Hostname: page-component-84b7d79bbc-x5cpj Total loading time: 0 Render date: 2024-07-26T22:06:31.238Z Has data issue: false hasContentIssue false

The Fe-Mg-saponite solid solution series – a hydrothermal synthesis study

Published online by Cambridge University Press:  27 February 2018

A. Baldermann*
Affiliation:
Institute of Applied Geosciences, Graz University of Technology, Graz, Austria
R. Dohrmann
Affiliation:
BGR, Bundesanstalt für Geowissenschaften und Rohstoffe, Hannover, Germany
S. Kaufhold
Affiliation:
BGR, Bundesanstalt für Geowissenschaften und Rohstoffe, Hannover, Germany
C. Nickel
Affiliation:
Institute of Applied Geosciences, Graz University of Technology, Graz, Austria
I. Letofsky-Papst
Affiliation:
Institute of Electron Microscopy and Nanoanalysis, Graz University of Technology, Graz, Austria
M. Dietzel
Affiliation:
Institute of Applied Geosciences, Graz University of Technology, Graz, Austria

Abstract

The boundary conditions of saponite formation are generally considered to be well known, but significant gaps in our knowledge persist in respect to the influence of solution chemistry, temperature, and reaction time on the mineralogy, structure, stability, and chemical composition of laboratory-grown ferrous saponite. In the present study, ferrous saponite and Mgsaponite were synthesized in Teflon-lined, stainless steel autoclaves at 60, 120 and 180°C, alkaline pH, reducing conditions, and initial solutions with molar Si:Fe:Mg ratios of 4:0:2, 4:1:1, 4:1.5:0.5, 4:1.75:0.25, and 4:1.82:0.18. The experimental solutions were prepared by dissolution of sodium orthosilicate (Na4SiO4), iron(II)sulfate (FeSO4·6H2O) and magnesium chloride salts (MgCl2·6H2O with ≤ 0.005 mass% of K and Ca) in 50 mL ultrapure water that contained 0.05% sodium dithionite as the reducing agent. The precipitates obtained at two, five and seven days of reaction time were investigated by X-ray diffraction techniques, transmission electron microscopy analysis, infra-red spectroscopy, and thermo-analytical methods.

The precipitates were composed mainly of trioctahedral ferrous saponite, with small admixtures of co-precipitated brucite, opal-CT, and 2-line ferrihydrite, and nontronite as the probable alteration product of ferrous saponite. The compositions of the obtained ferrous saponites were highly variable, (Na0.44−0.59 K0.00−0.05 Ca0.00−0.02) (Fe2+0.37−2.41 Mg0.24−2.44 Fe3+0.00−0.282.65−2.85 [(Fe3+0.00−0.37 Si3.63−4.00)O10](OH)2, but show similarities with naturally occurring trioctahedral Fe and Mg end members, except for the Al content. This suggests that a complete solid solution may exist in the Fe-Mg-saponite series.

A conceptual reaction sequence for the formation of ferrous saponite is developed based on the experimental solution and solid compositions. Initially, at pH ≥ 10.4, brucite-type octahedral template sheets are formed, where dissolved Si-O tetrahedra are condensed. Subsequent reorganization of the octahedra and tetrahedra via multiple dissolution-precipitation processes finally results in the formation of saponite structures, together with brucite and partly amorphous silica. The extent of Fe2+ incorporation in the octahedral template sheets via isomorphic substitution is suggested to stabilize the saponite structure, explaining (i) the abundance of saponite enriched in VIFe2+ at elevated Fe supply and (ii) the effect of structural Fe on controlling the net formation rates of ferrous saponite.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2014

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

Akbulut, A. & Kadir, S. (2003) The geology and origin of sepiolite, palygorskite and saponite in Neogene lacustrine sediments of the Serinhisar-Acipayam Basin, Denizli, SW Turkey. Clays and Clay Minerals, 51, 279292.Google Scholar
Badaut, D., Besson, G., Decarreau, A. & Rautureau, R. (1985) Occurrence of a ferrous, trioctahedral smectite in recent sediments of Atlantis II Deep, Red Sea. Clay Minerals, 20, 389404.CrossRefGoogle Scholar
Bailey, S.W., Alietti, A., Brindley, G.W., Formosa, M.L.L., Jasmund, K., Konta, J., MacKenzie, R.C., Nagasawa, K., Raussell-Colom, R.A. & Zvyagin, B.B. (1980) Summary of the recommendations of the AIPEA nomenclature committee. Clays and Clay Minerals, 28, 7378.Google Scholar
Baldermann, A., Warr, L.N., Grathoff, G.H. & Dietzel, M. (2013) The rate and mechanism of deep-sea glauconite formation at the Ivory Coast – Ghana Marginal Ridge. Clays and Clay Minerals, 61, 258276.Google Scholar
Breukelaar, J., Kellendonk, F.J.A. & van Santen, R.A. (1989) A process for the manufacture of synthetic saponites. European Patent 317,006.Google Scholar
Breukelaar, J., van Santen, R.A. & De Winter, A.W. (1990) Synthetic saponite-derivatives, a method for preparing such saponites and their use in catalytic (hydro)conversions. European Patent 398, 429.Google Scholar
Brigatti, (1983) Relationships between composition and structure in Fe-rich smectites. Clay Minerals, 18, 177186.CrossRefGoogle Scholar
Brigatti, M.F., Lugli, C., Poppi, L. & Venturelli, G. (1999) Iron-rich saponite: dissolution reactions and Cr uptake. Clay Minerals, 34, 637645.CrossRefGoogle Scholar
Caillère, S., Henin, S. & Esquevin, J. (1953) Synthèses à basse tempèrature de phyllite ferrifère. Comptes Rendus de l’Academie Sciences (Paris), 237, 17241726.Google Scholar
Caillère, S., Oberlin, A. & Henin, S. (1954) Etude au microscope electronique de quelques silicates phylliteux obtenus par synthèses à basse tempèrature. Clay Minerals Bulletin, 2, 146156.CrossRefGoogle Scholar
Caillère, S., Henin, S. & Esquevin, J. (1955) Synthèses à basse tempèrature de quelque minèraux ferrifère (silicates et oxydes). Bulletin de la Société Française de Minéralogie et Cristallographie, 78, 227241.Google Scholar
Che, C., Glotch, T.D., Bish, D.L., Michalski, J.R. & Xu, W. (2011) Spectroscopic study of the dehydration and/or dehydroxylation of phyllosilicate and zeolite minerals. Journal of Geophysical Research, 116, E05007.Google Scholar
Chukanov, N.V., Pekov, I.V., Zadov, A.E., Chukanova, V.N. & Mökkel, S. (2003) Ferrosaponite Ca0.3(Fe2+, Mg, Fe3+)3(Si,Al)4O10(OH)2·4H2O, the new trioctahedral smectite. Zapiski Vserossiyskogo Mineralogicheskogo Obshchestva, 132, 6874.Google Scholar
Corliss, J.B., Dymond, J., Gordon, L.I., Edmond, J.M. Von Herzen, R.P., Ballard, R.D., Green, K., Williams, D., Bainbridge, A., Crane, K. & Van Adel, T.H. (1979) Submarine thermal springs on Galapogos Rift. Science, 203, 10731083.Google Scholar
Cornell, R.M. & Schwertmann, U. (2003) The Iron Oxides: Structure, Properties, Reactions, Occurrences and Uses. 664 pp. Wiley-VCH, Weinheim.Google Scholar
Cuadros, J., Dekov, V.M. & Fiore, S. (2008) Crystalchemistry of the mixed-layer sequence talc–talcsmectite –smectite from submarine hydrothermal vents. American Mineralogist, 93, 13381348.Google Scholar
Cuadros, J., Michalski, J.R., Dekov, V.M., Bishop, J., Fiore, S. & Dyar, M.D. (2013) Crystal-chemistry of interstratified Mg/Fe-clay minerals from seafloor hydrothermal sites. Chemical Geology, 360–361, 142158.Google Scholar
Decarreau, A. (1980) Cristallogènese expérimentale des smectites magnésiennes: Hectorite, stévensite. Bulletin de Minéralogie, 103, 579590.CrossRefGoogle Scholar
Decarreau, A. (1985) Partitioning of divalent elements between octahedral sheets of trioctahedral smectites and water. Geochimica et Cosmochimica Acta, 49, 15371544.Google Scholar
Decarreau, A. & Bonnin, D. (1986) Synthesis and crystallogenesis at low temperature of Fe(III)- smectites by evolution of coprecipitated gels: Experiments in partially reducing conditions. Clay Minerals, 21, 861877.Google Scholar
Decarreau, A., Petit, S., Martin, F., Farges, F., Vieillard, P. & Joussein, E. (2008) Hydrothermal synthesis, between 75 and 150°C, of high-charge, ferric nontronites. Clays and Clay Minerals, 56, 322337.CrossRefGoogle Scholar
Deocampo, D.M., Cuadros, J., Wing-Dudek, T., Olives, J. & Amouric, M. (2009) Saline lake diagenesis as revealed by coupled mineralogy and geochemistry of multiple ultrafine clay phases: Pliocene Olduvai Gorge, Tanzania. American Journal of Science, 309, 834868.Google Scholar
Dietzel, M. & Letofsky-Papst, I. (2002) Stability of magadiite between 20 and 100°C. Clays and Clay Minerals, 50, 656665.Google Scholar
Dill, H.G., Dohrmann, R. & Kaufhold, S. (2011) Disseminated and faultbound autohydrothermal ferroan saponite in Late Paleozoic andesites of the Saar-Nahe Basin, SW Germany: Implications for the economic geology of intermediate (sub)volcanic rocks. Applied Clay Science, 51, 226240.Google Scholar
Drits, V.A., Plançon, A. Sakharov, B.A., Besson, G., Tsipursky, S.I. & Tchoubar, C. (1984) Diffraction effects calculated for structural models of Ksaturated montmorillonite containing different types of defects. Clay Minerals, 19, 541562.Google Scholar
Drits, V.A., Besson, G. & Muller, F. (1995) An improved model for structural transformations of heat-treated aluminous dioctahedral 2:1 layer silicates. Clays and Clay Minerals, 43, 718731.Google Scholar
Dymond, J., Corliss, J.B., Heath, G.R., Field, C.W., Dasch, E.J. & Veeh, H.H. (1973) Origin of metalliferous sediments from the Pacific Ocean. Geological Society of America Bulletin, 84, 33553372.Google Scholar
Eggleton, R.A. & Fitzpatrick, R.W. (1988) New data and a revised structural model for ferrihydrite. Clays and Clay Minerals, 36, 111124.Google Scholar
Farmer, V.C., McHardy, W.J., Elsass, F. & Robert, M. (1994) hk-ordering in aluminous nontronite and saponite synthesized near 90°C: Effects of synthesis conditions on nontronite composition and ordering. Clays and Clay Minerals, 42, 180186.Google Scholar
Garvie, L.A.J., Craven, A.J. & Brydson, R. (1994) Use of electron-energy loss near-edge fine structure in the study of minerals. American Mineralogist, 79, 411425.Google Scholar
Grauby, O., Petit, S., Decarreau, A. & Baronnet, A. (1994) The nontronite-saponite series: An experimental approach. European Journal of Mineralogy, 6, 99112.Google Scholar
Greenberg, S.A. (1958) The nature of the silicate species in sodium silicate solutions. Journal of The American Chemical Society, 80, 65086511.CrossRefGoogle Scholar
Greenberg, S.A. & Price, E.W. (1957) The solubility of silica in solutions of electrolytes. The Journal of Physical Chemistry, 61, 15391541.Google Scholar
Harder, H. (1972) The role of magnesium in the formation of smectite minerals. Chemical Geology, 10, 3139.Google Scholar
Harder, H. (1976) Nontronite synthesis at low temperatures. Chemical Geology, 18, 169180.Google Scholar
Harder, H. (1978) Synthesis of iron layer silicate minerals under natural conditions. Clays and Clay Minerals, 26, 6572.CrossRefGoogle Scholar
Henin, S. & Robichet, O. (1954) A study of the synthesis of clay minerals. Clay Minerals, 2, 110115.Google Scholar
Heuser, M., Andrieux, P., Petit, S. & Stanjek, H. (2013) Iron-bearing smectites: a revised relationship between structural Fe, b cell edge lengths and refractive indices. Clay Minerals, 48, 97103.CrossRefGoogle Scholar
Iler, R.K. (1979) The Chemistry of Silica – Solubility, Polymerization, Colloid and Surface Properties, and Biochemistry. 896 pp. Wiley-Interscience, New York.Google Scholar
Iriarte, P.I., Petit, S., Huertas, F.J., Fiore, S., Grauby, O., Decarreau, A. & Linares, J. (2005) Synthesis of kaolinite with a high level of Fe3+ for Al substitution. Clays and Clay Minerals, 53, 110.Google Scholar
Jasmund, K. & Lagaly, G. (1993) Tonminerale und Tone. Struktur, Eigenschaften, Anwendung und Einsatz in Industrie und Umwelt. 490 pp. Steinkopff Verlag, Darmstadt.Google Scholar
Kawano, M. & Tomita, K. (1991) Dehydration and rehydration of saponite and vermiculite. Clays and Clay Minerals, 39, 174183.Google Scholar
Kloprogge, J.T., Komarneni, S. & Amonette, J.E. (1999) Synthesis of smectite clay minerals: A critical review. Clays and Clay Minerals, 47, 529554.Google Scholar
Köster, H.M. (1993) Beschreibung einzelner Tonminerale. Pp. 33–89 in: Tonminerale und Tone (K. Jasmund & G. Lagaly, editors). Steinkopff Verlag, Darmstadt.Google Scholar
Kuchta, L. & Fajnor, V.S. (1988) Optimal conditions for hydrothermal synthesis of saponite. Chemicke Zvesti, 42, 339345.Google Scholar
Lanson, B., Lantenois, S., van Aken, P.A., Bauer, A. & Plançon, A. (2012) Experimental investigation of smectite interaction with metal iron at 80°C: Structural characterization of newly formed Fe-rich phyllosilicates. American Mineralogist, 97, 864871.CrossRefGoogle Scholar
Madejová, J., Bujdak, J., Gates, W.P. & Komadel, P. (1996) Preparation and infrared spectroscopic characterization of reduced-charge montmorillonite with variable Li contents. Clay Minerals, 31, 233241.CrossRefGoogle Scholar
Parthasarathy, G., Choudary, B.M., Sreedhar, B., Kunwar, A.C. & Srinivasan, R. (2003) Ferrous saponite from the Deccan Trap, India, and its application in adsorption and reduction of hexavalent chromium. American Mineralogist, 88, 19831988.Google Scholar
Porter, S., Vanko, D.A. & Ghazi, A.M. (2000) Major and trace element compositions of secondary clays in basalts altered at low temperature, eastern flank of the Juan de Fuca Ridge. Proceedings of the Ocean Drilling Program, Scientific Results, 168, 149157.Google Scholar
Post, J.L. (1984) Saponite from near Ballarat, California. Clays and Clay Minerals, 32, 147153.CrossRefGoogle Scholar
Ramachandran, V.S., Paroli, R.M., Beaudoin, J.J. & Delgado, A.H. (2002) Handbook of Thermal Analysis of Construction Materials. 680 pp. Noyes Publications, New York.Google Scholar
Russell, J.D. (1979) Infrared spectroscopy of ferrihydrite: Evidence for the presence of structural hydroxyl groups. Clay Minerals, 14, 109114.Google Scholar
Russell, J.D. & Fraser, A.R. (1994) Infrared methods. Pp. 11–67 in: Clay Mineralogy: Spectroscopic and Chemical Determinative Methods (M.J. Wilson, editor). Chapman & Hall, London.Google Scholar
Sandler, A., Nathan, Y., Eshet, Y. & Raab, M. (2001) Diagenesis of trioctahedral clays in a Miocene to Pleistocene sedimentary-magmatic sequence in the Dead Sea Rift, Israel. Clay Minerals, 36, 2947.Google Scholar
Schiffman, P. & Staudigel, H. (1995) The smectite to chlorite transition in a fossil seamount hydrothermal system: the Basement Complex of La Palma, Canary Islands. Journal of Metamorphic Geology, 13, 487498.CrossRefGoogle Scholar
Tischendorf, G., Förster, H.-J., Gottesmann, B. & Rieder, M. (2007) True and brittle micas: composition and solid-solution series. Mineralogical Magazine, 71, 285320.CrossRefGoogle Scholar
Tsipursky, S.I. & Drits, V.A. (1984) The distribution of octahedral cations in the 2:1 layers of dioctahedral smectites studied by oblique texture electron diffraction. Clay Minerals, 19, 177192.Google Scholar
Van Aken, P.A., Liebscher, B. & Styrsa, V.J. (1998) Quantitative determination of iron oxidation states in minerals using Fe L2,3-edge electron energy-loss near-edge structure spectroscopy. Physics and Chemistry of Minerals, 25, 323327.Google Scholar
Velde, B. (1992) Introduction to Clay Minerals: Chemistry, Origin, Uses and Environmental Significance. 198 pp. Chapman and Hall, London.Google Scholar
Wildman, W.E., Whittig, L.D. & Jackson, M.L. (1971) Serpentine stability in relation to formation of ironrich montmorillonite in some California soils. American Mineralogist, 56, 587602.Google Scholar
Wolters, F. & Emmerich, K. (2007) Thermal reactions of smectites – Relation of dehydroxilation temperature to octahedral structure. Thermochimica Acta, 462, 8088.Google Scholar