Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-10T19:48:38.454Z Has data issue: false hasContentIssue false

On the Thermodynamic Stability of Illite and I-S Minerals

Published online by Cambridge University Press:  01 January 2024

Stephen U. Aja*
Affiliation:
Department of Earth and Environmental Sciences, Brooklyn College and The Graduate Center, City University of New York, 2900, Bedford Avenue, Brooklyn, NY 11210-2889, USA
*
*E-mail address of corresponding author: suaja@brooklyn.cuny.edu

Abstract

A review of the models proffered to advance the notion of the metastability of illite shows that these models are not supported by the various data groups that have become available. Given that clay minerals are products of water–rock interactions, low-temperature hydrothermal experiments provide singular insights into their relative stabilities; such experiments with natural materials of contrasting pedigree (illites, sericites, muscovites, and chlorites) show that clay-mineral behaviors in low-temperature hydrothermal solutions are amenable to equilibrium thermodynamic conventions. The data from hydrothermal experiments coupled with data from geothermal fields indicate that muscovite is not a stable phase in the P-T-X range in which authigenic illite occurs; given that experimental data and field occurrence suggest that muscovite and illite have different P-T stability regimes, the continued use of muscovite as a proxy for illite in thermodynamic models is of questionable utility. Furthermore, morphometric studies of clays undergoing illitization show that crystal-size distributions exhibit log-normal patterns. Because log-normal distributions derive from maximum entropy effects, these crystal-size distributions may reflect the effects of entropy production during crystallization rather than kinetically driven Ostwald ripening of illitic phases; the small crystal size of clay minerals may derive from constraints imposed by the physicochemical conditions of their environments of formation. Presumably, irreversible thermodynamics provides the framework for a quantitative understanding of the evolution of complex clay minerals in space and time.

Type
Review
Copyright
Copyright © Clay Minerals Society 2020

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

Ahn, J. H., & Peacor, D. R. (1986). Transmission and analytical electron microscopy of the smectite-to-illite transition. Clays and Clay Minerals, 34, 165179.Google Scholar
Aja, S. U. (1989) A hydrothermal study of illite stability relationships between 25 and 250°C and Pv = PH2O. PhD thesis, Washington State University.Google Scholar
Aja, S. U. (1991). Illite equilibria in solutions: III. A reinterpretation of the data of Sass et al. (1987). Geochimica et Cosmochimica Acta, 55, 34313435.CrossRefGoogle Scholar
Aja, S. U. (1995). Thermodynamic properties of some 2: 1 layer clay minerals from solution-equilibration data. European Journal of Mineralogy, 7, 325333.CrossRefGoogle Scholar
Aja, S. U. (2002). The stability of Fe-Mg chlorites in hydrothermal solutions: II. Thermodynamic properties. Clays and Clay Minerals, 50, 591600.CrossRefGoogle Scholar
Aja, S. U., & Dyar, M. D. (2002). The stability of Fe-Mg chlorites in hydrothermal solutions I. Results of experimental investigations. Applied Geochemistry, 17, 12191239.CrossRefGoogle Scholar
Aja, S. U., & Rosenberg, P. E. (1992). The thermodynamic status of compositionally-variable clay minerals: A discussion. Clays and Clay Minerals, 40, 292299.CrossRefGoogle Scholar
Aja, S. U., & Rosenberg, P. E. (1996). The thermodynamic status of compositionally-complex clay minerals: Discussion of “Clay Mineral thermometry – A critical perspective”. Clays and Clay Minerals, 44, 560568.CrossRefGoogle Scholar
Aja, S. U., Rosenberg, P. E., & Kittrick, J. A. (1991a). Illite equilibria in solutions: I. Phase relationships in the system K2O-Al2O3-SiO2-H2O between 25 and 250°C. Geochimica et Cosmochimica Acta, 55, 13531364.CrossRefGoogle Scholar
Aja, S. U., Rosenberg, P. E., & Kittrick, J. A. (1991b). Illite equilibria in solutions: II. Phase relationships in the system K2O-Al2O3-MgO-SiO2-H2O. Geochimica et Cosmochimica Acta, 55, 13651374.CrossRefGoogle Scholar
Aja, S.U., & Small, J.S. (1999) The solubility of a low-Fe clinochlore between 25 and 175°C and Pv = PH2O European Journal of Mineralogy, 11, 829842.CrossRefGoogle Scholar
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, 517533.CrossRefGoogle Scholar
Anderson, G. M. (2002) Stable and metastable equilibrium: the third constraint. In Hellman, R., & Wood, S. A.. (eds), Water-Rock Interactions, Ore Deposits, and Environmental Geochemistry: a Tribute to David A. Crerar (pp. 181189). The Geochemical Society, Special Publications, 7.Google Scholar
Baldan, A. (2002). Review Progress in Ostwald ripening theories and their applications to nickel-base superalloys Part I: Ostwald ripening theories. Journal of Materials Science, 37, 21712202.CrossRefGoogle Scholar
Berrill, J. B., & Davis, R. O. (1980). Maximum entropy and the magnitude distribution. Bulletin of the Seismological Society of America, 70, 18231831.Google Scholar
Boles, J. R., & Franks, S. G. (1979). Clay diagenesis in Wilcox sandstones of Southwest Texas: implications of smectite diagenesis on sandstone cementation. Journal of Sedimentary Petrology, 49, 5570.Google Scholar
Brown, G., & Norrish, K. (1952). Hydrous micas. Mineralogical Magazine, 29, 929932.CrossRefGoogle Scholar
Chung, S.-Y., Kim, Y.-M., Kim, J.-G, & Kim, Y.-J. (2009). Multiphase transformation and Ostwald's rule of stages during crystallization of a metal phosphate. Nature Physics, 5, 6873.CrossRefGoogle Scholar
De Yoreo, J. J., & Vekilov, P. G. (2003). Principles of crystal nucleation and growth. Reviews in Mineralogy, 54, 5793.CrossRefGoogle Scholar
Dong, H., & Peacor, D. R. (1996). TEM observations of coherent stacking relations in smectite, IIS and illite of shales: evidence for MacEwan crystallites and dominance of 2M1 polytypism. Clays and Clay Minerals, 44, 257275.CrossRefGoogle Scholar
Dong, H., Peacor, D. R., & Freed, R. L. (1997). Phase relations among smectite, R1 illite-smectite, and illite. American Mineralogist, 82, 379391.CrossRefGoogle Scholar
Drits, V. A., Lindgreen, H., Sakharov, B. A., Jakobsen, H. J., Fallick, A. F., Salyn, A. I., Dainyk, I. G., Zviagina, B. B., & Barfod, D. N. (2007). Formation and transformation of mixed-layer minerals by Tertiary intrusives in Cretaceous mudstones, West Greenland. Clays and Clay Minerals, 55, 260283.CrossRefGoogle Scholar
Dubacq, D., Vidal, O., & Lewin, É. (2011). Atomistic investigation of the pyrophyllite substitution and implications on clay stability. American Mineralogist, 96, 241249.CrossRefGoogle 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., & Hower, J. (1977). The hydrothermal transformation of sodium and potassium smectite into mixed layer clay. Clays and Clay Minerals, 28, 161172.CrossRefGoogle 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, 914934.Google Scholar
Eberl, D. D., & Środoń, J. (1988). Ostwald ripening and interparticle diffraction effects for illite crystals. American Mineralogist, 73, 13351345.Google Scholar
Eberl, D. D., Środoń, J., Kralik, M., Taylor, B., & Peterman, Z. E. (1990). Ostwald ripening of clays and metamorphic minerals. Science, 248, 474477.CrossRefGoogle Scholar
Eberl, D. D., Drits, V. A., & Środoń, J. (1998). Deducing growth mechanisms for minerals from the shapes of crystal size distributions. American Journal of Science, 298, 499533.CrossRefGoogle Scholar
Essene, E. J., & Peacor, D. R. (1995). Clay mineral thermometry: A critical perspective. Clays and Clay Minerals, 43, 540553.CrossRefGoogle Scholar
Foscolos, A. E., & Kodama, H. (1974). Diagenesis of clay minerals from Lower Cretaceous shales of North Eastern British Columbia. Clays and Clay Minerals, 22, 319335.CrossRefGoogle Scholar
Gailhanou, H., Blanc, P., Rogez, J., Mikaelian, G., Kawaji, H., Olives, J., Amouric, M., Denoyel, R., Bourrelly, S., Montouillout, V., Viellard, P., Fialips, C. I., Michau, N., & Gaucher, E. C. (2012). Thermodynamic properties of illite, smectite and beidellite by calorimetric methods: Enthalpies of formation, heat capacities, entropies and Gibbs free energies of formation. Geochimica et Cosmochimica Acta, 89, 279391.CrossRefGoogle Scholar
Garrels, R. M., & Howard, P. (1957). Reactions of feldspar and mica with water at low temperature and pressure. Clays and Clay Minerals, 6, 6888.CrossRefGoogle Scholar
Gaudette, H. E. (1965). Illite from Fond du Lac County, Wisconsin. American Mineralogist, 50, 411417.Google Scholar
Gaudette, H. E., Eades, J. L., & Grim, R. E. (1966a). The nature of illites. Clays and Clay Minerals, 13, 3348.CrossRefGoogle Scholar
Gaudette, H. E., Grim, R. E., & Metzger, C. F. (1966b). Illite: a model based on the sorption behavior of cesium. American Mineralogist, 51, 16491656.Google Scholar
Grim, R. E., & Bradley, W. F. (1939). A unique clay from the Goose Lake, Illinois, area. Journal of American Ceramic Society, 22, 157164.CrossRefGoogle Scholar
Grim, R. E., Bray, R. H., & Bradley, W. F. (1937). The mica in argillaceous sediments. American Mineralogist, 22, 813829.Google Scholar
Gro$nDholm, T., & Annila, A. (2007). Natural distribution. Mathematical Biosciences, 210, 659667.CrossRefGoogle Scholar
Güven, N. (1972). Electron optical observations in Marblehead illite. Clays and Clay Minerals, 37, 111.Google Scholar
Hartman, P. (1973). Structure and morphology. In Hartman, P. (Ed.), Crystal Growth: an Introduction (pp. 358402). Amsterdam: North Holland Publishing Company.Google Scholar
Heling, D. (1974). Diagenetic alteration of smectite in argillaceous sediments of the Rhinegraben. Sedimentology, 21, 463472.CrossRefGoogle Scholar
Hemley, J. J. (1959). Some mineralogic equilibria in the system K2O-Al2O3-SiO2 -H2O. American Journal of Science, 57, 241270.CrossRefGoogle Scholar
Hower, J., & Mowatt, T. C. (1966). The mineralogy of illites and mixed-layer illite-montmorillonites. American Mineralogist, 51, 825854.Google Scholar
Hower, J., Eslinger, E. V., Hower, M. E., & Perry, E. A. (1976). Mechanism of burial metamorphism of argillaceous sediments: 1. Mineralogical and chemical evidence. Geological Society of America Bulletin, 87, 725737.2.0.CO;2>CrossRefGoogle Scholar
Inoue, A. (1983). Potassium fixation by clay minerals during hydro-thermal treatment. Clays and Clay Minerals, 31, 8191.CrossRefGoogle Scholar
Inoue, A., & Kitagawa, R. (1994). Morphological characteristics of illitic clay minerals from a hydrothermal system. American Mineralogist, 79, 700711.Google Scholar
Inoue, A., & Utada, M. (1983). Further investigations of a conversion series of dioctahedral mica/smectites in the Shinzan hydrothermal alteration area, northeast Japan. Clays and Clay Minerals, 31, 401412.CrossRefGoogle Scholar
Inoue, A., Minto, H., & Utada, M. (1978). Mineralogical properties and occurrence of illite/montmorillonite mixed layer minerals formed from Miocene volcanic glass in Waga-Omono district. Clay Science, 5, 123136.Google Scholar
Inoue, A., Kohyama, N., Kitagawa, R., & Watanabe, T. (1987). Chemical and morphological evidence for the conversion of smectite to illite. Clays and Clay Minerals, 35, 111120.CrossRefGoogle Scholar
Inoue, A., Velde, B., Meunier, A., & Touchard, G. (1988). Mechanism of illite formation during smectite-to-illite conversion in a hydrothermal system. American Mineralogist, 73, 13251334.Google Scholar
Jaisi, D. P., Eberl, D. D., Dong, H., & Kim, J. (2011). The formation of illite from nontronite by mesophilic and thermophilic bacterial reaction. Clays and Clay Minerals, 59, 2133.CrossRefGoogle Scholar
Jiang, W., Essene, E. J., & Peacor, D. R. (1990). Transmission electron microscopic study of coexisting pyrophyllite and muscovite: direct evidence for the metastability of illite. Clays and Clay Minerals, 38, 225240.CrossRefGoogle Scholar
Kittrick, J. A. (1984). Solubility measurements of phases in three illites. Clays and Clay Minerals, 32, 115124.CrossRefGoogle Scholar
Kim, J., & Peacor, D. R. (2002). Crystal-size distributions of clays during episodic diagenesis: The Salton Sea geothermal system. Clays and Clay Minerals, 50, 371380.CrossRefGoogle Scholar
Kim, J., Dong, H., Seabaugh, J., Newell, S., & Eberl, D. D. (2004). Role of microbes in the smectite-to-illite reaction. Science, 303, 830832.CrossRefGoogle ScholarPubMed
Kim, J., Dong, H., Yang, K., Park, H., Elliott, W. C., Spivack, A., Koo, T.-H., Kim, G., Morono, Y., Henkel, S., Inagaki, F., Zeng, Q., Hoshino, T., & Heuer, V. B. (2019). Naturally occurring, microbially induced smectite-to-illite reaction. Geology, 47, 535539.CrossRefGoogle Scholar
Kondepudi, D., & Prigogine, I. (1998). Modern Thermodynamics: From Heat Engines to Dissipative Structures. New York: John Wiley.Google Scholar
Kuila, U., & Prasad, M. (2013). Specific surface area and pore-size distribution in clays and shales. Geophysical Prospecting, 61, 341362.CrossRefGoogle Scholar
Kulik, D. A., & Aja, S. U. (1997). The hydrothermal stability of illite: implications of empirical correlations and Gibbs energy minimization. In Palmer, D. A. & Wesolowski, D. J. (Eds.), Proceedings of the Fifth International Symposium on Hydrothermal Reactions (pp. 288292). Gatlinburg, Tennessee, USA: ORNL.Google Scholar
Kuwahara, Y., & Uehara, S. (2008). AFM study on surface microtopography, morphology and crystal growth of hydrothermal illite in Izumiyama pottery stone from Arita, Saga Prefecture, Japan. The Open Mineralogy Journal, 2, 3437.CrossRefGoogle Scholar
Lanson, B., & Champion, D. (1991). The I/S-to-illite reaction in the late stage diagenesis. American Journal of Science, 291, 473506.CrossRefGoogle Scholar
Li, D., Nielsen, M. H., Lee, J. R. I., Frandsen, C., Banfield, J. F., & De Yoreo, J. J. (2012). Direction-specific interactions control crystal growth by oriented attachment. Science, 336, 10141018.CrossRefGoogle ScholarPubMed
Lifshitz, I. M., & Slyozov, V. V. (1961). The kinetics of precipitation from supersaturated solid solutions. Journal of Physics and Chemistry of Solids, 19, 3550.CrossRefGoogle Scholar
Lippmann, F. (1977). The solubility product of complex minerals, mixed-crystals and three-layer clay minerals. Neues Jahrbuch für Mineralogie – Abhandlungen, 130, 243263.Google Scholar
Lippmann, F. (1982). The thermodynamic status of clay minerals. In Olphen, H. van & Veniale, F. (Eds.), Proceedings of the 7th International Clay Conference Bologna, Pavia 1981 (pp. 475485). New York: Elsevier.Google Scholar
Lonker, S. W., Fitz Gerald, J. D., Hedenquist, J. W., & Walshe, J. (1990). Mineral-fluid interactions in the Broadlands-Ohaaki geothermal system, New Zealand. American Journal of Science, 290, 9951068.CrossRefGoogle Scholar
Loucks, R. R. (1991). The bound interlayer H2O content of potassic white micas: muscovite-hydromuscovite-hydropyrophyllite solutions. American Mineralogist, 76, 15631579.Google Scholar
Manning, D. A. C. (2003). Experimental studies of mineral occurrences. In Worden, R. H. & Morad, S. (Eds.), Clay Mineral Cements in Sandstones (pp. 177190). Oxford: International Association of Sedimentologists Special Publications 34, Blackwell Publishing company.Google Scholar
Mankin, C. J., & Dodd, C. G. (1963). Proposed reference illite from the Ouchita Mountains of Southeastern Oklahoma. Clays and Clay Minerals, 10, 372379.CrossRefGoogle Scholar
McDowell, D. S., & Elders, W. A. (1980). Authigenic layer silicate minerals in borehole Elmore 1, Salton Sea Geothermal Field, California, USA. Contributions to Mineralogy and Petrology, 74, 293310.CrossRefGoogle Scholar
Meunier, A. (2006). Why are clay minerals small? Clay Minerals, 41, 551566.CrossRefGoogle Scholar
Montoya, J. W., & Hemley, J. J. (1975). Activity relations and stabilities in alkali feldspar and mica alteration reactions. Economic Geology, 70, 577594.CrossRefGoogle Scholar
Morse, J. W., & Casey, W. H. (1988). Ostwald processes and mineral paragenesis in sediments. American Journal of Science, 288, 537560.CrossRefGoogle Scholar
Muffler, P. L. J., & White, D. E. (1969). Active metamorphism of Upper Cenozoic sediments in the Salton Sea geothermal field and the Salton Trough, southeastern California. Geological Society of America Bulletin, 80, 157182.CrossRefGoogle Scholar
Nadeau, P. H., & Reynolds, R. C. Jr. (1981). Burial and contact metamorphism in the Mancos Shale. Clays and Clay Minerals, 29, 249259.CrossRefGoogle Scholar
Nieto, F., Mellini, M., & Abad, I. (2010). The role of H 3O + in the crystal structure of illite. Clays and Clay Minerals, 58, 238246.CrossRefGoogle Scholar
Ostwald, W. Z. (1897). Studien über die Bildung und Umwandlung fester Körper. I. Abhandlung: Ubersattigung and Uberkaltung. Zeitschrift für Physikalische Chemie, 22, 289330.CrossRefGoogle Scholar
Perry, E., & Hower, J. (1970). Burial diagenesis in Gulf Coast pelitic sediments. Clays and Clay Minerals, 18, 165177.CrossRefGoogle Scholar
Pollastro, R. M. (1985). Mineralogical and morphological evidence for the formation of illite at the expense of illite/smectite. Clays and Clay Minerals, 33, 265274.CrossRefGoogle Scholar
Prigogine, I. (1961). Introduction to Thermodynamics of Irreversible Processes, 2e. New York: Interscience Publishers–John Wiley.Google Scholar
Primmer, T. J. (1994). Some comments on the chemistry and stability of interstratified illite-smectite and the role of Ostwald-type processes. Clay Minerals, 29, 6368.CrossRefGoogle Scholar
Primmer, T. J., Warren, E. A., Sharma, B. K., & Atkins, M. P. (1993). The stability of experimental grown clay minerals: implications for modelling the stability of neoformed clay minerals. In Manning, D. A. C., Hall, P. L., & Hughes, C. R. (Eds.), Geochemistry of Clay-Pore Fluid Interactions (pp. 163180). London: Chapman & Hall.Google Scholar
Pytte, A. M., & Reynolds, R. C. (1989). The thermal transformation of smectite to illite. In Naeser, N. D. & McCulloh, T. H. (Eds.), Thermal History of Sedimentary Basins (pp. 133140). New York, NY: Springer.CrossRefGoogle Scholar
Reesman, A. L., & Keller, W. D. (1968). Aqueous solubility studies of high alumina and clay minerals. American Mineralogist, 53, 929942.Google Scholar
Roberson, H. E., & Lahann, R. W. (1981). Smectite to illite conversion rates: effects of solution chemistry. Clays and Clay Minerals, 29, 129135.CrossRefGoogle Scholar
Rosenberg, P. E., Kittrick, J. A., & Sass, B. M. (1985). Implications of illite/smectite stability diagrams: A discussion. Clays and Clay Minerals, 33, 561562.CrossRefGoogle Scholar
Rosenberg, P. E., Kittrick, J. A., & Aja, S. U. (1990). Mixed-layer illite/smectite: A multiphase model. American Mineralogist, 75, 11821185.Google Scholar
Routson, R. C., & Kittrick, J. A. (1971) Illite solubility. Proceeding of Soil Science Society of America, 36, 714718.CrossRefGoogle Scholar
Sass, B. M., Rosenberg, P. E., & Kittrick, J. A. (1987). The stability of illite/smectite during diagenesis: An experimental study. Geochimica et Cosmochimica Acta, 51, 21032115.CrossRefGoogle Scholar
Sharma, V., & Annila, A. (2007). Natural process – natural selection. Biophysical Chemistry, 127, 123128.CrossRefGoogle ScholarPubMed
Smith, M. M., Dai, Z., & Caroll, S. A. (2017). Illite dissolution kinetics from 100 to 280 C and pH 3 to 9. Geochimica et Cosmochimica Acta, 209, 923.CrossRefGoogle Scholar
Środoh, J. (1984). X-ray powder diffraction identification of illitic materials. Clays and Clay Minerals, 32, 337349.CrossRefGoogle Scholar
Środoh, J., & Eberl, D. D. (1984) Illite. Reviews in Mineralogy, 13, 495544.Google Scholar
Środoh, J., Eberl, D. D., & Drits, V. A. (2000). Evolution of fundamental-particle size during illitization of smectite and implications for reaction mechanism. Clays and Clay Minerals, 48, 446458.CrossRefGoogle Scholar
Steiner, A. (1968). Clay minerals in hydrothermally altered rocks at Wairakei, New Zealand. Clays and Clay Minerals, 16, 193213.CrossRefGoogle Scholar
Tillick, D. A., Peacor, D. R., & Mauk, J. L. (2001). Genesis of dioctahedral phyllosilicates during hydrothermal alteration of volcanic rocks: I. the Golden Cross epithermal ore deposit, New Zealand. Clays and Clay Minerals, 49, 126140.CrossRefGoogle Scholar
Van Moort, J. C. (1971). A comparative study of the diagenetic alteration of clay minerals in Mesozoic shales from Papua, New Guinea, and in Tertiary shales from Louisiana, U.S.A. Clays and Clay Minerals, 19, 120.CrossRefGoogle Scholar
Vidal, O., & Dubacq, B. (2009). Thermodynamic modelling of clay dehydration, stability and compositional evolution with temperature, pressure and H2O activity. Geochimica et Cosmochimica Acta, 73, 65446564.CrossRefGoogle Scholar
Vidal, O., Dubacq, B., & Lanari, P. (2010). Comment on “The role of H 3O+ in the crystal structure of illite by F. Nieto, M. Melini, and I. Abad”. Clays and Clay Minerals, 58, 717720.CrossRefGoogle Scholar
Wagner, C. (1961). Theorie der Alterung von Neiderschliigen durch UmItisen (Ostwald Reifung). Zeitschrift fiir Electrochemie, 65, 581591.Google Scholar
White, G. N., & Zelazny, L. W. (1988). Analysis and implications of the edge structure of dioctahedral phyllosilicates. Clays and Clay Minerals, 36, 141146.CrossRefGoogle Scholar
Wolery, T. J. (1993) EQ3/6, A software package for geochemical modelling of aqueous systems (Version 7.2). Lawrence Livermore Nat. Lab. UCRL-MA 110662.Google Scholar
Yates, D. M. (1993) Experimental investigation of the formation and stability of endmember illite from 100 to 250°C and Pv.H20. PhD thesis, Washington State University.Google Scholar
Yates, D.M., & Rosenberg, P.E. (1996) Formation and stability of endmember illite: I. Solution equilibration experiments at 100° to 150° C and Pv, soln. Geochimica et Cosmochimica Acta, 60, 18731883.CrossRefGoogle Scholar
Yates, D.M., & Rosenberg, P.E. (1997) Formation and stability of endmember illite: II. Solid equilibration experiments at 100 to 250°C and Pv, soln. Geochimica et Cosmochimica Acta, 61, 31353144.CrossRefGoogle Scholar
Yates, D. M., & Rosenberg, P. E. (1998). Characterization of neoformed illite from hydrothermal experiments at 250 8C and Pv, soln: An HRTEM/ATEM study. American Mineralogist, 83, 11991208.CrossRefGoogle Scholar
Yau, S. -T., Petsev, D. N., Thomas, B. R., & Vekilov, P. G. (2000). Molecular-level thermodynamic and kinetic parameters for the self-assembly of apoferritin molecules into crystals. Journal of Molecular Biology, 303, 667678.CrossRefGoogle ScholarPubMed
Zhang, G., Kim, J., Dong, H., & Sommer, A. J. (2007). Microbial effects in promoting the smectite to illite reaction: role of organic matter intercalated in the interlayer. American Mineralogist, 92, 14011410.CrossRefGoogle Scholar