Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-11T22:03:40.566Z Has data issue: false hasContentIssue false

Arsenic-Bearing Serpentine-Group Minerals: Mineral Synthesis with Insights for the Arsenic Cycle

Published online by Cambridge University Press:  01 January 2024

P. C. Ryan*
Affiliation:
Department of Geology, Middlebury College, 276 Bicentennial Way, Middlebury, VT 05763, USA Instituto Andaluz de Ciencias de la Tierra, Avda. Palmeras 4, Armilla, 18100 Granada, Spain
F.J. Huertas
Affiliation:
Instituto Andaluz de Ciencias de la Tierra, Avda. Palmeras 4, Armilla, 18100 Granada, Spain
L. N. Pincus
Affiliation:
School of Forestry and Environmental Studies, Yale University, 195 Prospect Street, New Haven, CT 06511, USA
W. Painter
Affiliation:
Department of Geology, Middlebury College, 276 Bicentennial Way, Middlebury, VT 05763, USA
*
*E-mail address of corresponding author: pryan@middlebury.edu

Abstract

When present at elevated levels in drinking water, arsenic is toxic, and magnesian clays are gaining recognition as a source of elevated arsenic in groundwater. In the crust and upper mantle of Earth, arsenic incorporation into clay minerals is influenced by geochemical conditions associated with hydrothermal fluids and metamorphic processes (e.g. serpentinization), meaning that As is a useful tracer of fluid-flow in the deep Earth. To improve understanding of arsenic speciation in groundwater, sediments, soils, and hydrothermal-metamorphic systems, the present study examined arsenic incorporation into magnesian clays by synthesis of serpentine minerals (200oC, 10 d) with varied concentrations of Si, Al, As5+, and As3+. The synthesis experiments produced two distinct crystal types, tubular and platy serpentines, each with 10–15% randomly interstratified talc layers. X-ray absorption spectroscopy indicated that As5+ and As3+ occurred in the tetrahedral sheet. Single-crystal analysis revealed that tubular crystals contained up to 1 wt.% arsenic [Mg2.8(Si1.8As0.2)O5(OH)4] (mean 0.2 wt.% As). The mean composition of platy, high-Al crystals is (Mg1.8Al0.7)(Si2.0)O5(OH)4, and that of platy, medium-Al crystals with As3+ is (Mg2.07Al0.52) (Si1.97As3+0.03)O5(OH)4. Charge, geometry, and radius of tetrahedral AsO43– oxyanions are similar to tetrahedral SiO44–, and this facilitates fixation of As5+ into the tetrahedral sheet of clay minerals. The geometry and size of the larger As3+ in tetrahedral sites (as a pyramidal AsO33– oxyanion) may limit incorporation relative to As5+. Arsenic-bearing Mg clays crystallize in alkaline environments where AsO43– or AsO33– are the dominant As species and where high pH accompanies crystallization of serpentine, talc, chlorite, or Mg-smectite. The presence of tetrahedral As in these clays raises the possibility of tetrahedral As in other Mg clays (e.g. sepiolite or kerolite) as well.

Type
Article
Copyright
Copyright © Clay Minerals Society 2019

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.)

Footnotes

A note on terminology of arsenic species: to avoid confusion with the often imprecise terms arsenate and arsenite, this article uses “As5+” and “As3+” as much as possible. As used herein, these are effectively equivalent to As(V) and As(III).

References

Arai, Y., Elzinga, E. J., & Sparks, D. L. (2001). X-ray absorption spectroscopic investigation of arsenite and arsenate adsorption at the aluminum oxide–water interface. Journal of Colloid and Interface Science, 235, 8088.CrossRefGoogle ScholarPubMed
Ballantyne, M., & Moore, J. N. (1988). Arsenic geochemistry in geothermal systems. Geochimica et Cosmochimica Acta, 52, 475483.CrossRefGoogle Scholar
Barnes, I., & O'Neil, J. R. (1969). The relationship between fluids in some fresh alpine type ultramafics and possible modern serpentinization, western United States. Geological Society of America Bulletin, 80, 19471960.CrossRefGoogle Scholar
Bentabol, M., Ruiz Cruz, M. D., Huertas, F. J., & Linares, J. (2006). Hydrothermal synthesis of Mg- and Mg-Ni-rich kaolinite. Clays and Clay Minerals, 54, 667677.CrossRefGoogle Scholar
Bentabol, M., Ruiz Cruz, M. D., & Huertas, F. J. (2007). Synthesis of Ni-rich 1: 1 phyllosilicates. Clays and Clay Minerals, 55, 572582.CrossRefGoogle Scholar
Bentabol, M., Ruiz Cruz, M. D., & Huertas, F. J. (2009). Hydrothermal synthesis (200°C) of Co–kaolinite and Al–Co–serpentine. Applied Clay Science, 42, 649656.CrossRefGoogle Scholar
Bentabol, M., Ruiz Cruz, M., & Sobrados, I. (2010). Chemistry, morphology and structural characteristics of synthetic Al-lizardite. Clay Minerals, 45, 131143.CrossRefGoogle Scholar
Bhattacharya, P., Welch, A. H., Stollenwerk, K. G., McLaughlin, M., Bundschuh, J., & Panaullah, G. (2007). Arsenic in the environment: biology and chemistry. Science of the Total Environment, 379, 109120.CrossRefGoogle ScholarPubMed
Boskabadi, A., Pitcairn, I. K., Broman, C., Boyce, A., Teagle, D. A. H., Cooper, M. J., Azer, M. K., Stern, R. J., Mohamed, F. H., & Majka, J. (2017). Carbonate alteration of ophiolitic rocks in the Arabian–Nubian Shield of Egypt: sources and compositions of the carbonating fluid and implications for the formation of Au deposits. International Geology Review, 59, 391419.CrossRefGoogle Scholar
Breuer, C., & Pichler, T. (2013). Arsenic in marine hydrothermal fluids. Chemical Geology, 348, 214.CrossRefGoogle Scholar
Brindley, G. W., & Brown, G. (1980). Crystal Structures of Clay Minerals and their X-ray Identification. Monograph 5, Mineralogical Society, London.CrossRefGoogle Scholar
Capitani, G., & Mellini, M. (2004). The modulated crystal structure of antigorite: The m=17 polysome. American Mineralogist, 89, 147158.CrossRefGoogle Scholar
Čavajda, V., Uhlík, P., Derkowski, A., Čaplovičová, M., Madejová, J., Mikula, M., & Ifka, T. (2015). Influence of grinding and sonication on the crystal structure of talc. Clays and Clay Minerals, 63, 311327.CrossRefGoogle Scholar
Charnock, J. M., Polya, D. A., Gault, G., & Wogelius, R. (2007). Direct EXAFS evidence for incorporation of As5+ in the tetrahedral site of natural andraditic garnet. American Mineralogist, 92, 18561861.CrossRefGoogle Scholar
Chukanov, N. V. (2014). Infrared Spectra of Mineral Species: Extended Library, volume 1. Berlin: Springer.CrossRefGoogle Scholar
Cliff, G., & Lorimer, G. W. (1975). The quantitative analysis of thin specimens. Journal of Microscopy, 103, 203207.CrossRefGoogle Scholar
Craw, D., Landis, C. A., & Kelsey, P. I. (1987). Authigenic chrysotile formation in the matrix of Quaternary debris flows, northern Southland, New Zealand. Clays and Clay Minerals, 35, 4352.CrossRefGoogle Scholar
Deer, W. A., Howie, R. A., & Zussman, J. (2009). Rock-Forming Minerals Vol. 3B, Layered Silicates Excluding Micas and Clay Minerals. 2nd edition. London: Geological Society.Google Scholar
Deschamps, F., Guillot, S., Godard, M., Chauvel, C., Andreani, M., & Hattori, K. (2010). In situ characterization of serpentinites from forearc mantle wedges: timing of serpentinization and behavior of fluid-mobile elements in subduction zones. Chemical Geology, 269, 262277.CrossRefGoogle Scholar
Dódony, I., Pósfai, M., & Buseck, P. R. (2002). Revised structure models for antigorite: An HRTEM study. American Mineralogist, 87, 14431457.CrossRefGoogle Scholar
Farmer, V. C. (1974). The layer silicates. In Farmer, V. C. (Ed.), The Infrared Spectra of Minerals (pp. 331365). London: Mineralogical Society.CrossRefGoogle Scholar
Farmer, V. C., & Russell, J. D. (1967). Infrared absorption spectrometry in clay studies. Clays and Clay Minerals, 15, 121142.CrossRefGoogle Scholar
Frost, R. L., Xi, Y., Pogson, R. E., & Scholz, R. (2013). A vibrational spectroscopic study of philipsbornite PbAl3(AsO4)(OH)5.H2 Omolecular structural implications and relationship to the crandallite subgroup arsenates. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 104, 257261.CrossRefGoogle Scholar
Fuchs, Y., Linares, J., & Mellini, M. (1998). Mössbauer and infrared spectrometry of lizardite-1T from Monte Fico, Elba. Physics and Chemistry of Minerals, 26, 111115.CrossRefGoogle Scholar
Guillot, S., & Charlet, L. (2007). Bengal arsenic, an archive of Himalaya orogeny and paleohydrology. Journal of Environmental Science and Health A, 42, 17851794.CrossRefGoogle ScholarPubMed
Hattori, K., Takahashi, Y., Guillot, S., & Johanson, B. (2005). Occurrence of arsenic (V) in forearc mantle serpentinites based on X-ray absorption spectroscopy study. Geochimica et Cosmochimica Acta, 69, 55855596.CrossRefGoogle Scholar
Inoue, A. (1995). Formation of clay minerals in hydrothermal environments. In Velde, B. (Ed.), Origin and Mineralogy of Clays (pp. 268329). Berlin: Springer.CrossRefGoogle Scholar
Iriarte, 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.CrossRefGoogle Scholar
Ishimaru, S., & Arai, S. (2008). Arsenide in a metasomatized peridotite xenolith as a constraint on arsenic behavior in the mantle wedge. American Mineralogist, 93, 10611065.CrossRefGoogle Scholar
James-Smith, J., Cauzid, J., Testamale, D., Liu, W., Hazemann, J.-L., Proux, O., Etschmann, B., Philipott, P., Banks, D., Williams, P., & Brugger, J. (2010). Arsenic speciation in fluid inclusions using micro-beam X-ray absorption spectroscopy. American Mineralogist, 95, 921932.CrossRefGoogle Scholar
Jones, B. F., & Conko, K. M. (2011). Environmental influences on the occurrences of sepiolite and palygorskite: A brief review. Developments in Clay Science, 3, 6983.CrossRefGoogle Scholar
Lafay, R., Montes-Hernandez, G., Janots, E., Munoz, M., Auzende, A. L., Gehin, A., Chiriac, R., & Proux, O. (2016). Experimental investigation of As, Sb and Cs behavior during olivine serpentinization in hydrothermal alkaline systems. Geochimica et Cosmochimica Acta, 179, 177202.CrossRefGoogle Scholar
Liu, S., Jing, C., & Meng, X. (2008). Arsenic re-mobilization in water treatment adsorbents under reducing conditions: Part II. XAS and modeling study. Science of The Total Environment, 392, 137144.CrossRefGoogle ScholarPubMed
López, D. L., Birkle, P., Bundschuh, J., Sracek, O., Armienta, M. A., Cornejo, L., & Ormachea, M. (2012). Arsenic in geothermal waters of volcanic–magmatic systems of Latin America. Science of the Total Environment, 429, 5775.CrossRefGoogle ScholarPubMed
Masuda, H. (2018). Arsenic cycling in the Earth's crust and hydrosphere: interaction between naturally occurring arsenic and human activities. Progress in Earth and Planetary Science, 5, 68.CrossRefGoogle Scholar
Masuda, H., Shinoda, K., Okudaira, T., Takahashi, Y., & Noguchi, N. (2012). Chlorite – source of arsenic groundwater pollution in the Holocene aquifer of Bangladesh. Geochemical Journal, 46, 381391.CrossRefGoogle Scholar
Mellini, M., Fuchs, Y., Viti, C., Lemaire, C., & Linares, J. (2002). Insights into the antigorite structure from Mössbauer and FTIR spectroscopies. European Journal of Mineralogy, 14, 97104.CrossRefGoogle Scholar
Méring, J. (1949). L'intereférence des Rayons X dans les systèmes à stratification désordonée. Acta. Crystallographica, 2, 371377.CrossRefGoogle Scholar
Meunier, A. (2005). Clays. Berlin: Springer-Verlag.Google Scholar
Milham, L., & Craw, D. (2009). Two-stage structural development of a Paleozoic auriferous shear zone at the Globe-Progress deposit, Reefton, New Zealand. New Zealand Journal of Geology and Geophysics, 52, 247259.CrossRefGoogle Scholar
Moore, D. M., & Reynolds, R. C. Jr. (1997). X-ray Diffraction and the Identification and Analysis of Clay Minerals. New York: Oxford University Press.Google Scholar
Mukherjee, A., Verma, S., Gupta, S., Henke, K., & Bhattacharya, P. (2014). Influence of tectonics, sedimentation and aqueous flow cycles on the origin of global groundwater arsenic: paradigms from three continents. Journal of Hydrology, 518, 284299.CrossRefGoogle Scholar
Müller, K., Ciminelli, V. S. T., Dantas, M. S. S., & Willscher, S. (2010). A comparative study of As(III) and As(V) in aqueous solutions and adsorbed on iron oxy-hydroxides by Raman spectroscopy. Water Research, 44, 56605672.CrossRefGoogle Scholar
Myneni, S. C. B., Traina, S. J., Waychunas, G. A., & Logan, T. J. (1998). Experimental and theoretical vibrational spectroscopic evaluation of arsenate coordination in aqueous solutions, solids, and at mineral-water interfaces. Geochimica et Cosmochimica Acta, 62, 32853300.CrossRefGoogle Scholar
Navas-Acien, A., Silbergeld, E. K., Pastor-Barriuso, R., & Guallar, E. (2010). Arsenic exposure and prevalence of type 2 Diabetes in US adults. Journal of the American Medical Association, 300, 814822.CrossRefGoogle Scholar
Newville, M. (2001). IFEFFIT: interactive XAFS analysis and FEFF fitting. Journal of Synchrotron Radiation, 8, 322324.CrossRefGoogle ScholarPubMed
Niu, L. (2011). Arsenic distribution and speciation in antigorite-rich rocks from Vermont, USA. M.Sc. thesis, Univ. Ottawa, Ontario, Canada, 111 pp.Google Scholar
Opiso, E. M., Sato, T., Morimoto, K., Asai, A., Anraku, S., Numako, C., & Yoneda, T. (2010). Incorporation of arsenic during the formation of Mg-bearing minerals at alkaline condition. Minerals Engineering, 23, 230237.CrossRefGoogle Scholar
Pascua, C., Charnock, J., Polya, D. A., Sato, T., Yokoyama, S., & Minato, M. (2005). Arsenic bearing smectite from the geothermal environment. Mineralogical Magazine, 69, 897906.CrossRefGoogle Scholar
Petit, S., & Decarreau, A. (1990). Hydrothermal (200°C) synthesis and crystal chemistry of iron-rich kaolinites. Clay Minerals, 25, 181196.CrossRefGoogle Scholar
Petit, S., Baron, F., & Decarreau, A. (2017). Synthesis of nontronite and other Fe-rich smectites: a critical review. Clay Minerals, 52, 469483.CrossRefGoogle Scholar
Ravel, B., & Newville, M. (2005). ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. Journal of Synchrotron Radiation, 12, 537541.CrossRefGoogle ScholarPubMed
Rehr, J. J., Kas, J. J., Vila, F. D., Prange, M. P., & Jorissen, K. (2010). Parameter-free calculations of X-ray spectra with FEFF9. Physical Chemistry and Chemical Physics, 12, 55035513.CrossRefGoogle ScholarPubMed
Rozalén, M., Ramos, M. E., Fiore, S., Gervilla, F., & Huertas, F. J. (2014). Effect of oxalate and pH on chrysotile dissolution at 25 °C: An experimental study. American Mineralogist, 99, 589600.CrossRefGoogle Scholar
Ryan, P. C., Kim, J., Wall, A. J., Moen, J. C., Corenthal, L. G., Chow, D. R., Sullivan, C. M., & Bright, K. S. (2011). Ultramafic-derived arsenic in a fractured bedrock aquifer. Applied Geochemistry, 26, 444457.CrossRefGoogle Scholar
Serna, C. J., White, J. L., & Velde, B. D. (1979). The effect of aluminium on the infra-red spectra of 7 Å trioctahedral minerals. Mineralogical Magazine, 43, 141148.CrossRefGoogle Scholar
Shannon, R. D. (1976). Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallografica, A32, 751767.CrossRefGoogle Scholar
Shi, X., Ayotte, J. D., Onda, A., Miller, S., Rees, J., Gilbert-Diamond, D., Onega, T., Gui, J., Karagas, M., & Moeschler, J. (2015). Geospatial association between adverse birth outcomes and arsenic in groundwater in New Hampshire, USA. Environmental Geochemistry and Health, 37, 333351.CrossRefGoogle ScholarPubMed
Shock, E. L., Sassani, D. C., Willis, M., & Sverjensky, D. A. (1997). Inorganic species in geologic fluids: Correlations among standard molal thermodynamic properties of aqueous ions and hydroxide complexes. Geochimica et Cosmochimica Acta, 61, 907950.CrossRefGoogle ScholarPubMed
Smedley, P. L., & Kinniburgh, D. G. (2002). A review of the source, behavior and distribution of As in natural waters. Applied Geochemistry, 17, 517568.CrossRefGoogle Scholar
Suquet, H. (1987). Effects of dry grinding and leaching on the crystal structure of chrysotile. Clays and Clay Minerals, 37, 439445.CrossRefGoogle Scholar
Takahashi, Y., Ohtaku, N., Mitsunobu, S., Yuita, K., & Nomura, M. (2003). Determination of the As(III)/As(V) Ratio in Soil by X-ray Absorption Near-edge Structure (XANES) and Its Application to the Arsenic Distribution between Soil and Water. Analytical Sciences, 19, 891896.CrossRefGoogle Scholar
Viti, C., & Mellini, M. (1997). Contrasting chemical compositions in associated lizardite and chrysotile in veins from Elba, Italy. European Journal of Mineralogy, 9, 585596.CrossRefGoogle Scholar
Webster, J. G., & Nordstrom, D. K. (2003). Geothermal arsenic: the source, transport and fate of arsenic in geothermal systems. In Welch, A. H. & Stollenwerk, K. G. (Eds.), Arsenic in Ground Water: Geochemistry and Occurrence (pp. 101126). New York: Kluwer Academic Publishers.CrossRefGoogle Scholar
Wicks, F. F., & O'Hanley, D. S. (1988). Serpentine minerals: structures and properties. In Bailey, S. W. (Ed.), Hydrous Phyllosilicates, Reviews in Mineralogy 19 (pp. 91167). Washington, D.C.: Mineralogical Society of America.CrossRefGoogle Scholar
Yariv, S., & Heller-Kallai, L. (1975). The relationship between the IR spectra of serpentines and their structures. Clays and Clay Minerals, 23, 145152.CrossRefGoogle Scholar
Supplementary material: File

Ryan et al. supplementary material
Download undefined(File)
File 1.2 MB
Supplementary material: File

Fig S1.

Ryan et al. supplementary material
Download Fig S1.(File)
File 62.6 KB