Hostname: page-component-848d4c4894-r5zm4 Total loading time: 0 Render date: 2024-06-28T04:30:54.720Z Has data issue: false hasContentIssue false

Upper Jurassic travertine at El Macanudo, Argentine Patagonia: a fossil geothermal field modified by hydrothermal silicification and acid overprinting

Published online by Cambridge University Press:  23 June 2017

DIEGO M. GUIDO*
Affiliation:
CONICET and Facultad de Ciencias Naturales y Museo, Universidad Nacional de La Plata, Instituto de Recursos Minerales (INREMI), Calle 64 y 120, La Plata (1900), Argentina
KATHLEEN A. CAMPBELL
Affiliation:
School of Environment, University of Auckland, Private Bag 92019, Auckland 1142, New Zealand
*
Author for correspondence: diegoguido@yahoo.com

Abstract

The Deseado Massif hosts numerous Late Jurassic (150 Ma) fossil geothermal systems related to an extensive volcanic event developed in a diffuse extensional back-arc setting. Detailed mapping, petrography and mineralogical observations of El Macanudo outcrops verify that it represents a hot-spring-related travertine partially replaced by silica and delineated by six sedimentary facies. These are large concentric cones (F1), laminated vertical columnar structures (F2), porous layers (F3), shrubby and irregular lamination (F4), low-amplitude wavy bedding (F5) and mounds and breccias (F6). The Macanudo Norte Outcrop rocks constitute a silica-replaced travertine sequence, with development of large conical stromatolites in a deep pool or geothermally influenced shallow lacustrine environment, surrounded by a subaerial travertine apron terrace; whereas, the Macanudo Sur Outcrop is a subaerial travertine mound sequence. Structurally controlled vent areas occur in both northern (F1) and southern (F6) outcrops, mainly located along regional NNE- and ENE-trending faults. The other sedimentary units display a concentric distribution of travertine facies with respect to the interpreted vent areas. The El Macanudo palaeo-hot spring deposit is situated in an eroded Jurassic volcanic centre, and records a complex evolutionary-fluid history. The sediments archived three different Jurassic events, when large and long-lasting hydrothermal systems were active across the region. This relative temporal sequence was formed by: (1) travertine precipitation; (2) development of a silica cap, where early silicification was responsible for exceptional preservation of some stromatolitic fabrics; and (3) acid alteration, recorded by dissolution textures and clay formation, and caused by a palaeo-phreatic water-level drop.

Type
Original Article
Copyright
Copyright © Cambridge University Press 2017 

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

Cady, S. L. & Farmer, J. D. 1996. Fossilization processes in siliceous thermal springs: trends in preservation along thermal gradients. In Evolution of Hydrothermal Ecosystems on Earth (and Mars?) (eds Bock, G. R. & Goode, G. A.), pp. 150–73. Proceedings of the Ciba Foundation Symposium 202. Chichester: J. Wiley.Google Scholar
Campbell, K. A., Guido, D. M., Gautret, P., Foucher, F., Ramboz, C. & Westall, F. 2015a. Geyserite in hot-spring siliceous sinter: window on Earth's hottest terrestrial paleoenvironment and its extreme life. Earth-Science Reviews 148, 4464.Google Scholar
Campbell, K. A., Lynne, B. Y., Handley, K., Jordan, S., Farmer, J., Guido, D., Turner, S. & Perry, R. 2015b. Tracing biosignature preservation of geothermally silicified microbial textures into the geological record. Astrobiology 15, 858–82.Google Scholar
Campbell, K. A., Sannazzaro, K., Rodgers, K. A., Herdianita, N. R. & Browne, P. R. L. 2001. Sedimentary facies and mineralogy of the late Pleistocene Umukuri silica sinter, Taupo Volcanic Zone, New Zealand. Journal of Sedimentary Research 71, 727–46.Google Scholar
Chafetz, H. S. & Folk, R. L. 1984. Travertines: depositional morphology and the bacterially constructed constituents. Journal of Sedimentary Petrography 54, 289316.Google Scholar
Chafetz, H. S. & Guidry, S. A. 1999. Bacterial shrubs, crystal shrubs, and ray-crystal shrubs: bacterial vs. abiotic precipitation. Sedimentary Geology 126, 5774.Google Scholar
Channing, A., Zamuner, A., Edwards, D. & Guido, D. 2011. Equisetum thermale sp. nov. (Equisetales) from the Jurassic San Agustín hot spring deposit, Patagonia: anatomy, paleoecology and inferred paleoecophysiology. American Journal of Botany 98, 680–97.Google Scholar
Drake, B. D., Campbell, K. A., Rowland, J. V., Guido, D. M., Browne, P. R. L. & Rae, A. 2014. Evolution of a dynamic paleo-hydrothermal system at Mangatete, Taupo Volcanic Zone, New Zealand. Journal of Volcanology and Geothermal Research 282, 1935.Google Scholar
Echeveste, H., Fernández, R., Bellieni, G., Tessone, M., Llambias, E., Schalamuk, I., Piccirillo, E. & De Min, A. 2001. Relaciones entre las Formaciones Bajo Pobre y Chon Aike (Jurásico medio a superior) en el área de Estancia El Fénix-Cerro Huemul, zona centro-occidental del Macizo del Deseado, provincia de Santa Cruz. Revista de la Asociación Geológica Argentina 56 (4), 548–58.Google Scholar
Echeveste, H., Fernández, R., Llambias, E., Tessone, M., Schalamuk, I., Bellieni, G., Piccirillo, E. & De Min, A. 1999. Ignimbritas tardías de alto grado en la Formación Chon Aike (Jurásico). Macizo del Deseado, Santa Cruz. Argentina. In XIV Congreso Geológico Argentino, Actas II, pp. 182–5.Google Scholar
Ellis, A. J. & Wilson, S. H. 1961. Hot spring areas with acid–sulphate–chloride waters. Nature 191, 696–8.Google Scholar
Farmer, J. D. 2000. Hydrothermal systems: doorways to early biosphere evolution. GSA Today 10, 19.Google Scholar
Feruglio, E. 1949. Descripción Geológica de la Patagonia. Dirección Nacional de Yacimientos Petrolíferos Fiscales, Buenos Aires, Tomo 1, pp. 17–19.Google Scholar
Fouke, B. W., Farmer, J. D., Des Marais, D. J., Pratt, L., Sturchio, N. C., Burns, P. C. & Discipulo, M. K. 2000. Depositional facies and aqueous-solid geochemistry of travertine depositing hot springs (Angel Terrace, Mammoth Hot Springs, Yellowstone National Park, U.S.A.). Journal of Sedimentary Research 70, 565–85.Google Scholar
Fournier, R. O. 1985. The behaviour of silica in hydrothermal solutions. Reviews in Economic Geology 2, 4562.Google Scholar
Fournier, R. O. & Rowe, J. J. 1966. Estimation of underground temperatures from the silica content of water from hot springs and steam wells. American Journal of Science 264, 685–97.Google Scholar
García Massini, J., Channing, A., Guido, D. M. & Zamuner, A. B. 2012. First report of fungi and fungus-like organisms from Mesozoic hot springs. Palaios 27, 5562.Google Scholar
Giacosa, R., Zubia, M., Sánchez, M. & Allard, J. 2010. Meso-Cenozoic tectonics of the southern Patagonian foreland: structural evolution and implications for Au-Ag veins in the eastern Deseado region (Santa Cruz, Argentina). Journal of South American Earth Sciences 30, 134–50.Google Scholar
Gibert, R. O., Taberner, C., Sáez, A., Giralt, S., Alonso, R. N., Edwards, R. L. & Pueyo, J. J. 2009. Igneous origin of CO2 in ancient and recent hot-spring waters and travertines from the northern Argentinean Andes. Journal of Sedimentary Research 79, 554–67.Google Scholar
Guido, D. 2004. Subdivisión litofacial e interpretación del volcanismo jurásico (Grupo Bahía Laura) en el este del Macizo del Deseado, provincia de Santa Cruz. Revista de la Asociación Geológica Argentina 59 (4), 727–42.Google Scholar
Guido, D. M. & Campbell, K. A. 2009. Jurassic hot-spring activity in a fluvial setting at La Marciana, Patagonia, Argentina. Geological Magazine 146, 617–22.Google Scholar
Guido, D. M. & Campbell, K. A. 2011. Jurassic hot spring deposits of the Deseado Massif (Patagonia, Argentina): characteristics and controls on regional distribution. Journal of Volcanology and Geothermal Research 203, 3547.Google Scholar
Guido, D. M. & Campbell, K. A. 2012. Diverse subaerial and sublacustrine hot springs setting of the Cerro Negro epithermal (Jurassic, Deseado Massif), Patagonia, Argentina. Journal of Volcanology and Geothermal Research 229–230, 112.Google Scholar
Guido, D. M. & Campbell, K. A. 2014. A large and complete Jurassic geothermal field at Claudia, Deseado Massif, Santa Cruz, Argentina. Journal of Volcanology and Geothermal Research 275, 6170.Google Scholar
Guido, D. M., Channing, A., Campbell, K. A. & Zamuner, A. 2010. Jurassic geothermal landscapes and fossil ecosystems at San Agustín, Patagonia, Argentina. Journal of the Geological Society, London 167, 1120.Google Scholar
Guido, D., Delupí, R., López, R., de Barrio, R. & Schalamuk, I. 2002. Estromatolitos y mineralización epitermal en el área Marianas-Eureka, Macizo del Deseado, Santa Cruz. In XV Congreso Geológico Argentino, Calafate, Actas II, pp. 284–9.Google Scholar
Guido, D., Escayola, M., de Barrio, R., Schalamuk, I. & Franz, G. 2006. La Formación Bajo Pobre (Jurásico) en el este del Macizo del Deseado, Patagonia: vinculación con el Grupo Bahía Laura. Revista de la Asociación Geológica Argentina 61 (2), 187–96.Google Scholar
Guido, D. & Schalamuk, I. 2003. Genesis and exploration potential of epithermal deposits from the Deseado Massif, Argentinean Patagonia. In Mineral Exploration and Sustainable Development (eds Eliopoulos, D. et al.), pp. 489–92. Rotterdam: Millpress.Google Scholar
Guo, L. & Riding, R. 1994. Origin and diagenesis of Quaternary travertine shrub fabrics, Rapolano Terme, central Italy. Sedimentology 41, 499520.Google Scholar
Handley, K. & Campbell, K. A. 2011. Character, analysis and preservation of biogenicity in terrestrial siliceous stromatolites from geothermal settings. In Stromatolites: Interaction of Microbes with Sediments (eds Tewari, V. & Seckbach, J.), pp. 359–81. Cellular Origin, Life in Extreme Habitats and Astrobiology 18. Dordrecht: Springer.Google Scholar
Henley, R. W. & Ellis, A. J. 1983. Geothermal systems, ancient and modern: a geochemical review. Earth-Science Reviews 19, 150.Google Scholar
Herdianita, N. R., Browne, P. R. L., Rodgers, K. A. & Campbell, K. A. 2000. Mineralogical and textural changes accompanying ageing of silica sinter. Mineralium Deposita 35, 4862.Google Scholar
Hinman, N. W. & Lindstrom, R. F. 1996. Seasonal changes in silica deposition in hot spring systems. Chemical Geology 132, 237–46.Google Scholar
Jones, B. & Renaut, R. W. 2003a. Petrography and genesis of spicular and columnar geyserite from the Whakarewarewa and Orakeikorako geothermal areas, North Island, New Zealand. Canadian Journal of Earth Science 40, 1585–610.Google Scholar
Jones, B. & Renaut, R. W. 2003b. Hot spring and geyser sinters: the integrated product of precipitation, replacement, and deposition. Canadian Journal of Earth Science 40, 1549–69.Google Scholar
Jones, B. & Renaut, R. W. 2010. Calcareous spring deposits in continental settings. In Carbonates in Continental Settings: Facies, Environments & Processes (eds Alonso-Zarza, A. M. & Tanner, L. H.), pp. 177224. Developments in Sedimentology 61. Amsterdam: Elsevier.Google Scholar
Jones, B. & Renaut, R. W. 2011. Hot springs and geysers. In Encyclopedia of Geobiology (eds Reitner, J. & Thiel, V.), pp. 447–51. Heidelberg: Springer.Google Scholar
Jones, B. & Renaut, R. W. 2012. Facies architecture in depositional systems resulting from the interaction of acidic springs, alkaline springs, and acidic lakes: case study of Lake Roto-a-Tamaheke, Rotorua, New Zealand. Canadian Journal of Earth Science 49, 1217–50.Google Scholar
Jones, B., Renaut, R. W. & Owen, B. O. 2011. Life of a geyser discharge apron: evidence from Waikite Geyser, Whakarewarewa geothermal area, North Island, New Zealand. Sedimentary Geology 236, 7794.Google Scholar
Koban, C. G. & Schweigert, G. 1993. Microbial origin of travertine fabrics – two examples from Southern Germany (Pleistocene Stuttgart travertines and Miocene Riedöschingen travertine). Facies 29, 251–64.Google Scholar
Lynne, B. Y. 2012. Mapping vent to distal-apron hot spring paleo-flow pathways using siliceous sinter architecture. Geothermics 43, 324.Google Scholar
Lynne, B. Y. & Campbell, K. A. 2003. Diagenetic transformations (opal-A to quartz) of low and mid-temperature microbial textures in siliceous hot-spring deposits, Taupo Volcanic Zone, New Zealand. Canadian Journal of Earth Sciences 40, 1679–96.Google Scholar
Lynne, B. Y., Campbell, K. A., James, B. J., Browne, P. R. L. & Moore, J. 2007. Tracking crystallinity in siliceous hot-spring deposits. American Journal of Science 307, 612–41.Google Scholar
Lynne, B. Y., Campbell, K. A., Moore, J. N. & Browne, P. R. L. 2005. Diagenesis of 1900-year-old siliceous sinter (opal-A to quartz) at Opal Mound, Roosevelt Hot Springs, Utah, U.S.A. Sedimentary Geology 179, 249–78.Google Scholar
Lynne, B. Y., Campbell, K. A., Moore, J. & Browne, P. R. L. 2008. Origin and evolution of the Steamboat Springs siliceous sinter deposit, Nevada, U.S.A. Sedimentary Geology 210, 111–31.Google Scholar
Lynne, B. Y., Campbell, K. A., Perry, R. S., Browne, P. R. L. & Moore, J. N. 2006. Acceleration of sinter diagenesis in an active fumarole, Taupo Volcanic Zone, New Zealand. Geology 34, 749–52.Google Scholar
Mykietiuk, K. & Lanfrancini, M. 2004. Depósitos Jurásicos de un lago geotermal en el Cerro Tornillo, Macizo del Deseado, Santa Cruz. In 7 Congreso de Mineralogía y Metalogenia, pp. 261–6.Google Scholar
Nicholson, K. 1993. Geothermal Fluids: Chemistry and Exploration Technique. Berlin: Springer.Google Scholar
Pankhurst, R. J., Leat, P. T., Sruoga, P., Rapela, C. W., Marquez, M., Storey, B. C. & Riley, T. R. 1998. The Chon Aike province of Patagonia and related rocks in West Antarctica: a silicic large igneous province. Journal of Volcanology and Geothermal Research 81, 113–36.Google Scholar
Pentecost, A. 1990. The formation of travertine shrubs: Mammoth Hot Springs, Wyoming. Geological Magazine 127, 159–68.Google Scholar
Pentecost, A. 1993. British travertines: a review. Proceedings of the Geologists’ Association 104, 2939.Google Scholar
Pentecost, A. 2005. Travertine. Berlin: Springer, 445 pp.Google Scholar
Renaut, R. W. & Jones, B. 2000. Microbial precipitates around continental hot springs and geysers. In Microbial Sediments (eds Riding, R. & Awramik, S. M.), pp. 187–95. Berlin: Springer.Google Scholar
Renaut, R. W. & Jones, B. 2011a. Hydrothermal environments, terrestrial. In Encyclopedia of Geobiology (eds Reitner, J. & Thiel, V.), pp. 467–79. Heidelberg: Springer.Google Scholar
Renaut, R. W. & Jones, B. 2011b. Sinter. In Encyclopedia of Geobiology (eds Reitner, J. & Thiel, V.), pp. 808–13. Heidelberg: Springer.Google Scholar
Renaut, R. W., Morley, C. K. & Jones, B. 2002. Fossil hot-spring travertine in the Turkana basin, northern Kenya: Structure, facies, and genesis. In Sedimentation in Continental Rifts (eds Renaut, R. W. & Ashley, G. M.), pp. 123–41. SEPM Special Publication no. 73.Google Scholar
Richardson, N. J. & Underhill, J. R. 2002. Controls on the structural architecture and sedimentary character of syn-rift sequences, North Falkland Basin, South Atlantic. Marine and Petroleum Geology 19, 417–43.Google Scholar
Rodgers, K. A., Browne, P. R. L., Buddle, T. F., Cook, K. L., Greatrex, R. A., Hampton, W. A., Herdianita, N. R., Holland, G. R., Lynne, B. Y., Martin, R., Newton, Z., Pastars, D., Sannazarro, K. L. & Teece, C. I. A. 2004. Silica phases in sinters and residues from geothermal fields of New Zealand. Earth-Science Reviews 66, 161.Google Scholar
Rosen, M. R., Arehart, G. B. & Lico, M. S. 2004. Exceptionally fast growth rate of <100-year-old tufa, Big Soda Lake, Nevada: implications for using tufa as a paleoclimate proxy. Geology 32, 409–12.Google Scholar
Rowland, J. V. & Simmons, S. F. 2012. Hydrologic, magmatic, and tectonic controls on hydrothermal flow, Taupo Volcanic Zone, New Zealand: implications for the formation of epithermal vein deposits. Economic Geology 107 (3), 427–57.Google Scholar
Schalamuk, I. B., Guido, D. M., de Barrio, R. E. & Fernandez, R. R. 1999. Hot spring structures from El Macanudo-El Mirasol area, Deseado Massif, Argentina. In Mineral Deposits: Processes to Processing (eds Stanley, C. J., et al.), pp. 577–80. Rotterdam: Balkema.Google Scholar
Schalamuk, I. B., Zubia, M., Genini, A. & Fernández, R. R. 1997. Jurassic epithermal Au-Ag deposits of Patagonia, Argentina. Ore Geology Reviews 12, 173–86.Google Scholar
Schinteie, R., Campbell, K. A. & Browne, P. R. L. 2007. Microfacies of stromatolitic sinter from acid–sulphate–chloride springs at Parariki Stream, Rotokawa geothermal field, New Zealand. Palaeontologia Electronica 10 (1), 133 (4A), http://palaeoelectronica.org/paleo/2007_1/sinter/index.html.Google Scholar
Sibson, R. H. 1987. Earthquake rupturing as a mineralizing agent in hydrothermal systems. Geology 15, 701–4.Google Scholar
Sillitoe, R. H. 2015. Epithermal paleosurfaces. Mineralium Deposita 50, 767–93.Google Scholar
Sturchio, N. C., Dunkley, P. N. & Smith, M. 1993. Climate-driven variations in the geothermal activity of the northern Kenya Rift Valley. Nature 362, 233–4.Google Scholar
Trewin, N. H. 1996. The Rhynie Chert: an early Devonian ecosystem preserved by hydrothermal activity. In Evolution of Hydrothermal Ecosystems on Earth (and Mars?) (eds Bock, G. R. & Goode, G. A.), pp. 131–49. Proceedings of the Ciba Foundation Symposium 202. Chichester: J. Wiley.Google Scholar
Trewin, N. H., Fayers, S. R. & Kelman, R. 2003. Subaqueous silicification of the contents of small ponds in an early Devonian hot spring complex, Rhynie, Scotland. Canadian Journal of Earth Science 40, 1697–712.Google Scholar
White, D. E. 1967. Some principles of geyser activity, mainly from Steamboat Springs, Nevada. American Journal of Science 265, 641–84.Google Scholar