Hostname: page-component-848d4c4894-r5zm4 Total loading time: 0 Render date: 2024-06-30T00:31:52.559Z Has data issue: false hasContentIssue false

The Formation of Clay Minerals in the Mudflats of Bolivian Salars

Published online by Cambridge University Press:  01 January 2024

Jennifer L. Bentz*
Affiliation:
Department of Geological Sciences and Geological Engineering, Queen’s University, 36 Union Street, Kingston, ON K7L 3N6, Canada
Ronald C. Peterson
Affiliation:
Department of Geological Sciences and Geological Engineering, Queen’s University, 36 Union Street, Kingston, ON K7L 3N6, Canada
*
*E-mail address of corresponding author: jennifer.bentz@queensu.ca

Abstract

Understanding clay-mineral assemblages forming in saline lakes aids in reconstructing paleoenvironments on Earth and other terrestrial planets; this is because authigenic phyllosilicates are sensitive to the prevailing geochemical conditions present during formation. In most geochemical models, evaporative concentration favors sepiolite with increasing silica and Mg2+ concentrations without considering the role of the biogenic removal of silica from solution by diatoms. In the present study, phyllosilicates occurring in the mudflats of Bolivian salars were investigated to aid in understanding the geochemical factors that control mineral assemblages forming in (SO42–)- and (Cl)-rich environments in relation to dissolved silica. From transects across the mudflats, the physical, chemical, and mineralogical characteristics of the bulk sediment and the <2 μm fraction of each sedimentary layer were analyzed. From these analyses, three types of sediments were identified: (1) regolith sediments dominated by Al-dioctahedral smectite, illite, and chlorite; (2) detritus-rich mudflat sediments with Mg-trioctahedral smectite and Al-dioctahedral smectite along with illite and chlorite; and (3) authigenic mudflat sediments dominated by poorly formed Mg-trioctahedral smectite, kerolite, and biogenic silica. The absence of sepiolite-palygorskite in the salars is the result of excessively high Mg:Si ratios within the waters. In the surface water Mg becomes enriched relative to Si as diatoms remove dissolved Si from solution through biologically mediated uptake. The geochemical conditions present within the salars that act to preserve the diatom frustules and prevent their dissolution include: neutral–slightly alkaline pH solutions, cold temperatures, shallow water, and high salinity. Under these conditions the formation of sepiolite is restricted by the small amount of dissolved silica, despite the silica-rich environment. The formation of Mg-smectite and kerolite is favored under these conditions.

Type
Article
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

Akbulut, A., & Kadir, S. (2003). The geology & origin of sepiolite, palygorskite, & saponite in Neogene lacustrine sediments of the Serinhisar-Acipayam basin, Denizli, SW Turkey. Clays & Clay Minerals, 3, 279292.CrossRefGoogle Scholar
Allmendinger, R. W., Jordan, T. E., Kay, S. M., & Isacks, B. L. (1997). The evolution of the Altiplano-Puna plateau of the Central & es. Annual Review of Earth & Planetary Science, 25, 139174.CrossRefGoogle Scholar
Argollo, J., & Mourguiart, P. (2000). Late Quaternary climate history of the Bolivian Altiplano. Quaternary International, 72, 3751.CrossRefGoogle Scholar
Badaut, D., & Risacher, F. (1983). Authigenic smectite on diatom frustules in Bolivian saline lakes. Geochimica et Cosmochimica Acta, 47, 363375.CrossRefGoogle Scholar
Banfield, J. F., Jones, B. F., & Veblen, D. R. (1991). An AEM-TEM study of weathering & diagenesis, Abert Lake, Oregon: 2. Diagenetic modification of the sedimentary assemblage. Geochimica et Cosmochimica Acta, 55, 27952810.CrossRefGoogle Scholar
Bellanca, A., Calvo, J. P., Censi, P., Neri, R., & Pozo, M. (1992). Recognition of lake-level changes in Miocene lacustrine units Madrid basin, Spain. Evidence from facies analysis, isotope geochemistry & clay mineralogy. Sedimentary Geology, 76, 135153.CrossRefGoogle Scholar
Birsoy, R. (2002). Formation of sepiolite-palygorskite & related minerals from solution. Clays & Clay Minerals, 50, 736745.CrossRefGoogle Scholar
Bishop, J. L., Pieters, C. M., & Edwards, J. O. (1994). Infrared spectroscopic analysis on the nature of water in montmorillonite. Clays & Clay Minerals, 42, 702716.CrossRefGoogle Scholar
Bishop, J., Madejová, J., Komadel, P., & Fröschl, H. (2002a). The influence of structural Fe, Al & Mg on the infrared OH b&s in spectra of dioctahedral smectites. Clay Minerals, 37, 607616.CrossRefGoogle Scholar
Bishop, J., Murad, E., & Dyar, M. D. (2002b). The influence of octahedral & tetrahedral cation substitution on the structure of smectites & serpentines as observed through infrared spectroscopy. Clay Minerals, 37, 617628.CrossRefGoogle Scholar
Bishop, J. L., Rampe, E. B., Bish, D. L., Abidin, Z., Baker, L. L., Matsue, N., & Henmi, T. (2013). Spectral & hydration properties of allophane & imogolite. Clays & Clay Minerals, 61, 5774.CrossRefGoogle Scholar
Bowen, B. B., & Benison, K. C. (2009). Geochemical characteristics of naturally acid & alkaline saline lakes in southern Western Australia. Applied Geochemistry, 24, 268284.CrossRefGoogle Scholar
Brindley, G. W., Bish, D. L., & Wan, H.-M. (1977). The nature of kerolite, its relation to talc & stevensite. Mineralogical Magazine, 41, 443452.CrossRefGoogle Scholar
Bristow, T. F., & Milliken, R. E. (2011). Terrestrial perspective on authigenic clay mineral production in ancient Martian lakes. Clays & Clay Minerals, 59, 339358.CrossRefGoogle Scholar
Bustillo, M. A., & Bustillo, M. (2000). Miocence silcretes in argillaceous playa deposits, Madrid basin, Spain: petrological & geochemical features. Sedimentology, 47, 10231037.CrossRefGoogle Scholar
Cabrol, N. A., Grin, E. A., Chong, G., Minkley, E., Hock, A. N., Yu, Y., Bebout, L., Fleming, E., Häder, D. P., Demergasso, C., Gibson, J., Escudero, L., Dorador, C., Lim, D., Woosley, C., Morris, R. L., Tambly, C., Gaete, V., Galvez, M. E., Smith, E., Uskin-Peate, I., Salazar, C., Dawidowicz, G., & Majerowicz, J. (2009). The high-lakes project. Journal of Geophysical Research: Biogeosciences, 114, G00D06. https://doi.org/10.1029/2008JG000818.CrossRefGoogle Scholar
Calvo, J.P., Blanc-Valleron, M.M., Rodrigue-Ar&ia, J.P., Rouchy, J.M., & Sanz, M.E. (1999). Authigenic clay minerals in continental evaporitic environments. Pp 129151 in: Paleoweathering, Paleosurfaces & Related Continental Deposits (Thiry, M. & Simon-Coincon, R., editors). Special Publication of the International Association of Sedimentologists, 27, Blackwell Sciences, Oxford, UK.Google Scholar
Chahi, A., Fritz, B., Duplay, J., Weber, F., & Lucas, J. (1997). Textural transition & genetic relationship between precursor stevensite & sepiolite in lacustrine sediments (Jbel Rhassoul, Morocco). Clays & Clay Minerals, 45, 378389.CrossRefGoogle Scholar
Clark, R.N. (1999). Spectroscopy of rocks & minerals, & principles of spectroscopy. Pp. 358 in: Manual of Remote Sensing, Volume 3, Remote Sensing for the Earth Sciences (Rencz, A.N., editor). John Wiley & Sons, New York.Google Scholar
Cuevas, J., Vigil de la Villa, R., Ramirez, S., Petit, S., Meunier, A., & Leguey, S. (2003). Chemistry of Mg smectites in lacustrine sediments from the Vicalvaro sepiolite deposit, Madrid Neogene basin (Spain). Clays & Clay Minerals, 51, 457472.CrossRefGoogle Scholar
Darragi, F., & Tardy, Y. (1987). Authigenic trioctahedral smectites controlling pH, alkalinity, silica, & magnesium concentrations in alkaline lakes. Chemical Geology, 63, 5972.CrossRefGoogle Scholar
Degen, M., Sadki, M., Bron, E., Köng, U., & Nénert, G. (2014). The Highscore suite. Powder Diffraction, Supplement S2, 29 S13–S18.CrossRefGoogle Scholar
Dejoux, C. (1993). Benthic invertebrates of some saline lakes of the Sud Lipez region, Bolivia. Hydrobiologia, 267, 257267.CrossRefGoogle Scholar
DeMaster, D.J. (2003). The diagenesis of biogenic silica: chemical transformations occurring in the water column, seabed, & crust. Pp. 8796 in: Treatise on Geochemistry, Volume 7.04, Sediments, Diagenesis, & Sedimentary Rocks (H.D Holl& & K.K Turekian editor). Elsevier.Google Scholar
Deocampo, D. M. (2005). Evaporative evolution of surface waters & the role of aqueous CO2 in magnesium silicate precipitation: Lake Eyasi & Ngorongoro Crater, northern Tanzania. South African Journal of Geology, 108, 493504.CrossRefGoogle Scholar
Deocampo, D. (2015). Authigenic clay minerals in lacustrine mudstones. Pp. 4964 in: Paying Attention to Mudrocks: Priceless!: Geological Society of America Special Paper 515 (Larson, D., Egenhoff, S.O., & Fishman, N.S., editors). Geological Society of America.CrossRefGoogle Scholar
Deocampo, D. M., Cuadros, J., Wing-Dudek, T., Olives, J., & Amouric, M. (2009). Saline lake diagenesis as revealed by coupled mineralogy & geochemistry of multiple ultrafine clay phases: Pliocene Olduvai Gorge, Tanzania. American Journal of Science, 309, 834868.CrossRefGoogle Scholar
Deocampo, D. M., Behrensmeyer, A. K., & Potts, R. (2010). Ultrafine clay minerals of the Pleistocene Olorgesailie Formation, southern Kenya rift: diagenesis & paleoenvironments of early hominins. Clays & Clay Minerals, 58, 294310.CrossRefGoogle Scholar
Ece, Ö. I., & Çoban, F. (1994). Geology, occurrence, & genesis of Eskişehir sepiolite, Turkey. Clays & Clay Minerals, 42, 8192.CrossRefGoogle Scholar
Elzea, J. M., Odom, I. E., & Miles, W. J. (1994). Distinguishing well ordered opal-CT & opal C from high temperature cristobalite by X-ray diffraction. Analytica Chimica Acta, 286, 107116.CrossRefGoogle Scholar
Farr, W. H., Glotch, T.D., Rice, J. W., Hurowitz, J. A., & Swayze, G. A. (2009). Discovery of jarosite within the Mawrth Vallis region of Mars: implications for the geological history of the region. Icarus, 204, 478488.Google Scholar
Flower, R. J., & Ryves, D. B. (2009). Diatom preservation: differential preservation of sedimentary diatoms in two saline lakes. Acta Botanica Croatica, 68, 381399.Google Scholar
Furquim, S. A. C., Barbiéro, L., Graham, R. C., Neto, J. P. Q., Ferreira, R. P. D., & Furian, S. (2010). Neoformation of micas in soils surrounding an alkaline-saline lake of Pantanal Wetl & Brazil. Geoderma, 158, 331342.CrossRefGoogle Scholar
Gac, J. Y., Droubi, A., Fritz, B., & Tardy, Y. (1977). Geochemical behaviour of silica & magnesium during the evaporation of waters in Chad. Chemical Geology, 19, 215228.CrossRefGoogle Scholar
Galán, E. & Castillo, A. (1984). Sepiolite-palygorskite in Spanish tertiary basins: Genetical patterns in continental environments. Pp. 87124 in: Palygorskite-Sepiolite Occurrences, Genesis, & Uses (Singer, A. & Galán, E., editors). Developments in Sedimentology, 37, Elsevier, Amsterdam, The Netherl&s.Google Scholar
Galán, E. & Pozo, M. (2011). Palygorskite & sepiolite deposits in continental environments. Description, genetic patterns, & sedimentary settings. Pp. 125173 in: Developments in Clay Science, Vol. 3, Developments in Palygorskite-Sepiolite Research, A New Outlook on these Nanomaterials (Galán, E., & Singer, A., editors). Elsevier, Amsterdam, the Netherl&s.Google Scholar
García-Romero, E., Suárez, M., Santaré, J., & Alvarez, A. (2007). Crystallochemical characterization of the palygorskite & sepiolite from the Allou Kagne deposit, Senegal. Clays & Clay Minerals, 55, 606617.CrossRefGoogle Scholar
Gürel, A., & Özcan, S. (2016). Paleosol & dolocrete associated clay mineral occurrences in siliciclastic red sediments of the Late Miocene Kömişini Formation of the Tuzgölü basin in central Turkey. Catena, 143, 102113.CrossRefGoogle Scholar
Hardie, L. A., & Eugster, H. P. (1970). The evolution of closed-basin brines. Mineralogical Society of America Special Paper, 3, 273290.Google Scholar
Hay, R. L. (1970). Silicate reactions in three lithofacies of a semi-arid basin, Olduvai George, Tanzania. Mineralogical Society of America Special Paper, 3, 237255.Google Scholar
Hay, R. L., & Kyser, T. K. (2001). Chemical sedimentology & paleoenvironmental history of Lake Olduvai, a Pliocene lake in northern Tanzania. Bulletin of the Geological Society of America, 113, 15051521.2.0.CO;2>CrossRefGoogle Scholar
Hay, R. L., Guldman, S. G., Matthews, J. C., L&er, R. H., Duffin, M. E., & Kyser, T. K. (1991). Clay mineral diagenesis in Core Km-3 of Searles Lake, California. Clays & Clay Minerals, 39, 8496.CrossRefGoogle Scholar
Hecky, R. E., & Kilham, P. (1973). Diatoms in alkaline saline lakes: ecology & geochemical implications. Limnology & Oceanography, 18, 5371.CrossRefGoogle Scholar
Hover, V. C., & Ashley, G. M. (2003). Geochemical signatures of paleodepositional & diagenetic environments: a STEM/AEM study of authigenic clay minerals from an arid rift basin, Olduvai Gorge, Tanzania. Clays & Clay Minerals, 51, 231251.CrossRefGoogle Scholar
Hover, V. C., Walter, L. M., Peacor, D. R., & Martini, A. M. (1999). Mg-smectite authigenesis in a marine evaporative environment, Saline Ometepec, Baja, California. Clays & Clay Minerals, 47, 252268.CrossRefGoogle Scholar
Huggett, J., Cuadros, J., Gale, A., Wray, D., & Adetunji, J. (2016). Low temperature, authigenic illite & carbonates in a mixed dolomite-clastic lagoonal & pedogenic setting, Spanish Central System, Spain. Applied Clay Science, 132–133, 296312.CrossRefGoogle Scholar
Jones, B.F. (1986). Clay mineral diagenesis in lacustrine sediments. Pp. 291–300 in: Studies in Diagenesis (Mumpton, F.A., editor). Bulletin of the US Geological Survey, 1578.Google Scholar
Jones, B.F. & Galán, E. (1988). Sepiolite & palygorskite. Pp. 631674 in: Hydrous Phyllosilicates (Exclusive of Micas) (Bailey, S.W., editor). Reviews in Mineralogy, 19, Mineralogical Society of America, Washington, D.C.CrossRefGoogle Scholar
Jones, B. F., & Weir, A. H. (1983). Clay mineral of Lake Abert, an alkaline, saline lake. Clays & Clay Minerals, 31, 161172.CrossRefGoogle Scholar
Kadir, S., Eren, M., & Atabey, E. (2010). Dolocretes & associated palygorskite occurrences in siliciclastic red mudstones of the Sariyer Formation (Middle Miocene), southeastern side of the Çanakkale Strait, Turkey. Clays & Clay Minerals, 58, 205219.CrossRefGoogle Scholar
Kadir, S., Erkoyun, H., Erin, M., Hugettt, J., & Önalgil, N. (2016). Palygorskite in Neogene lacustrine sediments, Eskişehir Province, West Central Anatolia, Turkey. Clays & Clay Minerals, 64, 145166.CrossRefGoogle Scholar
Kalro, Y. P. (1995). Determination of pH of soils by different methods: collaborative study. Journal of AOAC International, 78, 310324.CrossRefGoogle Scholar
Larsen, D. (2008). Revisiting silicate authigenesis in the Pliocence-Pleistocene Lake Tecopa beds, southeastern California: Depositional & hydrological controls. Geosphere, 4, 612639.CrossRefGoogle Scholar
Léveillé, R. J., Longstaffe, F. J., & Fyfe, W. S. (2002). Kerolite in carbonate-rich speleothems & microbial deposits from basaltic caves, Kauai, Hawaii. Clays & Clay Minerals, 50, 517527.CrossRefGoogle Scholar
Loucaides, S., Van Cappellen, P., Roubeix, V., Moriceau, B., & Ragueneau, O. (2012). Controls on the recycling & preservation of biogenic silica from biomineralization to burial. Silicon, 4, 722.CrossRefGoogle Scholar
Martin De Vidales, J. L., Pozo, M., Alia, J. M., Garcia-Navarro, F., & Rull, F. (1991). Kerolite-stevensite in mixed layers from the Madrid Basin, central Spain. Clay Minerals, 26, 329342.CrossRefGoogle Scholar
Meunier, A. (2005). Clays. Springer, Berlin, 472 pp.Google Scholar
Michalopouos, P., Aller, R. C., & Reeder, R. J. (2000). Conversion of diatoms to clays during early diagenesis in tropical, continental shelf muds. Geology, 28, 10951098.2.0.CO;2>CrossRefGoogle Scholar
Milliken, R. E., Swayze, G. A., Arvidson, R. E., Bishop, J. L., Clark, R. N., Ehlmann, B. L., Green, R. O., Grotzinger, J. P., Morris, R. V., Murchie, S. L., Mustard, J. F., & Weitz, C. (2008). Opaline silica in young deposits on Mars. Geology, 36, 847850. https://doi.org/10.1130/G24967A.1.CrossRefGoogle Scholar
Milliken, R.E., Grotzinger, J.P., & Thomson, B.J. (2010). Paleoclimate of Mars as captured by the stratigraphic record in Gale Crater. Geophysical Research Letters 37, https://doi.org/10.1029/2009GL041870.fCrossRefGoogle Scholar
Moore, D. M., & Reynolds, R. C. (1989). X-Ray Diffraction & the Identification & Analysis of Clay Minerals (pp. 179201). New York: Oxford University Press.Google Scholar
Papke, K. G. (1972). A sepiolite-rich playa deposit in southern Nevada Clays & Clay Minerals, 20, 211215.CrossRefGoogle Scholar
Pozo, M., & Casas, J. (1999). Origin of kerolite & associated Mg clays in palustrine-lacustrine environments. The Esquivias deposit (Neogene Madrid basin, Spain). Clay Minerals, 34, 395418.CrossRefGoogle Scholar
Pozo, M. & Calvo, J.P. (2018). An overview of authigenic magnesian clays. Minerals, 8, https://doi.org/10.3390/min8110520.CrossRefGoogle Scholar
Risacher, F., & Fritz, B. (1991). Geochemistry of Bolivian salars, Lipez, southern Altiplano: Origin of solutes & brine evolution. Geochimica et Cosmochimica Acta, 55, 687705.CrossRefGoogle Scholar
Ryves, D. B., Battarbee, R. W., Juggins, S., Fritz, S. C., & & erson, N. J. (2006). Physical & chemical predictors of diatom dissolution in freshwater & saline lake sediments in North America & West Greenl. Limnology & Oceanography, 51, 13551368.CrossRefGoogle Scholar
Santiago de Buey, C., Barrios, M. S., Romero, E. G., & Montoya, M. D. (2000). Mg-rich smectite “precursor” phase in the Tagus basin, Spain. Clays & Clay Minerals, 48, 366373.CrossRefGoogle Scholar
Servant-Vildary, S., & Roux, M. (1990). Multivariate analysis of diatoms & water chemistry in Bolivian saline lakes. Hydrobiologia, 197, 267290.CrossRefGoogle Scholar
Stoessell, R. K., & Hay, R. L. (1978). The geochemical origin of sepiolite & kerolite at Amboseli, Kenya. Contributions to Mineralogy & Petrology, 65, 255267.CrossRefGoogle Scholar
Story, S., Bowen, B.B., Benison, K.C. & Schulze, D.G. (2010). Authigenic phyllosilicates in modern acid saline lake sediments & implications for Mars. Journal of Geophysical Research, 115, https://doi.org/10.1029/2010JE003687CrossRefGoogle Scholar
Sylvestre, F., Servant-Vildary, S., & Roux, M. (2001). Diatom-based ionic concentration & salinity models from the south Bolivian Altiplano (15–23°S). Journal of Paleolimnology, 25, 279295.CrossRefGoogle Scholar
Tosca, N.J. (2015). Geochemical pathways to Mg-silicate formation. Pp. 283330 in: Magnesian Clays: Characterization, Origin, & Applications (Pozo, M. & Galan, E., editors). Bari, Italy.Google Scholar
Tosca, N. J., & Masterson, A. L. (2014). Chemical controls onincipient Mg-silicate crystallization at 25°C: Implications for early & late diagenesis. Clay Minerals, 49, 165194.CrossRefGoogle Scholar
Tosca, N. J., Macdonald, F. A., Strauss, J. V., Johnston, D. T., & Knoll, A. H. (2011). Sedimentary talc in Neoproterozoic carbonate successions. Earth & Planetary Science Letters, 306, 1122.CrossRefGoogle Scholar
Turner, C. E., & Fishman, N. S. (1991). Jurassic Lake T'oo‘dichi’: A large alkaline, saline, lake, Morrison Formation, eastern Colorado Plateau. Bulletin of the Geological Society of America, 103, 538558.2.3.CO;2>CrossRefGoogle Scholar
Tutulo, B. M., & Tosca, N. J. (2018). Experimental examination of the Mg-silicate-carbonate system at ambient temperature: Implications for alkaline chemical sedimentation & lacustrine carbonate formation. Geochimica et Cosmochimica Acta, 225, 80101.CrossRefGoogle Scholar
Weaver, C.E. (1989). Clays, Muds, & Shales. 820 pp. Elsevier, Amsterdam.Google Scholar
Wray, J. J., Squyres, S. W., Roach, L. H., Bishop, J. L., Mustard, J. F., & Dobrea, E. Z. N. (2010). Identification of the Ca-sulfate bassanite in Mawrth Vallis, Mars. Icarus, 209, 416421.CrossRefGoogle Scholar
Wray, J.J., Milliken, R.E., Dundas, C.M., Swayze, G.A., rews-Hanna, J.C., Baldridge, A.M., Chojnacki, M., Bishop, J.L., Ehlmann, B.L., Murchie, S.L., Clark, R.N., Seelos, F.P., Tornabene, L.L., & Squyres, S.W. (2011). Columbus crater & other possible groundwater-fed paleolakes of Terra Sirenum, Mars. Journal of Geophysical Research, 116, https://doi.org/10.1029/2010JE003694CrossRefGoogle Scholar