Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-25T04:43:38.381Z Has data issue: false hasContentIssue false

Geschieberite, K2(UO2)(SO4)2(H2O)2, a new uranyl sulfate mineral from Jáchymov

Published online by Cambridge University Press:  02 January 2018

J. Plášil*
Affiliation:
Institute of Physics ASCR, v.v.i., Na Slovance 2, CZ–182 21, Prague 8, Czech Republic
J. Hloušek
Affiliation:
U Roháčových kasáren 24, CZ–100 00, Prague 10, Czech Republic
A. V. Kasatkin
Affiliation:
Fersmann Mineralogical Museum of the Russian Academy of Sciences, Leninsky Prospekt 18-2, 119071 Moscow, Russia
R. Škoda
Affiliation:
Department of Geological Sciences, Faculty of Science, Masaryk University, Kotlářská 2, 611 37 Brno, Czech Republic
M. Novák
Affiliation:
Department of Geological Sciences, Faculty of Science, Masaryk University, Kotlářská 2, 611 37 Brno, Czech Republic
J. Čejka
Affiliation:
Department of Mineralogy and Petrology, National Museum, Cirkusová 1740, CZ–193 00, Prague 9, Czech Republic
*
* E-mail: plasil@fzu.cz

Abstract

The new mineral geschieberite (IMA2014-006), K2(UO2)(SO4)2(H2O)2, was found in the Svornost mine, Jáchymov, Czech Republic, where it occurs as a secondary alteration phase after uraninite in association with adolfpateraite and gypsum. Geschieberite forms crystalline aggregates of bright green colour (when thick) composed of multiply intergrown prismatic crystals elongated on [001] typically reaching 0.1–0.2 mm across; observable forms are {010} and {001}. Crystals are translucent to transparent with a vitreous lustre. The mineral is brittle, with perfect cleavage on {100} and an uneven fracture. It has a greenish-white streak and a probable Mohs hardness of ∼2. The mineral is slightly soluble in cold H2O. The calculated density is 3.259 g cm–3. The mineral exhibits strong yellowish-green fluorescence under both shortwave and longwave UV radiation. Optically, geschieberite is biaxial (–), with β = 1.596(2) and γ = 1.634(4) (measured at 590 nm), with X = a. Electron-microprobe analyses provided Na2O 0.23, K2O 14.29, MgO 2.05, CaO 0.06, UO3 49.51, SO3 27.74, H2O 6.36 (structure), total 100.24 wt.%, yielding the empirical formula (K1.72Mg0.29Na0.04Ca0.01)Σ2.06(U0.98O2)(S0.98O4)2(H2O)2 based on 12 O atoms per formula unit. The Raman spectrum is dominated by the symmetric stretching vibrations of UO22+, SO42– and weaker O–H stretching vibrations. Geschieberite is orthorhombic, Pna21, with a = 13.7778(3), b = 7.2709(4), c = 11.5488(2) Å, V = 1156.92(7) Å3, Z = 4. The eight strongest powder X-ray diffraction lines are [dobs in Å (hkl) Irel]: 6.882 (200) 100, 5.622 (111) 53, 4.589 (211) 12, 4.428 (202) 16, 3.681 (311) 18, 3.403 (013) 12, 3.304 (401,1̄13) 15 and 3.006 (122) 17. The structure, refined to R = 0.028 for 1882 I > 3σ(I) reflections, contains [(UO2)(SO4)2(H2O)]2– sheets that are based on the protasite anion topology. Sheets are stacked perpendicular to a. Potassium atoms and one H2O molecule are located between these sheets, providing an interlayer linkage. The remaining H2O molecule is localized within the structural unit, at the free vertex of the uranyl pentagonal bipyramid; this vertex does not link to sulfate tetrahedra. The mineral is named for one of the most important ore veins in Jáchymov – the Geschieber vein.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2015

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

deceased, April 27, 2014

References

Alekseev, E.V., Sulemanov, E.V., Chuprunov, E.V., Marychev, M.O., Ivanov, V.A. and Fukin, G.K. (2006) Crystal structure and nonlinear optical properties of the K2UO2(SO4)2.2H2O compound at 293K. Crystallography Reports 51, 29-33.CrossRefGoogle Scholar
Bartlett, J.R. and Cooney, R.P. (1989) On the determination of uranium-oxygen bond lengths in dioxouranium(VI) compounds by Raman spectroscopy. Journal of Molecular Structure 193, 295-300.CrossRefGoogle Scholar
Brown, I.D. (1981) The bond-valence method: an empirical approach to chemical structure and bonding. Pp. 1-30. in: Structure and Bonding in Crystals II (M. O’Keeffe and A. Navrotsky, editors). Academic Press, New York.Google Scholar
Brown, I.D. (2002) The Chemical Bond in Inorganic Chemistry: The Bond Valence Model. Oxford University Press, Oxford, UK.Google Scholar
Brown, I.D. and Altermatt, D. (1985) Bond-valence parameters obtained from a systematic analysis of the inorganic crystal structure database. Acta Crystallographica, B41, 244-247. with updated parameters from http://www.ccp14.ac.uk/ccp/web-mirrors/i_d_brown/. CrossRefGoogle Scholar
Burns, P.C. (2001) A new uranyl sulfate chain in the structure of uranopilite. The Canadian Mineralogist 39, 1139-1146.CrossRefGoogle Scholar
Burns, P.C. (2005) U6+ minerals and inorganic compounds: insights into an expanded structural hierarchy of crystal structures. The Canadian Mineralogist 43, 1839-1894.CrossRefGoogle Scholar
Burns, P.C., Ewing, R.C. and Hawthorne, F.C. (1997) The crystal chemistry of hexavalent uranium: polyhedron geometries, bond-valence parameters, and polymerization of polyhedra. The Canadian Mineralogist 35, 1551-1570.Google Scholar
Čejka, J. (1999) Infrared and thermal analysis of the uranyl minerals. Pp. 521-622. in: Uranium: Mineralogy, Geochemistry, and the Environment (P.C. Burns and R. Finch, editors). Reviews in Mineralogy, 38. Mineralogical Society of America, Washington, DC.CrossRefGoogle Scholar
Čejka, J. (2004) Vibrational spectroscopy of uranyl mineral – infrared and Raman spectra of uranyl minerals. I. Uranyl. Bulletin mineralogicko-petrologicke ´ho oddeˇlení Národního muzea (Praha) 12, 44-51.Google Scholar
Čejka, J. (2007) Vibrational spectroscopy of uranyl minerals – infrared and Raman spectra of uranyl minerals. III. Uranyl sulfates. Bulletin mineralogicko/ petrologického oddeˇlení Národního muzea (Praha), 14–15. 40–46.Google Scholar
Edwards, K.J., Bond, P.L., Druschel, G.K., McGuire, M.M., Hamers, R.J. and Banfield, J.F. (2000) Geochemical and biological aspects of sulfide mineral dissolution: lessons from Iron Mountain, California. Chemical Geology 169, 383-397.CrossRefGoogle Scholar
Evangelou, V.P. and Zhang, Y.L. (1995) A review – pyrite oxidation mechanisms and acid mine drainage prevention. Critical Reviews in Environmental Sciences and Technology 25, 141-199.CrossRefGoogle Scholar
Frondel, C., Ito, J., Honea, R.M. and Weeks, A.M. (1976) Mineralogy of the zippeite-group. The Canadian Mineralogist 14, 429-436.Google Scholar
Jambor, J.L., Nordstrom, D.K. and Alpers, C.N. (2000) Metal–sulfate salts from sulfide mineral oxidation. Pp. 303-350. in: Sulfate Minerals: Crystallography, Geochemistry, and Environmental (Alpers, C.N., Jambor, J.L. and Nordstrom, D.K., editors). Reviews in Mineralogy & Geochemistry, 40. Mineralogical Society of America and the Geochemical Society, Washington, DC.Google Scholar
Kampf, A.R., Plášil, J., Kasatkin, A.V. and Marty, J. (2014) Bel akovskiite , Na7(UO2 ) (SO4 ) 4 (SO3OH)(H2O)3, a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. Mineralogical Magazine 78, 639-649.CrossRefGoogle Scholar
Kraus, W. and Nolze, G. (1996) POWDER CELL – a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. Journal of Applied Crystallography 29, 301-303.CrossRefGoogle Scholar
Krivovichev, S.V. (2010) Actinyl compounds with hexavalent elements (S, ,Cr, Se, Mo): structural diversity, nanoscale chemistry, and cellular automata modelling. European Journal of Inorganic Chemistry 2010, 2594-2603.CrossRefGoogle Scholar
Laugier, J. and Bochu, B. (2004) LMPG Suite of Programs for the Interpretation of X-ray Experiments. ENSP/Laboratoire des Matériaux et du Génie Physique, BP 46. 38042 Saint Martin d’Hères, France. URL: http://www.ccp14.ac.uk/tutorial/lmgp/ (Accessed: 16 January 2014).Google Scholar
Libowitzky, E. (1999) Correlation of O-H stretching frequencies and O-H_O hydrogen bond lengths in minerals. Monatshefte für Chemie 130, 1047-1059.CrossRefGoogle Scholar
Ling, J., Sigmon, G.E., Ward, M., Roback, N. and Burns, P.C. (2010) Syntheses, structures, and IR spectroscopic characterization of new uranyl sulfate/ selenate 1D-chain, 2D-sheet and 3D framework. Zeitschrift für Kristallographie 225, 230-239.Google Scholar
Mandarino, J.A. (1981) The Gladstone-Dale relationship: Part IV. The compatibility concept and its application. The Canadian Mineralogist 19, 441-450.Google Scholar
Nakamoto, K. (1986) Infrared and Raman Spectra of Inorganic and Coordination Compounds. John Wiley and Sons, New York.Google Scholar
Niinistö, L., Toivonen, J. and Valkonen, J. (1978) Uranyl(VI) compounds. I. The crystal structure of ammonium sulfate dihydrate, (NH4)UO2 (SO4)2.2H2O, Acta Chemica Scandinavica, A32, 647-651.Google Scholar
Niinistö, L., Toivonen, J. and Valkonen, J. (1979) Uranyl(VI) compounds. II. The crystal structure of potassium uranyl sulphate dehydrate K2UO2 (SO4)2·2H2O. Acta Chemica Scandinavica, A33, 621-624.CrossRefGoogle Scholar
Oszlányi, G. and Süto, A. (2004) Ab-initio structure solution by charge flipping. Acta Crystallographica, A60, 134-141.CrossRefGoogle Scholar
Oszlányi, G. and Süto, A. (2008) The charge flipping algorithm. Acta Crystallographica, A64, 123-134.CrossRefGoogle Scholar
Palatinus, L. (2013) The charge-flipping algorithm in crystallography. Acta Crystallographica, B69, 1-16.Google Scholar
Palatinus, L. and Chapuis, G. (2007) Superflip – a computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. Journal of Applied Crystallography 40, 451-456.CrossRefGoogle Scholar
Palatinus, L. and van der Lee, A. (2008) Symmetry determination following structure solution in P1. Journal of Applied Crystallography 41, 975-984.CrossRefGoogle Scholar
Pekov, I.V., Krivovichev, S.V., Yapaskurt, V.O., Chukanov, N.V. and Belakovskyi, D.I. (2014) Beshtauite, (NH4)2(UO2)(SO4)2·2H2O, a new mineral from Mount Beshtau, Northern Caucasus, Russia. American Mineralogist 99, 1783-1787.CrossRefGoogle Scholar
PetříČek, V., Dušek, M. and Palatinus, L. (2006) Jana2006. The Crystallographic Computing System. Institute of Physics, Prague, Czech Republic.Google Scholar
PetříČek, V., Dušek, M. and Palatinus, L. (2014) Crystallographic computing system JANA2006: general features. Zeitschrift für Kristallographie 229, 345-352.Google Scholar
Plášil, J. (2014) Oxidation–hydration weathering of uraninite: the current state-of-knowledge. Journal of Geosciences 59, 99-114.CrossRefGoogle Scholar
Plášil, J., Buixaderas, E., Čejka, J., JehliČka, J. and Novák, M. (2010) Raman spectroscopic study of the uranyl sulphate mineral zippeite: low wavenumber and U–O stretching regions. Analytical and Bioanalytical Chemistry 397, 2703-2715.CrossRefGoogle ScholarPubMed
Plášil, J., Dušek, M., Novák, M., Čejka, J., Císařová, I. and Š, koda, R. (2011) Sejkoraite-(Y), a new member of the zippeite group containing trivalent cations from Jáchymov (St. Joachimsthal), Czech Republic: description and crystal structure refinement. American Mineralogist 96, 983-991.CrossRefGoogle Scholar
Plášil, J., Hloušek, J., Veselovský , F., Fejfarová, K., Dušek, M., Š koda, R., Novák, M., Čejka, J., Sejkora, J. and Ondruš , P. (2012) Adolfpateraite, K(UO2)(SO4)(OH)(H2O), a new uranyl sulfate mineral from Jáchymov, Czech Republic. American Mineralogist 97, 447-454.CrossRefGoogle Scholar
Plášil, J., Kasatkin, A.V., Š koda, R., Novák, M., Kallistová, A., Dušek, M., Skála, R., Fejfarová, K., Čejka, J., Meisser, N., Goethals, H., MachoviČ, V. and Lapčák, L. (2013a) Leydetite, Fe(UO2) (SO4)2(H2O)11, a new uranyl sulfate mineral from Mas d’Alary, Lodève, France. Mineralogical Magazine 77, 429-441.CrossRefGoogle Scholar
Plášil, J., Kampf, A.R., Kasatkin, A.V., Marty, J., Š koda, R., Silva, S. and Čejka, J. (2013b) Meisserite, Na5(UO2)(SO4)3(SO3OH)(H2O), a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. Mineralogical Magazine 77, 2975-2988.CrossRefGoogle Scholar
Plášil, J., Veselovský , F., Hloušek, J., Š koda, R., Novák, M., Sejkora, J., Čejka, J., Škácha, P. and Kasatkin, A.V.(2014a) Mathesiusite, K5 (UO2)4(SO4)4 (VO5)(H2O)4, a new uranyl vanadate-sulfate from Jáchymov, Czech Republic. American Mineralogist 99, 625-632.CrossRefGoogle Scholar
Plášil, J., Kampf, A.R., Kasatkin, A.V. and Marty, J. (2014b) Bluelizardite, Na7(UO2)(SO4)4Cl(H2O)2, a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. Journal of Geosciences 59, 145-158.CrossRefGoogle Scholar
Plášil, J., Hloušek, J., Kasatkin, A.V., Š koda, R., Novák, M. and Čejka, J. (2014c) Geschieberite, IMA 2014- 006. CNMNC Newsletter No. 20, June 2014, page 555; Mineralogical Magazine 78, 549-558.Google Scholar
Plášil, J., Hloušek, J., Kasatkin, A.V., Novák, M. and Čejka, J. and Lapčák, L. (2015) Svornostite, K2Mg[(UO2)(SO4)2]2·8H2O, a new uranyl sulfate mineral from Jáchymov, Czech Republic. Journal of Geosciences 60, 113-121.CrossRefGoogle Scholar
Pouchou, J.L. and Pichoir, F. (1985) ‘PAP’ (j rZ) procedure for improved quantitative microanalysis. Pp. 104-106. in: Microbeam Analysis (J.T. Armstrong, editor). San Francisco Press, San Francisco, California, USA.Google Scholar
Supplementary material: File

Plášil et al. supplementary material

CIF

Download Plášil et al. supplementary material(File)
File 156.4 KB