Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-26T11:43:23.968Z Has data issue: false hasContentIssue false

Fe concentration in palaeosols and in clayey marine sediments: two case studies in the Variscan basement of Sardinia (Italy)

Published online by Cambridge University Press:  09 July 2018

P. Mameli*
Affiliation:
Istituto di Scienze Geologico-Mineralogiche, Università degli Studi di Sassari, Corso Angioj 10, I-07100 Sassari, Italy
G. Mongelli
Affiliation:
Dipartimento di Scienze Geologiche, Università degli Studi della Basilicata Campus di Macchia Romana, I-85100 Potenza, Italy
G. Oggiano
Affiliation:
Istituto di Scienze Geologico-Mineralogiche, Università degli Studi di Sassari, Corso Angioj 10, I-07100 Sassari, Italy
R. Sinisi
Affiliation:
Dipartimento di Scienze Geologiche, Università degli Studi della Basilicata Campus di Macchia Romana, I-85100 Potenza, Italy
*

Abstract

Within the Variscan basement of Sardinia (Italy), two main Fe concentrations occur in the low-grade metamorphic tectonic units: (1) an uppermost-Ordovician oolitic ironstone of shallow anoxic water environment; and (2) a concentration of Fe oxyhydroxides lying on a palaeosurface. Two sets of samples were picked from the marine ironstone and from the continental Fe concentration following stratigraphic criteria. Chemical analysis, X-ray diffraction, scanning electron microscopy-energy dispersive X-ray analysis and thin-section studies were performed on 34 samples.

Marine ironstones formed under a highly reducing, anoxic, non-sulphidic methanic environment, and their Fe phases are chamosite, siderite and magnetite. Detrital chlorite and illite, produced during physical weathering, were chamosite precursors. Using the V/Cr proxy, an emergence stage that caused a transition to an oxic environment is documented. In contrast, continental ironstones formed under oxic conditions and the dominant Fe phase is goethite, which can adsorb Zn2+ and U6+. Unexpected negative Ce anomalies occur in this set of samples, suggesting that the oxyhydroxides originated from Ce-depleted solutions. Although the ironstones of Sardinia formed in different environments (marine vs. continental) and under contrasting climatic conditions (sub-glacial vs. tropical) they share similar geochemical features. These dramatic palaeoenvironmental differences did not result in large differences between the geochemistry of the chemical sediments.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2008

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

Algeo, T.J. & Maynard, J.B. (2004) Trace-element behavior and redox facies in core shales of Upper Pennsylvanian Kansas-type cyclothems. Chemical Geology, 206, 289318.Google Scholar
Amadesi, E., Cantelli, C., Carloni, G.C. & Rabbi, E. (1961) Ricerche geologiche sui terreni sedimentari del foglio 208 Dorgali. Giornale di Geologia, 28, 5987.Google Scholar
Bau, M. (1994) Modeling of rare-earth element partitioning between particles and solution in aquatic environments — comment. Geochimica et Cosmochimica Acta, 58, 45214523.CrossRefGoogle Scholar
Bau, M. (1999) Scavenging of dissolved yttrium and rare earths by precipitating iron oxyhydroxide: experimental evidence for Ce oxidation, Y—Ho fractionation and lanthanide tetrad effect. Geochimica et Cosmochimica Acta, 63, 6777.Google Scholar
Bau, M., Usui, A., Pracejus, B., Mita, N., Kanai, Y., Irber, W. & Dulski, P. (1998) Geochemistry of low-temperature water—rock interaction: evidence from natural waters, andesite, and iron oxyhydroxide precipitation at Nishiki-numa iron spring, Hokkaido, Japan. Chemical Geology, 151, 293307.Google Scholar
Becq-Giraudon, J.F., Bouille, S. & Chauvel, J.J. (1992) Genesis and significance of the silico-aluminous nodules in the Ordovician of the Montagne Noire and the Massif Armoricain (France). Sedimentary Geology, 77, 7787.Google Scholar
Berner, R.A. (1980) A new geochemical classification of sedimentary environments. Journal of Sedimentary Petrology, 51, 359365.Google Scholar
Braun, J.J., Pagel, M., Muller, J.P., Bilong, P., Micherd, A. & Guillet, B. (1990) Cerium anomalies in lateritic profiles. Geochimica Cosmochimica Acta, 54, 781795.Google Scholar
Brookins, D.G. (1988) Eh-pH Diagrams for Geochemistry. Springer-Verlag, Berlin.Google Scholar
Bruno, J., De Pablo, J., Duro, L., Figuerola, E. (1995) Experimental study and modeling of the U(VI)-Fe(OH)3 surface precipitation/coprecipitation equilibria. Geochimica et Cosmochimica Acta, 59, 41134123.Google Scholar
Buzzi, L., Gaggero, L., Funedda, A., Oggiano, G. & Tiepolo, M. (2007) Zircon geochronology and Sr-Nd isotopic study of the Ordovician magmatic events in the southern Variscides (Sardinia). Geologie de la France, 2, 73.Google Scholar
Capelli, B., Carmignani, L., Castorina, F., Di Pisa, A., Oggiano, G. & Petrini, R. (1992) A Hercynian suture zone in Sardinia: geological and geochemical evidence. Geodinamica Acta, 5, 101118.Google Scholar
Carmignani, L., Carosi, R., Di Pisa, A., Gattiglio, M., Musumeci, G., Oggiano, G. & Pertusati, P.C. (1994) The Hercynian chain of Sardinia (Italy). Geodinamica Acta, 7, 3147.Google Scholar
Dieni, I. & Massari, F. (1966) I Foraminiferi del Valanginiano superiore di Orosei (Sardegna). Paleontologia Italiana, 61, 75186.Google Scholar
Drever, J.I. (1997) The geochemistry of natural waters: surface and groundwater environments, 3rd edition, Prentice Hall, Upper Saddle River, NJ.Google Scholar
Duff, M.C., Coughlin, J.U. & Hunter, D.B. (2002) Uranium co-precipitation with iron oxide minerals. Geochimica et Cosmochimica Acta, 66, 35333547.Google Scholar
Ehrmann, W., Setti, M. & Marinoni, L. (2005) Clay minerals in Cenozoic sediments off Cape Roberts (McMurdo Sound, Antarctica) reveal palaeoclimatic history. Palaeogeography, Palaeoclimatology, Palaeoecology, 229, 187211.Google Scholar
Franceschelli, M., Puxeddu, M. & Carta, M. (2000) Mineralogy and geochemistry of late Ordovician Phosphate-bearing oolitic ironstones from NW Sardinia, Italy. Mineralogy and Petrology, 69, 267293.Google Scholar
Franzini, M., Leoni, L. & Saitta, M. (1972) A simple method to evaluate the matrix effects in X-ray fluorescence analysis. X-ray Spectrometry, 1, 151154.Google Scholar
Franzini, M., Leoni, L. & Saitta, M. (1975) Revisione di una metodologia analitica per fluorescenza X basata sulla correzione degli effetti di matrice. Societa Italiana di Mineralogia e Petrologia, 31, 365378.Google Scholar
Hren, M.T., Lowe, D.R., Tice, M.M., Byerly, G. & Chamberlain, C.P. (2006) Stable isotope and rare earth element evidence for recent ironstone pods within the Archean Barberton greenstone belt, South Africa. Geochimica et Cosmochimica Acta, 70, 14571470.Google Scholar
Jones, B. & Manning, D.A.C. (1994) Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chemical Geology, 111, 111129.CrossRefGoogle Scholar
Kim, Y. & Lee, Y.I. (2000) Ironstones and green marine clays in the Dongjeom Formation (Early Ordovician) of Korea. Sedimentary Geology, 130, 6580.Google Scholar
Koeppenkastrop, D. & De Carlo, E.H. (1992) Sorption of rare-earth elements from seawater onto synthetic mineral particles — an experimental approach. Chemical Geology, 95, 251263.Google Scholar
Leoni, L. & Saitta, M. (1976) Determination of yttrium and niobium on standard silicate rocks by X-ray fluorescence analysis. X-ray Spectrometry, 5, 2930.CrossRefGoogle Scholar
Marshall, C.P. & Fairbridge, R.W. (1999) Encyclopedia of Geochemistry, Kluwer Academy Publishers, Dordrecht.Google Scholar
Maynard, J.B. (1982) Extension of Berner's ‘New geochemical classification of sedimentary environments’ to ancient sediments. Journal of Sedimentary Petrology, 52, 13251331.Google Scholar
Mongelli, G. (1997) Ce-anomalies in the textural components of Upper Crétacéous karst bauxites from the Apulian Carbonate platform (southern Italy). Chemical Geology, 140, 6979.Google Scholar
Mücke, A. (2000) Environmental conditions in the late Crétacéous African Tethys: conclusion from a microscopic—microchemical study of ooidal iron-stones from Egypt, Sudan and Nigeria. Journal of African Earth Sciences, 30, 2546.CrossRefGoogle Scholar
Mücke, A. (2006) Chamosite, siderite and the environmental conditions of their formation in chamosite-type Phanerozoic ooidal ironstones. Ore Geology Reviews, 28, 235249.Google Scholar
Mücke, A. & Farshad, F. (2005) Whole-rock and mineralogical composition of Phanerozoic ooidal ironstones: comparison and differentiation of types and subtypes. Ore Geology Review, 26, 227262.Google Scholar
Oggiano, G. (1994) Lineamenti stratigrafico-strutturali del basamento del Goceano (Sardegna centro-settentrionale). Bollettino della Societa Geologica Italiana, 113, 471480.Google Scholar
Oggiano, G. & Mameli, P. (2006) Diamictite and oolitic ironstones, a sedimentary association at Ordovician—Silurian transition in the North Gondwana Margin: new evidence from the Inner Nappe of Sardinia Variscides (Italy). Gondwana Research, 9, 500511.Google Scholar
Plank, T. & Langmuir, C.H. (1998) The chemical composition of subducting sediment and its composition for the crust and mantle. Chemical Geology, 145, 325394.CrossRefGoogle Scholar
Schwertmann, U. & Murad, E. (1983) Effect of pH on the formation of goethite and hematite from ferrihydrite. Clays and Clay Minerals, 31, 277284.Google Scholar
Stumm, W. (1992) Chemistry of the Solid-Water Interface. Wiley & Sons, New Jersey.Google Scholar
Tait, J., Schatz, M., Bachtadse, V. & Soffel, H. (2000) Palaeomagnetism and Palaeozoic palaeogeography of Gondwana and European terranes. Pp. 2134 in. Orogenic Processes: Quantification and Modelling in the Variscan Belt (Franke, W., Haak, V., Oncken, O. & Tanner, D., editors). Special Publication, 179, Geological Society of London.Google Scholar
Thiessen, H., Lo Monaco, S., Ramirez, A., Santos, M.C.D. & Shang, C. (1996) Phosphate minerals in a lateritic crust from Venezuela. Biogeochemistry, 34, 117.Google Scholar
von Raumer, J.F., Stampfli, G.M. & Bussy, F. (2003) Gondwana-derived microcontinents — the constituents of the Variscan and Alpine collisional orogens. Tectonophysics, 365, 722.CrossRefGoogle Scholar
Yapp, C.J. (1993) Paleoenvironment and the oxygen isotope geochemistry of ironstone of the Upper Ordovician Neda Formation, Wisconsin, USA. Geochimica et Cosmochimica Acta, 57, 23192327.Google Scholar