Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-26T03:14:00.764Z Has data issue: false hasContentIssue false

Alteration zonation in the Loma Blanca kaolin deposit, Los Menucos, Province of Rio Negro, Argentina

Published online by Cambridge University Press:  09 July 2018

S. A. Marfil*
Affiliation:
Department of Geology–INGEOSUR, Universidad Nacional del Sur, San Juan 670, 8000 Bahía Blanca, Argentina Independent Researcher at CIC de la Provincia de Buenos Aires, Argentina
P. J. Maiza
Affiliation:
Department of Geology–INGEOSUR, Universidad Nacional del Sur, San Juan 670, 8000 Bahía Blanca, Argentina Principal Researcher at CONICET. Argentina
N. Montecchiari
Affiliation:
Department of Geology–INGEOSUR, Universidad Nacional del Sur, San Juan 670, 8000 Bahía Blanca, Argentina
*

Abstract

The Loma Blanca mine is in one of the northwest kaolinized zones of the Los Menucos area (Patagonia, Argentina). The parent rocks are andesites and andesitic tuffs from the Vera Formation, Los Menucos Group (Lower Triassic). Hayase & Maiza (1974) proposed a concentric zonation model. From the parent rock outward, four different alteration patterns were recognized: Zone 1, with sericite, chlorite and montmorillonite; Zone 2, with kaolinite and dickite; Zone 3, with dickite, pyrophyllite and alunite; and Zone 4, with quartz, disseminated sulphides and diaspore.

The relationship between the chemical composition of major, minor and trace elements and the mineralogical alteration zonation was evaluated to confirm the genesis of the deposit. Fe2O3, CaO, Na2O and K2O contents decrease from Zone 1 to Zone 3, whereas Al2O3 and LOI increase in the kaolinite-alunite zone. In the chemical composition of alunite, Na > K. Large Ba, Sr, V and Zr contents were observed mainly in Zones 2 and 3. Co, Ni, Cu, Zn and Rb are more common in Zone 1. LREE are more abundant than HREE in Zones 2 and 3. In kaolinites, δ18O values range from 10.8% to 13.2%, and δD from –83% to –85%.

The mineral assemblage (dickite-alunite-pyrophyllite-diaspore), the alteration zonation pattern (laterally concentric), the geochemistry of trace elements, the relation between LREE and HREE and the small δ18O values suggest that the Loma Blanca deposit was formed by hydrothermal processes.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2010

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

Cravero, F., Domínguez, E. & Iglesias, C. (2001) Genesis and applications of the Cerro Rubio kaolin deposit, Patagonia (Argentina). Applied Clay Science, 18, 157172.Google Scholar
Dill, H. (2001) The geology of aluminium phosphates and sulphates of the alunite group minerals: a review. Earth Science Reviews, 53, 3593.Google Scholar
Dill, H.G., Bosse, H.-R., Henning, K.-H. & Fricke, A. (1997) Mineralogical and chemical variations in hypogene and supergene kaolin deposits in a mobile fold belt — The Central Andes of northwestern Peru. Mineralium Deposita, 32, 149163.Google Scholar
Dill, H., Bosse, H.-R. & Kassbohm, J. (2000) Mineralogical and chemical studies of volcanicrelated argillaceous industrial minerals of the Central American Cordillera (Western El Salvador). Economic Geology, 95, 517538.Google Scholar
Ece, Ö., Schroeder, P.A., Smilley, M.J. & Wampler, J.M. (2008) Acid-sulphate hydrothermal alteration of andesitic tuffs and genesis of halloysite and alunite deposits in the Biga Peninsula, Turkey. Clay Minerals, 43, 281315.Google Scholar
Grim, H. (1969) Clay Mineralogy. 2 nd Edition. Pp. 131137. McGraw Hill Book Co, New York, USA.Google Scholar
Hayase, K. & Maiza, P. (1974) Génesis del yacimiento de caolín de mina Loma Blanca, Los Menucos, Provincia de Río Negro, Argentina. V Congreso Geológico Argentino. Buenos Aires, Argentina, 139-151.Google Scholar
Hemley, J.J. (1959) Some mineralogical equilibria in the system K2O-Al2O3-SlO2-H2O. American Journal of Science, 257, 241270.Google Scholar
Labudia, C. & Bjerg, E. (1994) Geología del sector oriental de la hoja Bajo Hondo (39e), Provincia de Río Negro. Revista de la Asociacion Geológica Argentina, 49, 284296.Google Scholar
Maiza, P. (1972) Los yacimientos de caolín originados por la actividad hidrotermal en lo principales distritos caoliniferos de la Patagonia, República Argentina. PhD thesis, Universidad Nacional del Sur, Argentina.Google Scholar
Maiza, P. & Mas, G. (1980) Estudio de los sulfatos alunita—natroalunita. Sintesis de la serie. Revista de la Asociación Argentina de Mineralogia, Sedimentologia y Petrologia, 11, 3241.Google Scholar
Maiza, P. & Mas, G. (1981) Presencia de natroalunita en Mina Equivocada, Río Negro. Su significado. 8th Congreso Geologico Argentino, San Luis, Actas IV, 285-292.Google Scholar
Maiza, P.J., Pieroni, D. & Marfil, S.A. (2003) Geochemistry of the hydrothermal kaolins in the SE area of Los Menucos. Prov. de Río Negro. A clay odyssey. Proceedings of the 12th International Clay Conference. Argentina, 2001, Elsevier, pp. 123130.Google Scholar
Maiza, P.J., Marfil, S.A. & Montecchiari, N. (2009) Alteration zonation ‘Loma Blanca’ kaolin deposit, Los Menucos, Río Negro Province, Argentina. XIV International Clay Conference, Abstracts Vol. 1, p. 188.Google Scholar
Marfil, S.A., Pieroni, D. & Maiza, P.I. (2000) Dickita y alunita en mina Blanquita. Los Menucos. (Province de Río Negro). V Congreso de Mineralogía y Metalogenia. MINMET 2000, La Plata. Argentina, 281286.Google Scholar
Marfil, S.A., Maiza, P.I., Cardellach, E., & Corbella, M. (2005) Origin of kaolin deposits in the ‘Los Menucos', Río Negro Province, Argentina. Clay Minerals, 40, 283293.Google Scholar
Murray, H. & lanssen, I. (1984) Oxygen isotopes—indicators of kaolin genesis. Proceedings of the 27th International Geological Congress, 15, 287303.Google Scholar
Pandarinath, K., Dulski, P., Torres Alvarado, I.S. & Verma, S.P. (2008) Element mobility during the hydrothermal alteration of rhyolitic rocks of the Los Azufres geothermal field, Mexico. Geothermics, 37, 5372.Google Scholar
Pankurst, R.J.C., Rapela, W., Caminos, R., Llambías, E. & Pariea, C. (1992) A revised age for the granites of the central Somoncura Batholith, North Patagonian Massif. Journal of South American Earth Sciences, 5, 321325.Google Scholar
Papoulis, D. & Tsolis-Katagas, P. (2008) Formation of alteration zones and kaolin genesis, Limnos Island, northeast Aegean Sea, Greece. Clay Minerals, 43, 631646.Google Scholar
Papoulis, D., Tsolis-Katagas, P. & Katagas, C. (2004). Monazite alteration mechanisms and depletion measurements in kaolins. Applied Clay Science, 24, 271285.Google Scholar
Reed, B.L. & Hemley, J.J. (1967) Ocurrence of pyrophyllite in Kekiktuk Conglomerate. Pp. 162166 in: Book Range, Northeastern Alaska. U.S. Geological Survey. Google Scholar
Rollinson, H. (1992) Using Geochemical Data: Evaluation, Presentation, Interpretation. University of Zimbabwe.Google Scholar
Roy, R. & Osborn, E. (1954) The system Al2O3-SiO2-H2O. American Mineralogist, 39, 853885.Google Scholar
Terakado, Y. & Fujitani, T. (1998) Behavior of the rare earth elements and other trace elements during interactions between acidic hydrothermal solutions and silicic volcanic rocks, southwestern Japan. Geochimica et Cosmochimica Ada, 62, 19031998.Google Scholar
Tzuzuki, Y. & Mizutani, S. (1971) A study of rock alternation process based on kinetics of hydrothermal experiment. Contributions to Mineralogy and Petrology, 30, 1533.Google Scholar