Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-23T12:33:26.133Z Has data issue: false hasContentIssue false

Tripuhyite and schafarzikite: two of the ultimate sinks for antimony in the natural environment

Published online by Cambridge University Press:  05 July 2018

P. Leverett
Affiliation:
School of Science and Health, University of Western Sydney, Locked Bag 1797, Penrith, New South Wales 2751, Australia
J. K. Reynolds
Affiliation:
School of Science and Health, University of Western Sydney, Locked Bag 1797, Penrith, New South Wales 2751, Australia
A. J. Roper
Affiliation:
School of Science and Health, University of Western Sydney, Locked Bag 1797, Penrith, New South Wales 2751, Australia
P. A. Williams*
Affiliation:
School of Science and Health, University of Western Sydney, Locked Bag 1797, Penrith, New South Wales 2751, Australia

Abstract

Studies of the stability of the oxides schafarzikite, FeSb2O4, and tripuhyite, FeSbO4, have been undertaken to clarify the roles these secondary minerals may have in determining the dispersion of antimony in oxidizing environments. Solubilities were determined at 298.15 K in aqueous HNO3, and these data were used to calculate values of ΔGfϴ at the same temperature. The derived Δ Gfϴ (s, 298.15 K) values for FeSb2O4 and FeSbO4 are – 959.4±4.3 and – 836.8±2.2 kJ mol–1, respectively. These results have been compared with electrochemically derived data, extrapolated from 771–981 K. The present study shows conclusively that although the mobility of Sb above the water table is limited by simple Sb(III) and Sb(V) oxides and stibiconite-group minerals, depending upon the prevailing redox potential and pH, tripuhyite is an important ultimate sink for Sb in the supergene environment. It is highly insoluble even in strongly acidic conditions and its anomalous stability at ambient temperatures causes the common mineral goethite, FeOOH, to react to form tripuhyite at activities of Sb(OH)5(aq) as low as 10–11. The comparatively limited numbers of reported occurrences of tripuhyite in the supergene zone are almost certainly due to the fact that its physical properties, especially colour and habit, are remarkably similar to those of goethite. In contrast, the small number of reported occurrences of schafarzikite can be related to its decomposition to tripuhyite as redox potentials rise at the top of the supergene zone and the fact that it decomposes to sénarmontite, Sb2O3, in acidic conditions, releasing Fe2+ ions into solution. In general, the findings confirm the immobility of Sb in near-surface conditions. Geochemical settings favouring the formation of the above minerals have been assessed using the results of the present study and data from the literature.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2012

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

Accornero, M., Marini, L. and Lelli, M. (2008) The dissociation constant of antimonic acid at 10-40°C. Journal of Solution Chemistry, 37, 785800.CrossRefGoogle Scholar
Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1997) Handbook of Mineralogy. III. Halides, Hydroxides, Oxides. Mineral Data Publishing, Tucson.Google Scholar
Atencio, D., Andrade, M., Christy, A.G., Giere, R. and Kartashov, P.M. (2010) the pyrochlore supergroup of minerals: nomenclature. The Canadian Mineralogist, 48, 673698.CrossRefGoogle Scholar
Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Wiley Interscience, New York.Google Scholar
Barin, I. (1989) Thermochemical Data of Pure Substances. Part II VCH Publishers, Weinheim.Google Scholar
Basso, R., Cabella, R, Lucchetti, G., Marescotti, P. and Martinelli, A. (2003) Structural studies on synthetic and natural Fe-Sb-oxides of MO2 type. Neues Jahrbuch fur Mineralogie, Monatshefte, 2003, 407420.CrossRefGoogle Scholar
Berlepsch, P., Armbruster, T., Brugger, J. and Graeser, S. (2003) Tripuhyite, FeSbO4, revisited. Mineralogical Magazine, 67, 3146.CrossRefGoogle Scholar
Brookins, D.G. (1988) Eh-pH Diagrams for Geochemistry. Springer, Berlin.CrossRefGoogle Scholar
Cabella, R, Basso, R., Lucchetti, G, Marescotti, P., Martinelli, A. and Nayak, V.K. (2003) Squawcreekite—rutile solid solution from the Kajlidongri mine (India). European Journal of Mineralogy, 15, 427433.Google Scholar
Chater, R., Gavarri, J.R. and Hewat, A. (1985) Structures isomorphes MeX2O4—evolution stuctur-ale entre 2K et 300K l'antimonite FeSb2O4: elasticite et ordre magnetique anisotropes. Journal of Solid State Chemistry, 60, 7886.CrossRefGoogle Scholar
Coppola, V., Boni, M., Gilg, A. and Strzelska-Smakowska, B. (2009) Nonsulfide zinc deposits in the Silesia-Cracow district, Southern Poland. Mineralium Deposita, 44, 559580.CrossRefGoogle Scholar
Diemar, G.A., Filella, M., Leverett, P. and Williams, P.A. (2009) Dispersion of antimony from oxidizing ore deposits. Pure and Applied Chemistry, 81, 15471553.CrossRefGoogle Scholar
Filella, M. and May, P.M. (2003) Computer simulation of the low-molecular-weight inorganic species distribution of antimony(III) and antimony(V) in natural waters. Geochimica et Cosmochimica Acta, 67, 40134031.CrossRefGoogle Scholar
Filella, M., Belzile, N. and Chen, Y.-W. (2002a) Antimony in the environment: a review focused on natural waters I. Occurrence. Earth-Science Reviews, 57, 125176.CrossRefGoogle Scholar
Filella, M., Belzile, N. and Chen, Y.-W. (2002b) Antimony in the environment: a review focused on natural waters II. Relevant solution chemistry. Earth-Science Reviews, 59, 265285 CrossRefGoogle Scholar
Filella, M., Williams, P.A. and Belzile, N. (2009) Antimony in the environment: knowns and un-knowns. Environmental Chemistry, 6, 95105.CrossRefGoogle Scholar
Fischer, R. and Pertlik, F. (1975) Verfeinerung der kristallstruktur des schafarzikites, FeSb2O4 . Tschermaks Mineralogische und Petrographische Mitteilungen, 22, 236241.CrossRefGoogle Scholar
Foord, E.E., Hlava, P.F., Fitzpatrick, J.J., Erd, R.C. and Hinton, RW. (1991) Maxwellite and squawcreekite, two new minerals from the Black Range tin district, Catron County, New Mexico U.S.A. Neues Jahrbuch fur Mineralogie, Monatshefte, 1991, 363384.Google Scholar
Gayer, K.H. and Garrett, A.B. (1952) The equilibria of antimonous oxide (rhombic)in dilute solutions of hydrochloric acid and sodium hydroxide at 25°C. Journal of the American Chemical Society, 74, 23532354.CrossRefGoogle Scholar
Grube, G. and Schweigardt, F. (1923) Uber das elektrochemische Verhalten von Wismut und Antimon in alkalischer Losung. Zeitschrift fur Elektrochemie und angewandte physikalische Chemie, 29, 257264.Google Scholar
Hussak, E. and Prior, GT. (1897) On tripuhyite, a new antimonite of iron, from Tripuhy, Brazil. Mineralogical Magazine, 11, 302303.CrossRefGoogle Scholar
Klimko, T., Lalinska, B. and Chovan, M. (2010) Secondary Sb mineral phases from abandoned Sb deposit Dubrava (Slovakia). Abstracts of the 20th General Meeting of the International Mineralogical Association, abstract EM60G_P013_S 1.Google Scholar
Konopik, N. and Zwiauer, J. (1952) Uber Antimontetroxid. Monatshefte fur Chemie, 83, 189196.CrossRefGoogle Scholar
Krupka, K.M. and Serne, R.J. (2002) Geochemical factors affecting the behaviour of antimony, cobalt, europium, technetium, and uranium in vadose sediments. United States Department of Energy Pacific Northwest Laboratory Report, PNNL-14126.Google Scholar
Langford, J.I. (1973) Least-squares refinement of cell dimensions from powder data by Cohen's method. Journal of Applied Crystallography, 6, 190196.CrossRefGoogle Scholar
Lefebvre, J. and Maria, H. (1963) Etude des equilibres dans les solutions recentes de polyantimoniates. Comptes rendus de VAcademie des sciences de Paris, 256, 31213124.Google Scholar
Majzlan, J., Lalinska, B., Chovan, M., BlaP, U, Brecht, B., Gottlicher, J., Steininger, R., Hug, K., Ziegler, S. and Gescher, J. (2011) A mineralogical, geochemical, and microbiological assessment of the antimony-and arsenic-rich neutral mine drainage tailings near Pezinok, Slovakia. American Mineralogist, 96, 113.CrossRefGoogle Scholar
Martinelli, A., Ferretti, M., Buscaglia, V., Cabella, R. and Lucchetti, G (2002) Formation and decomposition of the rutile-type compound FeSbO4: a TG-DTA study. Journal of Thermal Analysis and Calorimetry, 70, 123127.CrossRefGoogle Scholar
Martinelli, A., Ferretti, M., Basso, R, Cabella, R, Lucchetti, G, Marescotti, P. and Buscaglia, V. (2004) Solid state miscibility in the pseudo-binary TiO2-(FeSb)O4 system at 1373 K. Zeitschrift fur Kristallographie, 219, 487493.Google Scholar
Martinelli, A., Ferretti, M., Basso, R, Cabella, R, Lucchetti, G. (2006) Solid state solubility between SnO2 and (FeSb)O4 at high temperature. Zeitschrift fur Kristallographie, 221, 716721.Google Scholar
Mason, B. and Vitaliano, C.J. (1953) The mineralogy of the antimony oxides and antimonates. Mineralogical Magazine, 30, 100112.CrossRefGoogle Scholar
May, P.M. and Murray, K. (2000) Database of chemical reactions designed to achieve thermodynamic consistency automatically. Journal of Chemical Engineering Data, 46, 10351040.CrossRefGoogle Scholar
Mishra, S.K. and Gupta, Y.K. (1968) Spectrophotometric study of hydrolytic equilibria of Sb(III) in aqueous perchloric acid solution. Indian Journal of Chemistry, 6, 757758.Google Scholar
Mitsunobu, S., Takahashi, Y. and Sakata, M. (2010) Antimony (V) incorporation into iron oxyhydrox-ides. Environmental Science and Technology, 44, 37123718.CrossRefGoogle Scholar
Mitsunobu, S., Takahashi, Y., Utsunomiya, S., Marcus, M.A., Terada, Y., Iwamura, T. and Sakata, M. (2011) Identification and characterization of nanosized tripuhyite in soil near Sb mine tailings. American Mineralogist, 96, 11711181.CrossRefGoogle Scholar
Nordstrom, D.K., Alpers, C.N., Ptacek, C.J. and Blowes, D.W. (2000) Negative pH and extremely acidic mine waters from Iron Mountain, California. Environmental Science and Technology, 34, 254258.CrossRefGoogle Scholar
Orlandi, P. and Dini, A. (2004) Die Mineralien der Buca della Vena-Mine, Apuaner Berge, Toskana (Italien). Lapis, 29 (1), 1124.Google Scholar
Pankajavalli, R. and Sreedharan, O.M. (1987) Thermodynamic stability of Sb2O4 by a solid oxide electrolyte e.m.f. method. Journal of Materials Science, 22, 177180 CrossRefGoogle Scholar
Parker, and Khodakovskii, (1995) Thermodynamic properties of the aqueous ions (2+ and 3+) of iron and the key compounds of iron. Journal of Physical and Chemical Reference Data, 24, 16991745.CrossRefGoogle Scholar
Past, V. (1985) Antimony. Pp. 172179 in: Standard Potentials in Aqueous Solution (Bard, A.J., Parsons, R. and Jordan, J., editors) Marcel Dekker, New York.Google Scholar
Philippo, S. and Hanson, A. (2007) La mineralisation en antimoine de Goesdorf. Ferrantia, 49, 111146.Google Scholar
Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of minerals and related substances at 298.15 K and 1 bar (105 Pascals) pressure and at higher temperatures. United States Geological Survey Bulletin, 2131.Google Scholar
Robinson, G.W. and King, V.T (1991) What's new in minerals. Mineralogical Record, 22, 381393.Google Scholar
Robinson, G.W. and Normand, C. (1996) The Lac Nicolet antimony mine, South Ham, Quebec. Mineralogical Record, 27, 121134.Google Scholar
Schulze, H. (1883) Antimontrisulfid in wassriger Losung. Journal fur praktische Chemie, 27, 320332.CrossRefGoogle Scholar
Schuhmann, R. (1924) The free energy of antimony trioxide and the reduction potential of antimony. Journal of the American Chemical Society, 46, 5258.CrossRefGoogle Scholar
Sejkora, S., Ozdin, D., Vitalos, J., Tucek, P. and Dud'a R (2004) Schafarzikit von Pernek, Revier Pezinok (Slowakei). Lapis, 29, 2736.Google Scholar
Sejkora, J., Ozdin, D., Vitalos, J., Tucek, P., Cejka, I and Dud'a, R (2007) Schafarzikite from the type locality Pernek (Male Kaparty mountains, Slovak) revisited. European Journal of Mineralogy, 19, 419427.CrossRefGoogle Scholar
Stefansson, A. and Seward, T.M. (2008) A spectro-photometric study of iron(III) hydrolysis in aqueous solutions to 200°C. Chemical Geology, 249, 227235.CrossRefGoogle Scholar
Swaminathan, K. and Sreedharan, O.M. (2003) High temperature stabilities of interoxides in the system Fe—Sb—O and their comparison with the interoxides in other M—Sb—O (M = Cr, Ni or Co) systems. Journal of Alloys and Compounds, 358, 4855.CrossRefGoogle Scholar
Vink, B.W. (1996) Stability relations of antimony and arsenic compounds in the light of revised and extended Eh—pH diagrams. Chemical Geology, 130, 2130.CrossRefGoogle Scholar
Wagman, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halow, I., Bailey, S.M., Churney, K.I. and Nuttall, R.I., (1982) The NBS tables of chemical thermodynamic properties: selected values for inorganic and C1 and C2 organic substances in SI units. Journal of Physical and Chemical Reference Data, 11, Supplement Number 2.Google Scholar
Williams, PA. (1990) Oxide Zone Geochemistry. Ellis Horwood, Chichester, UK.Google Scholar
Wilson, N.J., Craw, D. and Hunter, K (2004) Antimony distribution and environmental mobility at an historic antimony smelter site, New Zealand. Environmental Pollution, 129, 257266.CrossRefGoogle ScholarPubMed
Zakaznova-Herzog, V.P. and Seward, T.M. (2006) Antimonous acid protonation/deprotonation equili-bria in hydrothermal solutions to 300°C. Geochimica et Cosmochimica Acta, 70, 22982310.CrossRefGoogle Scholar
Zotov, A.V., Shikina, N.D. and Akinfiev, N.N. (2003) Thermodynamic properties of the Sb(III) hydroxide complex Sb(OH)3(aq) at hydrothermal conditions. Geochimica et Cosmochimica Acta, 67, 18211836.CrossRefGoogle Scholar