Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-19T18:46:57.263Z Has data issue: false hasContentIssue false

The crystal structures of the mixed-valence tellurium oxysalts tlapallite, (Ca,Pb)3CaCu6[Te4+3Te6+O12]2(Te4+O3)2(SO4)2·3H2O, and carlfriesite, CaTe4+2Te6+O8

Published online by Cambridge University Press:  12 February 2019

Owen P. Missen*
Affiliation:
Geosciences, Museums Victoria, GPO Box 666, Melbourne 3001, Victoria, Australia School of Earth, Atmosphere and Environment, 9 Rainforest Walk, Monash University, Clayton 3800, Victoria, Australia
Anthony R. Kampf
Affiliation:
Mineral Sciences Department, Natural History Museum of Los Angeles County, 900 Exposition Boulevard, Los Angeles, CA 90007, USA
Stuart J. Mills
Affiliation:
Geosciences, Museums Victoria, GPO Box 666, Melbourne 3001, Victoria, Australia
Robert M. Housley
Affiliation:
Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA
John Spratt
Affiliation:
Department of Core Research Laboratories, Natural History Museum, Cromwell Road, London SW7 5BD, UK
Mark D. Welch
Affiliation:
Department of Earth Sciences, Natural History Museum, Cromwell Road, London SW7 5BD, UK
Mark F. Coolbaugh
Affiliation:
Renaissance Gold Inc., 4750 Longley Lane, Suite 106, Reno, NV 89502, USA
Joe Marty
Affiliation:
5199 East Silver Oak Road, Salt Lake City, UT 84108, USA
Marek Chorazewicz
Affiliation:
124 Pineplank Lane, Simi Valley, CA 93065, USA
Cristiano Ferraris
Affiliation:
Laboratoire de Physique des Milieux Condensés (LPMC), CNRS-UMR 7590, Muséum National d'Histoire Naturelle (MNHN), 61 rue Buffon, 75005 Paris, France
*
*Author for correspondence: Owen P. Missen, Email: omissen@museum.vic.gov.au

Abstract

The crystal structure of tlapallite has been determined using single-crystal X-ray diffraction and supported by electron probe micro-analysis, powder diffraction and Raman spectroscopy. Tlapallite is trigonal, space group P321, with a = 9.1219(17) Å, c = 11.9320(9) Å and V = 859.8(3) Å3, and was refined to R1 = 0.0296 for 786 reflections with I > 2σ(I). This study resulted from the discovery of well-crystallised tlapallite at the Wildcat prospect, Utah, USA. The chemical formula of tlapallite has been revised to (Ca,Pb)3CaCu6[Te4+3Te6+O12]2(Te4+O3)2(SO4)2·3H2O, or more simply (Ca,Pb)3CaCu6Te4+8Te6+2O30(SO4)2·3H2O, from H6(Ca,Pb)2(Cu,Zn)3(TeO3)4(TeO6)(SO4). The tlapallite structure consists of layers containing distorted Cu2+O6 octahedra, Te6+O6 octahedra and Te4+O4 disphenoids (which together form the new mixed-valence phyllotellurate anion [Te4+3Te6+O12]12−), Te4+O3 trigonal pyramids and CaO8 polyhedra. SO4 tetrahedra, Ca(H2O)3O6 polyhedra and H2O groups fill the space between the layers. Tlapallite is only the second naturally occurring compound containing tellurium in both the 4+ and 6+ oxidation states with a known crystal structure, the other being carlfriesite, CaTe4+2Te6+O8. Carlfriesite is the predominant secondary tellurium mineral at the Wildcat prospect. We also present an updated structure for carlfriesite, which has been refined to R1 = 0.0230 for 874 reflections with I > 2σ(I). This updated structural refinement improves upon the one reported previously by refining all atoms anisotropically and presenting models of bond valence and Te4+ secondary bonding.

Type
Article
Copyright
Copyright © Mineralogical Society of Great Britain and Ireland 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

Associate Editor: Oleg I Siidra

References

Brown, I.D. (2002) The Chemical Bond in Inorganic Chemistry. Oxford University Press, UK.Google Scholar
Brugger, J., Liu, W., Etschmann, B., Mei, Y., Sherman, D.M. and Testemale, D. (2016). A review of the coordination chemistry of hydrothermal systems, or do coordination changes make ore deposits? Chemical Geology, 447, 219253.Google Scholar
Christy, A.G. and Mills, S.J. (2013) Effect of lone-pair stereoactivity on polyhedral volume and structural flexibility: Application to TeIVO6 octahedra. Acta Crystallographica, B69, 446456.Google Scholar
Christy, A.G., Mills, S.J. and Kampf, A.R. (2016 a) A review of the structural architecture of tellurium oxycompounds. Mineralogical Magazine, 80, 415545.Google Scholar
Christy, A.G., Mills, S.J., Kampf, A.R., Housley, R.M., Thorne, B. and Marty, J. (2016 b) The relationship between mineral composition, crystal structure and paragenetic sequence: The case of secondary Te mineralization at the Bird Nest Drift, Otto Mountain, California, USA. Mineralogical Magazine, 80, 291310.Google Scholar
Degen, T., Sadki, M., Bron, E., König, U. and Nénert, G. (2014) The HighScore suite. Powder Diffraction, 29, S13S18.Google Scholar
Eby, R. and Hawthorne, F. (1993) Structural relations in copper oxysalt minerals. I. Structural hierarchy. Acta Crystallographica, B49, 2856.Google Scholar
Effenberger, H., Zemann, J. and Mayer, H. (1978) Carlfriesite; crystal structure, revision of chemical formula, and synthesis. American Mineralogist, 63, 847852.Google Scholar
Gagné, O.C. and Hawthorne, F.C. (2015) Comprehensive derivation of bond-valence parameters for ion pairs involving oxygen. Acta Crystallographica, B71, 562578.Google Scholar
Gaines, R.V. (1968) Poughite, a new tellurite mineral from Mexico and Honduras. American Mineralogist, 53, 10751080.Google Scholar
Grundler, P.V., Brugger, J., Etschmann, B.E., Helm, L., Liu, W., Spry, P.G., Tian, Y., Testemale, D. and Pring, A. (2013) Speciation of aqueous tellurium (IV) in hydrothermal solutions and vapors, and the role of oxidized tellurium species in Te transport and gold deposition. Geochimica et Cosmochimica Acta, 120, 298325.Google Scholar
Higashi, T. (2001) ABSCOR. Rigaku Corporation, Tokyo, Japan.Google Scholar
Kampf, A.R., Housley, R.M., Rossman, G.R., Marty, J. and Chorazewicz, M. (2018) Bodieite, Bi3+2(Te4+O3)2(SO4), a new mineral from the Tintic district, Utah, and the Masonic district, California, USA. The Canadian Mineralogist, 56, 110.Google Scholar
Kampf, A.R. and Mills, S.J. (2011) The role of hydrogen in tellurites: Crystal structure refinements of juabite, poughite and rodalquilarite. Journal of Geosciences, 56, 235247.Google Scholar
Kampf, A.R., Mills, S.J., Housley, R.M., Rossman, G.R., Marty, J. and Thorne, B. (2013 a) Lead-tellurium oxysalts from Otto Mountain near Baker, California: X. Bairdite, Pb2Cu2+4Te6+2O10(OH)2(SO4)(H2O), a new mineral with thick HCP layers. American Mineralogist, 98, 13151321.Google Scholar
Kampf, A.R., Mills, S.J., Housley, R.M., Rossman, G.R., Marty, J. and Thorne, B. (2013 b) Lead-tellurium oxysalts from Otto Mountain near Baker, California: XI. Eckhardite, (Ca,Pb)Cu2+Te6+O5(H2O), a new mineral with HCP stair-step layers. American Mineralogist, 98, 16171623.Google Scholar
Kampf, A.R., Mills, S.J. and Rumsey, M.S. (2017) The discreditation of girdite. Mineralogical Magazine, 81, 11251128.Google 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, 301303.Google Scholar
Krivovichev, S.V. and Brown, I. (2001) Are the compressive effects of encapsulation an artifact of the bond valence parameters? Zeitschrift für Kristallographie, 216, 245247.Google Scholar
Lee, D.W. and Ok, K.M. (2014) New polymorphs of ternary sodium tellurium oxides: hydrothermal synthesis, structure determination, and characterization of β-Na2Te4O9 and Na2Te2O6·1.5H2O. Inorganic Chemistry, 53, 1064210648.Google Scholar
Libowitzky, E. (1999). Correlation of O-H stretching frequencies and O–H···O hydrogen bond lengths in minerals. Monatshefte für Chemie, 130, 10471059.Google Scholar
Lindqvist, O. and Moret, J. (1973) The crystal structure of ditellurium pentoxide, Te2O5. Acta Crystallographica, B29, 643650.Google Scholar
Marty, J., Kampf, A.R., Housley, R.M., Mills, S.J. and Weiß, S. (2010) Seltene neue Tellurmineralien aus Kalifornien, Utah, Arizona und New Mexico (USA). Lapis, 35, 4252.Google Scholar
McPhail, D. (1995) Thermodynamic properties of aqueous tellurium species between 25 and 350°C. Geochimica et Cosmochimica Acta, 59, 851866.Google Scholar
Mills, S.J. and Christy, A.G. (2013) Revised values of the bond-valence parameters for TeIV–O, TeVI–O and TeIV–Cl. Acta Crystallographica, B69, 145149.Google Scholar
Mills, S.J., Christy, A.G., Chen, E.C.-C. and Raudsepp, M. (2009 a) Revised values of the bond valence parameters for [6]Sb(V)–O and [3–11]Sb(III)–O. Zeitschrift für Kristallographie, 224, 423431.Google Scholar
Mills, S.J., Kolitsch, U., Miyawaki, R., Groat, L.A. and Poirier, G. (2009 b) Joëlbruggerite, Pb3Zn3(Sb5+,Te6+)As2O13(OH,O), the Sb5+ analog of dugganite, from the Black Pine mine, Montana. American Mineralogist, 94, 10121017.Google Scholar
Mills, S.J., Kampf, A.R., Christy, A.G., Housley, R.M., Rossman, G.R., Reynolds, R.E. and Marty, J. (2014) Bluebellite and mojaveite, two new minerals from the central Mojave Desert, California, USA. Mineralogical Magazine, 78, 13251340.Google Scholar
Materials Data, Inc. (2010) JADE 2010 Complete XRD Analysis. Livermore, CA, USA.Google Scholar
Missen, O.P., Mills, S.J., Spratt, J., Welch, M.D., Birch, W.D., Rumsey, M.S. and Vylita, J. (2018) The crystal structure determination and redefinition of eztlite, Pb2+2Fe3+3(Te4+O3)3(SO4)O2Cl. Mineralogical Magazine, 82, 13551367.Google Scholar
Pasero, M. (2018) The New IMA List of Minerals, http://cnmnc.main.jp/.Google Scholar
Pawley, G.S. (1981) Unit cell refinement from powder diffraction scans. Journal of Applied Crystallography, 14, 357361.Google Scholar
Sheldrick, G.M. (2015 a) Crystal structure refinement with SHELXL. Acta Crystallographica, C71, 38.Google Scholar
Sheldrick, G.M. (2015 b) SHELXT – integrated space-group and crystal-structure determination. Acta Crystallographica, A71, 38.Google Scholar
Weil, M. and Stöger, B. (2007) Redetermination of SrTe3O8 from a hydrothermally grown single crystal. Acta Crystallographica, E63, i116i118.Google Scholar
Weil, M., Shirkhanlou, M. and Stürzer, T. (2018) Phase formation studies of lead (II) copper (II) oxotellurates: The crystal structures of dimorphic PbCuTeO5, PbCuTe2O6, and [Pb2Cu2(Te4O11)](NO3)2. Zeitschrift für anorganische und allgemeine Chemie, 645, 347353.Google Scholar
Williams, S.A. (1975) Xocomecatlite, Cu3TeO4(OH)4, and tlalocite, Cu10Zn6(TeO3)(TeO4)2Cl (OH)25·27H2O, two new minerals from Moctezuma, Sonora, Mexico. Mineralogical Magazine, 40, 221226.Google Scholar
Williams, S.A. (1979) Girdite, oboyerite, fairbankite, and winstanleyite, four new tellurium minerals from Tombstone, Arizona. Mineralogical Magazine, 43, 453457.Google Scholar
Williams, S.A. and Cesbron, F.P. (1985) Yecoraite, Fe3Bi5(TeO3)(TeO4)2O9·nH2O, a new mineral from Sonora, Mexico. Boletín de Mineralogía, 1, 1016.Google Scholar
Williams, S.A. and Duggan, M. (1978) Tlapallite, a new mineral from Moctezuma, Sonora, Mexico. Mineralogical Magazine, 42, 183186.Google Scholar
Williams, S.A. and Gaines, R.V. (1975) Carlfriesite, H4Ca(TeO3)3, a new mineral from Moctezuma, Sonora, Mexico. Mineralogical Magazine, 40, 127130.Google Scholar
Supplementary material: File

Missen et al. supplementary material

Missen et al. supplementary material 1

Download Missen et al. supplementary material(File)
File 98.6 KB