Skip to main content
Log in

Dehydroxylation, Rehydroxylation, and Stability of Kaolinite

  • Published:
Clays and Clay Minerals

Abstract

From hydrothermal experiments three pressure-temperature-time curves have been refined for the system Al2O3−SiO2−H2O and reversal temperatures established for two of the principal reactions involving kaolinite. The temperatures of three isobaric invariant points enable the Gibbs free energy of formation of diaspore and pyrophyllite to be refined and the stability field of kaolinite to be calculated. The maximal temperature of stable kaolinite decreases from 296°C at 2 kb water pressure to 284°C at water’s liquid/vapor pressure, and decreases rapidly at lower pressures. On an isobaric plot of [H4SiO4] vs. °K-1, kaolinite has a wedge-shaped stability field which broadens toward lower temperature to include much of the [H4SiO4] range of near-surface environments. If [H4SiO4] is above kaolinite’s stability field and the temperature is < 100°C, halloysite forms rather than pyrophyllite, an uncommon pedogenic mineral. Pyrophyllite forms readily instead of kaolinite above 150°C if [H4SiO4] is controlled by cristobalite or noncrystalline silica.

Kaolinite and a common precursor, halloysite, are characteristic products of weathering and hydro-thermal alteration. In sediments, relatively little halloysite has survived due to its low dehydration temperature and instability at low water pressure, but kaolinite commonly has survived since the Devonian Period. In buried sediments, the water pressure and [H4SiO4] requisite for stable kaolinite generally are maintained. In oxidized sediments and in pyritic reduced sediments, kaolinite commonly has survived, but where alkalies, alkaline earths, or aqueous iron has concentrated in the pore fluid, kaolinite has tended to transform to illite, zeolites, berthierine, or other minerals.

Резюме

На основе гидротермальных экспериментов были усовершенствованы три кривые давление-температура-время для системы Al2O3-SiO2-H2O и были определены реверсные температуры для двух из числа основных реакций, включающих каолинит. Величины температуры трех изобарных инвариантных точек позволили усовершенствовать величину свободной энергии Гиббса образования диаспора и пирофиллита, а также рассчитать поле стабильности каолинитов. Максимальная температура стабильного каолинита уменьшается от 296°С при давлении воды 2 кбар до 284°С при давлении жидкость/пар (для воды) и уменьшается быстро при низших давлениях. На изобарной кривой зависимости [H4SiO4] от °K-1, каолинит имеет клинообразное поле стабильности, которое расширяется по направлению к низшим температурам, чтобы включить большую часть [H4SiO4] области близких к поверхности сред. Если [H4SiO4] больше, чем для поля стабильности каолинита и температура < 100°С, галлуазит образуется вместо пирофиллита, необычного педогенического материала. Пирофиллит легко образуется вместо каолинита при температуре свыше 150°С, если [H4SiO4] контролируется кристобалитом или некристаллическим кремнеземом.

Каолинит и обычний предшественник, галлуазит, являются характерными продуктами выветривания и гидротермальных изменений пород. В осадочных отложениях сохранилось сравнительно небольшое количество галлуазита вследствие его низкой температуры дегидратации и нестабильности при низких давлениях воды, тогда как каолинит обычно сохраняется со времени девонского периода. В захороненных осадочных отложениях необходимые для стабильного каолинита давление воды и количество [H4SiO4] в основном поддерживаются. В окисленных отложениях и в отложениях с уменьшенным количеством пирита каолинит обычно сохраняется, но каолинит стремится видоизмениться в иллит, цеолит, бертьерин или другие минералы там, где в жидкости пор сосредотачиваются щелочи, щелочные почвы или осадочное железо. [E.G.]

Resümee

Aus hydrothermalen Experimenten wurden drei Druck-Temperatur-Zeit-Kurven für das System Al2O3-SiO2-H2O bestimmt, und die Temperaturen für zwei der wichtigsten Kaolinitreaktionen gewonnen. Die Temperaturen von drei isobar invarianten Punkten ermöglichen die Bestimmung der Gibbs’schen Freien Energie für die Bildung von Diaspor und Pyrophyllit und die Berechnung des Stabilitätsfeldes von Kaolinit. Die maximale Temperatur für stabilen Kaolinit nimmt von 296°C bei 2 kBar Wasserdampfdruck auf 284°C bei gesättigtem Wasserdampfdruck ab und verringert sich sehr schnell bei niedrigeren Drucken. Auf einem isobaren Diagramm, in dem [H4SiO4] gegen °K-1 aufgetragen ist, hat Kaolinit ein keilförmiges Stabilitätsfeld, das sich gegen niedrigere Temperaturen hin verbreitert, um viel von [H4SiO4]-Bereich der Oberflächenzone mit einzuschließen. Wenn [H4SiO4] über dem Kaolinitstabilitätsfeld liegt, und die Temperatur unter 100°C ist, dann bildet sich eher Halloysit als Pyrophyllit, ein unübliches Bodenmineral. Pyrophyllit bildet sich sehr leicht anstelle von Kaolinit bei Temperaturen über 150°C,wenn [H4SiO4] durch Cristobalit oder nichtkristallisiertes SiO2 kontrolliert wird.

Kaolinit und eine häufige Übergangsphase, Halloysit, sind typische Produkte der Verwitterung und hydrothermalen Umwandlung. In Sedimenten ist relativ wenig Halloysit aufgrund seiner niedrigen Dehydratationstemperatur und seiner Instabilität bei niedrigem H2O-Druck zu finden, während Kaolinit im allgemeinen seit dem Devon überlebt hat. In Versenkungssedimenten bleiben der für stabilen Kaolinit geforderte H2O-Druck und die notwendige [H4SiO4]-Aktivität im allgemeinen erhalten. In oxidierten Sedimenten und in pyritisch reduzierten Sedimenten bleibt Kaolinit gewöhnlich erhalten. Wenn jedoch Alkalien, Erdalkalien oder hydratisiertes Eisen in den Porenlösungen konzentriert sind, dann wandelt sich Kaolinit leicht in Illit, Zeolithe, Berthierit und anderen Minerale um. [U.W.]

Résumé

A partir d’expériences hydrothermiques, 3 courbes pression-température-temps ont été rafinées pour le système Al2O3-SiO2-H2O et des températures de revers ont été établies pour deux des réactions principales impliquant la kaolinite. Les températures de trois points invariants isobariques permet le rafinement de l’énergie libre de Gibbs de formation de la diaspore et de la pyrophyllite et le calcul du champ de stabilité de la kaolinite. La température maximale de kaolinite stable décroit de 296°C a 2 kb de pression d’eau a 284°C à la pression liquide/vapeur d’eau, et décroit rapidement à des pressions plus basses. Sur un diagramme isobarique de [H4SiO4] vs. °K-1, la kaolinite a un champ de stabilité éffilé à trois coins qui s’élargit vers la température plus basse pour inclure une grande partie de la gamme [H4SiO4] d’environements proches de la surface. Si [H4SiO4] est au delà du champ de stabilité de la kaolinite et la température est < 100°C, l’halloysite est formée plutôt que la pyrophyllite, un minéral pédogénique peu commun. La pyrophyllite est formée promptement à la place de la kaolinite au delá de 150°C si [H4SiO4] est contrôlée par la cristobalite ou par la silice non cristalline.

La kaolinite et un précurseur commun, l’halloysite, sont des produits caractéristiques de l’altération à l’air et hydrothermique. Dans des sédiments, relativement peu d’halloysite a survécu à cause de sa température de déshydratation basse et de son instabilité à de basses pressions d’eau, mais la kaolinite a communément survécu depuis la période dévonienne. Dans des sédiments ensevelis, la pression d’eau et l’[H4SiO4] nécéssaires pour la kaolinite stable sont généralement maintenues. Dans des sédiments oxidés et dans des sédiments pyritiques réduits, la kaolinite a communément survécu, mais là où des alkalins, des terres alkalines, ou du fer aqueux a été concentré dans les fluides de pores, la kaolinite a eu tendance à se transformer en illite, zéolite, berthierine ou en d’autres minéraux. [D.J.]

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Achar, N. N., Brindley, G. W., and Sharp, J. H. (1966) Kinetics and mechanism of dehydroxylation processes, III. Applications and limitations of dynamic methods: in Broc. Int. Clay Conf. Jerusalem, 1966, 1, L. Heller and A. Weiss, eds., Israel Prog. Sci. Transi., Jerusalem, Israel, 67–73.

    Google Scholar 

  • Anderson, G. M. and Burnham, C. W. (1965) The solubility of quartz in supercritical water: Amer. J. Sci. 263, 494–511.

    Article  Google Scholar 

  • Aramaki, Shigeo and Roy, Rustu. (1963) A new polymorph of Al2SiO5 and further studies in the system Al2O3-SiO2-H2O: Amer. Mineral. 48, 1322–1347.

    Google Scholar 

  • Barany, R. and Kelley, K. K. (1961) Heats and free energies of formation of gibbsite, kaolinite, halloysite, and dickite: U.S. Bur. Mines Rept. Inv. 5825, 1–13.

    Google Scholar 

  • Bhattacharyya, D. P. (1983) Origin of berthierine in ironstones: Clays & Clay Minerals 31, 173–182.

    Article  Google Scholar 

  • Bricker, O. P., Godfrey, A. E., and Cleaves, E. T. (1968) Mineral-water interaction during the chemical weathering of silicates: in Trace Inorganics in Water, R. F. Gould, ed., Advances in Chem. Ser., 73, Amer. Chem. Soc, Washington, D.C., 128–142.

    Chapter  Google Scholar 

  • Busenberg, Eurybiade. (1978) The product of the interaction of feldspars with aqueous solutions at 25°C: Geochim. Cosmochim. Acta 42, 1679–1686.

    Article  Google Scholar 

  • Carr, R. M. and Fyfe, W. S. (1960) Synthesis fields of some aluminum silicates: Geochim. Cosmochim. Acta 21, 99–109.

    Article  Google Scholar 

  • Crerar, D. A. and Anderson, G. M. (1971) Solubility and solvation reactions of quartz in dilute hydrothermal solutions: Chem. Geology 8, 107–122.

    Article  Google Scholar 

  • Day, H. D. and Kumin, H. J. (1980) Thermodynamic analysis of the aluminum silicate triple point: Amer. J. Sci. 280, 265–287.

    Article  Google Scholar 

  • von Dietzel, A. and Dhekne, B. (1957) Über die Rehydratation von Metakaolin: Ber. Deut. Keram. Ges. 34, 366–377.

    Google Scholar 

  • Eberl, D. D. and Hower, J. (1975) Kaolinite synthesis: the role of the Si/Al and (alkali)/H+ ratio in hydrothermal systems: Clays & Clay Minerals 23, 301–309.

    Article  Google Scholar 

  • Fisher, J. R. and Zen, E-A. (1971) Thermochemical calculations from hydrothermal phase equilibrium data and the free energy of H2O: Amer. J. Sci. 270, 297–314.

    Article  Google Scholar 

  • Fyfe, W. S. and Hollander, M. A. (1964) Equilibrium dehydration of diaspore at low temperature: Amer. J. Sci. 262, 709–712.

    Article  Google Scholar 

  • Garrels, R. M. and Christ, C. L. (1965) Solutions, Minerals, and Equilibria: Harper and Row, New York, 450 pp.

    Google Scholar 

  • Haas, Herber. (1972) Diaspore-corundum equilibrium determined by epitaxis of diaspore on corundum: Amer. Mineral. 57, 1375–1385.

    Google Scholar 

  • Haas, Herbert and Holdaway, M. J. (1973) Equilibrium in the system Al2O3-SiO2-H2O involving the stability limits of pyrophyllite and thermodynamic data of pyrophyllite: Amer. J. Sci. 273, 449–464.

    Article  Google Scholar 

  • Hass, J. L., Jr. (1970) Fugacity of H2O from 0° to 350°C at the liquid-vapor equilibrium and at 1 atmosphere: Geochim. Cosmochim. Acta 34, 920–932.

    Article  Google Scholar 

  • Helgeson, H. (1969) Thermodynamics of hydrothermal systems at elevated temperatures and pressures: Amer. J. Sci. 267, 729–804.

    Article  Google Scholar 

  • Helgeson, H. C. and Kirkham, D. H. (1974) Theoretical prediction of the thermodynamic behavior of aqueous electrolytes at high pressures and temperatures: I. Summary of the thermodynamic/electrostatic properties of the solvent: Amer. J. Sci. 274, 1089–1198.

    Article  Google Scholar 

  • Hemingway, B. S., Robie, R. A., and Kittrick, J. A. (1978) Revised values for the Gibbs free energy of formation of Al(OH)4 aq., diaspore, boehmite, and bayerite at 298.15°K and 1 bar, the thermodynamic properties of kaolinite to 800°K and 1 bar, and the heats of solution of several gibbsite samples: Geochim. Cosmochim. Acta 42, 1533–1554.

    Article  Google Scholar 

  • Hemley, J. J. (1959) Some mineralogical equilibria in the system K2O-Al2O3-SiO2-H2O: Amer. J. Sci. 257, 241–270.

    Article  Google Scholar 

  • Hemley, J. J., Montoya, J. W., Marinenko, J. W., and Luce, R. W. (1980) Equilibria in the system Al2O3-SiO2-H2O and some general implications for alteration/mineralization processes: Econ. Geol. 75, 210–228.

    Article  Google Scholar 

  • Henley, R. W. (1973) Solubility of gold in hydrothermal chloride solutions: Chem. Geol. 11, 73–87.

    Article  Google Scholar 

  • Hill, R. D. (1953) The rehydration of fired clay and associated minerals: Trans. Brit. Ceram. Soc. 52, 589–613.

    Google Scholar 

  • Hill, R. D. (1955) 14Å spacing in kaolin minerals: Acta Crystallogr. 8, p. 120.

    Article  Google Scholar 

  • Hughes, I. R. (1966) Mineral changes of halloysite on drying: N.Z.J. Sci. 9, 103–113.

    Google Scholar 

  • Hurst, V. J. and Basio, N. J. (1975) Rio Capim kaolin deposits, Brazil: Econ. Geol. 70, 990–992.

    Article  Google Scholar 

  • Keller, W. D., Pickett, E. E., and Reesman, A. L. (1966) Elevated dehydroxylation temperature of the Keokuk geode kaolinite—a possible reference mineral: in Proc. Inter. Clay Conf. Jerusalem 1966, 1, L. Heller and A. Weiss, eds., Israel Prog. Sci. Transi., Jerusalem, Israel, 75–85.

    Google Scholar 

  • Keller, W. D. (1978) Kaolinization of feldspars as displayed in scanning electron micrographs: Geology 6, 184–188.

    Article  Google Scholar 

  • Kennedy, G. C. (1944) The hydrothermal solubility of silica: Econ. Geol. 39, 25–36.

    Article  Google Scholar 

  • Kittrick, J. A. (1969) Soil minerals in the Al2O3-SiO2-H2O system and a theory of their formation: Clays & Clay Minerals 17, 157–167.

    Article  Google Scholar 

  • Kremer, Thoma. (1983) SEM and XRD investigation of mineralogical transformations during weathering: M.S. thesis, Univ. Georgia, Athens, Georgia, 115 pp.

    Google Scholar 

  • La Iglesia, A. and Galan, E. (1975) Halloysite-kaolinite transformation at room temperature: Clays & Clay Minerals 23, 109–113.

    Article  Google Scholar 

  • Laubengayer, A. W. and Weisz, R. S. (1943) A hydrothermal study of equilibria in the system alumina-water: J. Amer. Chem. Soc. 65, 247–250.

    Article  Google Scholar 

  • Marshall, C. E. (1977) The Physical Chemistry and Mineralogy of Soils: Wiley-Interscience, New York, 313 pp.

    Google Scholar 

  • McKeague, J. A. and Cline, M. G. (1963) Silica in soils: Adv. Agron. 15, 339–396.

    Article  Google Scholar 

  • Minato, H. and Aoki, M. (1979) Rate of transformation of halloysite to metahalloysite under hydrothermal conditions: in Proc. Inter. Clay Conf., Oxford, 1978, M. M. Mortland and V. C Farmer, eds., Elsevier, Amsterdam, 619–627.

    Google Scholar 

  • Minato, Hideo; Kusakabee, Hirotatsu; and Inoue, Atsuyuk. (1982) Alteration reactions of halloysite under hydrothermal conditions with acidic solutions: in Proc. Int. Clay Conf, Bologna, Pavia, 1981, H. van Olphen and F. Veniale, eds., Elsevier, Amsterdam, 565–571.

    Google Scholar 

  • Morey, G. W., Fournier, R. O., and Rowe, J. J. (1962) The solubility of quartz in water in the temperature interval from 25°C to 300°C: Geochim. Cosmochim. Acta 26, 1029–1043.

    Article  Google Scholar 

  • Parham, W. E. (1969) Formation of halloysite from feldspar: low temperature, artificial weathering versus natural weathering: Clays & Clay Minerals 17, 13–22.

    Article  Google Scholar 

  • Reed, B. L. and Hemley, J. J. (1966) Occurrences of pyrophyllite in the Kekiktuk conglomerate, Brooks Range, Northeast Alaska: U.S. Geol. Surv. Prof Pap. 550-C, 162–166.

    Google Scholar 

  • Robie, R. A., Hemingway, B. S., and Fisher, J. R. (1979) Thermodynamic properties of minerals and related substances at 298.15°K and 1 bar (155 Pascals) pressure and at higher temperatures: U.S. Geol. Surv. Bull. 1452, 456 pp.

  • Roy, Rustum and Brindley, G. W. (1956) A study of the hydrothermal reconstitution of the kaolin minerals: in Clays and Clay Minerals, Proc. 4th Natl. Conf, State College, Pennsylvania, 1955, Ada Swineford, ed., Natl. Acad. Sci. Natl. Res. Counc. Publ. 456, Washington, D.C., 125–132.

    Google Scholar 

  • Roy, Rustum and Osborn, E. F. (1954) The system Al2O3-SiO2-H2O: Amer. Mineral. 39, 853–885.

    Google Scholar 

  • Saalfeld, H. (1955) The hydrothermal formation of clay minerals from metakaolin: Ber. Deut. Keram. Ges. 32, 150–152.

    Google Scholar 

  • Schachtschabel, P. (1930) Dehydration and rehydration of kaolin: Chem. Erde 4, 375–419.

    Google Scholar 

  • Siever, R. (1962) Silica solubility, 0°-200°C, and the diagenesis of siliceous sediments: J. Geol. 70, 127–150.

    Article  Google Scholar 

  • Van Nieuwenberg, C. J. and Pieters, H. A. J. (1929) Rehydration of metakaolin and the synthesis of kaolin: Ber. Deut. Keram. Ges. 10, 260–263.

    Google Scholar 

  • Velde, B. and Kornprobst, J. (1969) Stabilité des silicates d’alumine hydrates: Contrib. Mineral Petrol. 21, 63–74.

    Article  Google Scholar 

  • Walther, J. V. and Helgeson, H. C. (1977) Calculation of the thermodynamic properties of aqueous silica and the solubility of quartz and its polymorphs at high pressures and temperatures: Amer. J. Sci. 277, 1315–1351.

    Article  Google Scholar 

  • Wefers, Karl and Bell, G. M. (1972) Oxides and hydroxides of aluminum: Technical Paper No. 19, Alcoa Research Laboratories, Pittsburgh, Pennsylvania, 51 pp.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hurst, V.J., Kunkle, A.C. Dehydroxylation, Rehydroxylation, and Stability of Kaolinite. Clays Clay Miner. 33, 1–14 (1985). https://doi.org/10.1346/CCMN.1985.0330101

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1346/CCMN.1985.0330101

Key Words

Navigation