Hostname: page-component-7479d7b7d-rvbq7 Total loading time: 0 Render date: 2024-07-10T12:20:20.903Z Has data issue: false hasContentIssue false

Nolzeite, Na(Mn,□)2[Si3(B,Si)O9(OH)2]·2H2O, a new pyroxenoid mineral from Mont Saint-Hilaire, Québec, Canada

Published online by Cambridge University Press:  02 January 2018

Monika M. M. Haring*
Affiliation:
Department of Earth Sciences, Laurentian University, Sudbury, Ontario P3E 2C6, Canada
Andrew M. McDonald*
Affiliation:
Department of Earth Sciences, Laurentian University, Sudbury, Ontario P3E 2C6, Canada

Abstract

Nolzeite, Na(Mn,□)2[Si3(B,Si)O9(OH)2]·2H2O, is a new mineral found in altered sodalite syenite at the Poudrette quarry, La Vallée-du-Richelieu, Montérégie (formerly Rouville County), Québec, Canada. Crystals are colourless to pale green and are acicular with average dimensions of 5 μm × 8 μm × 55 μm. They occur as radiating to loose, randomly oriented groupings within vugs associated with aegirine, nepheline, sodalite, eudialyte-group minerals, analcime, natron, pyrrhotite, catapleiite, steedeite and the unidentified mineral, UK80. Nolzeite is non-pleochroic, biaxial, with nmin = 1.616(2) and nmax = 1.636(2) and has a positive elongation. The average of six chemical analyses gave the empirical formula: Na1.04(Mn1.690.24Fe0.05Ca0.02)∑=2.00(Si2.96S0.04)∑=3.00(B0.70Si0.30)∑=1.00O9(OH)2·2H2O based on 13 anions. The Raman spectrum shows six distinct bands occurring at ∼3600–3300 cm–1 and 1600–1500 cm–1 (O–H and H–O–H bending), 1300–1200 cm–1 (B–OH bending), 1030–800 cm–1 (Si–O–Si stretching) as well as 700–500 cm–1 and 400–50 cm–1 (Mn–O and Na–O bonding, respectively). The FTIR spectrum for nolzeite shows bands at ∼2800 –3600 cm–1(O–H) stretching, a moderately sharp band at 1631 cm –1(H–O–H) bending, strong, sharp bands at ∼650 –700 cm–1, ∼800 –840 cm–1, and ∼900–1100 cm–1(Si–O and B –O) bonds. Nolzeite is triclinic, crystallizing in space group P with a = 6.894(1), b = 7.632(2), c = 11.017(2) Å, α= 108.39(3), β= 99.03, γ = 103.05(3)°, V = 519.27 Å3, and Z = 2. The crystal structure was refined to R = 12.37% and wR2 = 31.07% for 1361 reflections (Fo > 4σFo). It is based on chains of tetrahedra with a periodicity of three (i.e. a dreier chain) consisting of three symmetrically independent SiO4 tetrahedra forming C-shaped clusters closed by BO2(OH)2 tetrahedra, producing single loop-branched dreier borosilicate chains. The chains are linked through shared corners to double chains of edge-sharing MnO5(OH) octahedra. Nolzeite is a chain silicate closely related to steedeite and members of the sérandite–pectolite series. Paragenetically, nolzeite is late-stage, probably forming under alkaline conditions and over a narrow range of low pressures and temperatures.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2017

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

Bakakin, YY and Solo'eva, L.P. (1971) Crystal structure of Fe3BeSi3O9(F,OH)2, an example of a wollastonite-like silicon-oxygen chain based on Fe. Soviet Physics Crystallography, 15, 9991005.Google Scholar
Blakeman, E.A., Gard, J.A., Ramsay, C.G. and Taylor, H. FW. (1974) Studies on the system sodium oxide -calcium oxide - silica - water. Journal of Chemical Technology and Biotechnology, 24, 239245.Google Scholar
Brese, N.E. and O'Keeffe, M. (1991) Bond-valence parameters for solids. Acta Crystallographica, B47, 192197.CrossRefGoogle Scholar
Brugger, J., Krivovichev, S., Meisser, N., Ansermet, S. and Armbruster, T. (2006) Scheuchzerite, Na(Mn, Mg)9[VSi9O28(OH)](OH)3, a new single-chain silicate. American Mineralogist, 91, 937943.CrossRefGoogle Scholar
Chao, G.Y., Conlon, R.P. and Velthuizen, J. (1990) Mont Saint-Hilaire Unknowns. The Mineralogical Record, 21, 363368.Google Scholar
Clark, L.M. and Bunn, C.W. (1940) The scaling of boilers. Pt. IV. Identification of phases in calcium silicate scales. Journal of the Society of Chemical Industry, 59, 155158.Google Scholar
Cromer, D.T. and Liberman, D. (1970) Relativistic calculation of anomalous scattering factors for X rays. Journal of Physical Chemistry, 53, 18911898.CrossRefGoogle Scholar
Cromer, D.T. and Mann, J.B. (1968) X-ray scattering factors computed from numerical Hartree-Frock wave functions. Acta Crystallographica, A24, 321324.CrossRefGoogle Scholar
Czank, M. and Bissert, G. (1993) The crystal structure of Li2Mg2[Si4O11], a loop-branched dreier single chain silicate. Zeitschrift für Kristallographie, 204, 129142.Google Scholar
Dowty, E. (2009) VIBRATZfor Windows and Macintosh Version 2.2. Shape Software, Kingsport, Tennessee, USA.Google Scholar
Frisch, M.J., Trucks, G.W. Schlegel, H.B., et al. (2013) Gaussian 09, Revision D.01. Gaussian, Inc., Wallingford, Connecticut, USA.Google Scholar
Frost, R.L., Bouzaid, J.M., Martens, W.N. andReddy, J.B. (2007) Raman spectroscopy of the borosilicate mineral ferroaxinite. Journal of Raman Spectroscopy, 38, 135141.CrossRefGoogle Scholar
Haring, M.M.M. and McDonald, A.M. (2014) Steedeite, NaMn2[Si3BO9](OH)2: Characterization, crystal-structure determination, and origin. The Canadian Mineralogist, 52, 4760.CrossRefGoogle Scholar
Hawthorne, EC, Burns, P.C. and Grice, J.D. (1996) The crystal chemistry of boron. Pp. 41-110 in: Boron: Mineralogy, Petrology, and Geochemistry (L.M. Anovitz and E.S. Grew, editors) Reviews in Mineralogy, 33. Mineralogical Society of America, Washington DC.Google Scholar
Jacobsen, S.D., Smyth, J.R., Swope, J.R., and Sheldon, R. I (2000) Two proton positions in the very strong hydrogen bond of sérandite, NaMn2[Si3O8(OH)]. American Mineralogist, 85, 745752.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, 301303.CrossRefGoogle Scholar
Liebau, F. (1978) Silicates with branched anions: a crystallochemically distinct class. American Mineralogist, 63, 918923.Google Scholar
Mandarino, J.A. (1981) The Gladstone—Dale relationship. IV. The compatibility concept and its application. The Canadian Mineralogist, 19, 441450.Google Scholar
Maresch, W.V. and Czank, M. (1985) The optical and X-ray properties of Li2Mg2[Si4O11], a new type of chain-silicate. Neues Jahrbuch für Mineralogie, 7, 289297.Google Scholar
Marion, G.M. (2001) Carbonate mineral solubility at low temperatures in the Na-K-Mg-Ca-H-Cl-SO4-OH-HCO3-CO3-CO2-H2O system. Geochimica et Cosmochimica Acta, 65, 18831896.CrossRefGoogle Scholar
McDonald, A.M. and Chao, G.Y. (2005) Bobtraillite, (Na,Ca)13Srn(Zr,Y,Nb)14Si42B6Oi32(OH)12-12H2Oa new mineral species from Mont Saint-Hilaire, Québec: description, structure determination and relationship to benitoite and wadeite. The Canadian Mineralogist, 43, 747758.CrossRefGoogle Scholar
McDonald, A.M. and Chao, G.Y (2010) Rogermitchellite, Na12(Sr,Na)24Ba4Zr26Si78(B,Si)12O246(OH)24-18H2O, a new mineral species from Mont Saint-Hilaire, Québec: description, structure determination and relationship with HFSE-bearing cyclosilicates. The Canadian Mineralogist, 48, 267278.CrossRefGoogle Scholar
Nolze, G., Geist, V., Neumann, R.S. and Buchheim, M. (2005) Investigation of orientation relationships by EBSD and EDS on the example of the Watson iron meteorite. Crystallographic Research and Technology, 40, 791804.CrossRefGoogle Scholar
Prewitt, C.T (1967) Refinement of the crystal structure of pectolite, Ca2NaHSi3O9. Zeitschriftfür Kristallographie, 125, 298316.CrossRefGoogle Scholar
Schilling, J., Marks, M.A.W., Wenzel, T., Vennemann, T., Horvath, L., Tarassoff, P., Jacob, D.E., and Markl, G. (2011) Magmatic to hydrothermal evolution of the intrusive Mont Saint-Hilaire complex: insights into the late-stage evolution of peralkaline rocks. Journal of Petrology, 52, 21472185.CrossRefGoogle Scholar
Sheldrick, G.M. (2008) A short history of SHELX. Acta Crystallographica, A64, 112122.Google Scholar
Thompson, J.B. (1970) Geometrical possibilities for amphibole structures: Model biopyriboles. American Mineralogist, 63, 239249.Google Scholar
Waldemar, T.S. (1955) The pectolite-serandite series. American Mineralogist, 40, 10221031.Google Scholar
Williams, Q. (1995) Infrared, Raman and optical spectroscopy of Earth materials. Pp. 291-302 in: Mineral Physics and Crystallography: A Handbook of Physical Constants, (T.J. Ahrens, editor). American Geophysical Union, Washington, DC.Google Scholar
Xi, Y and Glasser, L.S.D. (1984) Hydrothermal study in the system Na2O-CaO-SiO2-H2O at 300 degrees celcius. Cement Concrete Research, 14, 741748.CrossRefGoogle Scholar