Flint clays, tonsteins and the kaolin-ite clayrock facies Clay Miner. 13 387400.CrossRefGoogle Scholar
MacEwan, D. M., (1956) Fourier transform methods for studying scattering from lamellar systems. I. A direct method for analysing interstratifled mixtures Kolloid Z. 149 96108.CrossRefGoogle Scholar
Mueller, R. F. and Saxena, S. K., (1977) Chemical Petrology New York Springer-Verlag.CrossRefGoogle Scholar
Muffler, P. and White, D. E., (1969) Active metamorphism of Upper Cenozoic sediments in the Salton Sea Geothermal Field and the Salton Trough, southeastern California Geol. Soc. Amer. Bull. 80 157182.CrossRefGoogle Scholar
Mumpton, F. A., (1960) Clinoptilolite redefined Amer. Mineral. 45 351369.Google Scholar
Murata, K. J. and Whiteley, K. R., (1973) Zeolites in the Miocene Briones Sandstone and related formations of the Central Coast Ranges, California J. Res. U.S. Geol. Surv. 1 255265.Google Scholar
Okulitch, A. V., Price, R. A., and Richards, T. A. (1977) A guide to the geology of the southern Canadian Cordillera: Geol. Assoc. Can-Mineral. Assoc. Can.-Soc. Econ. Geol. Field Trip Guidebook 8, 135 pp.Google Scholar
Perry, E. A. and Hower, J., (1970) Burial diagenesis in Gulf Coast pelitic sediments Clays & Clay Minerals 18 165177.CrossRefGoogle Scholar
Reynolds, R. C. Jr., (1967) Interstratifled clay systems: Calculation of the total one-dimensional diffraction function Amer. Mineral. 52 661672.Google Scholar
Reynolds, R. C. Jr. and Anderson, D. M., (1967) Cristobalite and clinoptilolite in bentonite beds of the Colville Group, northern Alaska J. Sediment. Petrology 37 966969.CrossRefGoogle Scholar
Reynolds, R. C. Jr. and Hower, J., (1970) The nature of in-terlayering in mixed-layer illite-montmorillonite Clays & Clay Minerals 18 2536.CrossRefGoogle Scholar
Sheppard, R. A. and Gude, A. J. 3rd (1969) Diagenesis of tuffs in the Barstow Formation, Mud Hills, San Bernardino County, California: U.S. Geol. Surv. Prof. Pap. 634, 36 pp.Google Scholar
Spears, D. A. and Kanaris-Sotiriou, R., (1979) Ageochemical and mineralogical investigation of some British and other European tonsteins Sedimentology 26 407425.CrossRefGoogle Scholar
Środoń, J., Mortland, M. M. and Farmer, V. C., (1979) Correlation between coal and clay diagenesis in the Carboniferous of the Upper Silesian Coal Basin International Clay Conference 1978 Amsterdam Elsevier 251260.Google Scholar
Środoń, J. (1980) Precise X-ray identification of illite/smec-tites: Clays & Clay Minerals 28, (in press).CrossRefGoogle Scholar
Stach, E. MacKowsky, M Th Teichmuller, M. Taylor, G. H. Chandra, D. and Teichmuller, R., (1975) Stach’s Textbook of Coal Petrology Berlin Gebruder Borntraeger.Google Scholar
Steiner, A., (1968) Clay minerals in hydrothermally altered rocks at Wairakei, New Zealand Clays & Clay Minerals 16 193213.CrossRefGoogle Scholar
Thompson, G. Fields, R. W. and Alt, D., (1978) Major Tertiary climate variations in the western United States as indicated by paleosol mineralogy and sedimentation patterns Program and Abstracts, 15th Annual Meeting, Clay Minerals Soc Indiana Bloomington 35.Google Scholar
Williams, V. E., (1978) Coal Petrology of the Tulameen Coalfield, South-Central British Columbia Bellingham, Washington M.Sc. thesis, Western Washington University.Google Scholar
Williams, V. E. and Ross, C. A., (1979) Depositional setting and coal petrology of Tulameen Coalfield, South-Central British Columbia Amer. Ass. Petrol. Geol. Bull. 63 20582069.Google Scholar
Wright, T. L., (1968) X-ray and optical study of alkali feldspar—Pt. 2 Amer. Mineral. 53 88104.Google Scholar