Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-19T01:52:50.464Z Has data issue: false hasContentIssue false

Marshallsussmanite, NaCaMnSi3O8(OH), a new pectolite-group mineral providing insight into hydrogen bonding in pyroxenoids

Published online by Cambridge University Press:  21 February 2018

Marcus J. Origlieri*
Affiliation:
Department of Geosciences, University of Arizona, Tucson, Arizona85721-0077, USA
Robert T. Downs
Affiliation:
Department of Geosciences, University of Arizona, Tucson, Arizona85721-0077, USA
Derek R. Hoffman
Affiliation:
Department of Geosciences, University of Arizona, Tucson, Arizona85721-0077, USA
Mihai N. Ducea
Affiliation:
Department of Geosciences, University of Arizona, Tucson, Arizona85721-0077, USA Faculty of Geology and Geophysics, University of Bucharest, Bulevardul N. Bălcescu 1, Bucharest030167, Romania
Jeffrey E. Post
Affiliation:
Department of Mineral Sciences, Smithsonian Institution, P.O. Box 37012, MRC 0119, Washington, D.C.20013-7012, USA
*
*Author for correspondence: Marcus J. Origlieri, Email: marcus@mineralzone.com

Abstract

Marshallsussmanite (IMA2013-067) is a new pyroxenoid mineral from the Wessels mine, Kalahari Manganese Field, Northern Cape Province, South Africa. Marshallsussmanite has ideal formula NaCaMnSi3O8(OH) and triclinic P$\bar{1}$ symmetry. Marshallsussmanite forms vitreous pink bladed crystals to 2.1 cm. The mineral shows perfect cleavage on both {100} and {001}. The chemical composition from electron microprobe (average of 20 analyses) and inductively coupled plasma mass spectrometer analysis (average of three analyses) is Li2O 0.43, Na2O 8.06, MgO 0.08, CaO 15.33, MnO 21.79, SiO2 51.71; totalling 97.40 wt.%. The empirical formula, normalized to 3 Si and assuming 1 H apfu is Li0.100Na0.906Ca0.953Mg0.007Mn1.071Si3O8(OH). Unit-cell parameters from single crystal X-ray diffraction are a = 7.7854(4), b = 6.9374(4), c = 6.8516(3) Å, α = 90.683(3)°, β = 94.330(3)°, γ = 102.856(3)°, V = 359.59(3) Å3; Z = 2. The crystal structure refinement converged with Robs = 0.0248 and site occupancy refinement gives crystal chemistry [Na0.948Li0.052][Ca0.793Mn0.207] [Mn0.937Ca0.063]Si3O8(OH). Marshallsussmanite is a single chain silicate with a repeat interval of three tetrahedra (i.e. dreier chain). Marshallsussmanite is a member of the pectolite group of pyroxenoids, which also includes barrydawsonite-(Y), cascandite, pectolite, serandite and tanohataite. Parallel silicate chains form layers, intercalated with well-ordered cation layers. Calcium and Mn both exhibit octahedral coordination, while Na has four bonded interactions in a coordination sphere (radius 3 Å) of seven separate O atoms. Procrystal electron density and bond valence modelling results are compared. The mineral has an unusually strong hydrogen bond with O4⋅⋅⋅O3 separation of 2.458(2) Å. Unlike pectolite and serandite, O4 in marshallsussmanite acts as an H-bond donor and O3 is an H-bond acceptor. Cation ordering in pyroxenoids has a substantial impact on the H position and corresponding H-bonding schemata.

Type
Article
Copyright
Copyright © The Author(s), 2018. Published by Cambridge University Press on behalf of The Mineralogical Society of Great Britain and Ireland

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: Michael Rumsey

This new mineral was originally approved by IMA in CNMNC newsletter 18, published in 2013 (IMA2013-067, Min. Mag., 77, 3256). Subsequent to prepublication on the journal's website, new information came to light and in May 2018 the CNMNC of the IMA passed memorandum IMA 18-B: Discreditation of the mineral name ‘marshallsussmanite’ with a reinstatement of the name schizolite (Newsletter 43, Min. Mag., 82, 785). An alternative paper was published by Grice et al. (2019). The original Mineralogical Magazine paper could not be withdrawn as it has already been cited. It is published here to complete the publication record. Readers' attention is directed to the updated paper (Grice et al., 2019, Min. Mag., 83, 473–478).

References

Arakcheeva, A., Pattison, P., Meisser, N., Chapuis, G., Pekov, I. and Thélin, P. (2007) New insight into the pectolite – serandite series: a single crystal diffraction study of Na(Ca1.73Mn0.27)[HSi3O9] at 293 and 100 K. Zeitschrift für Kristallographie, 222, 696704.10.1524/zkri.2007.222.12.696CrossRefGoogle Scholar
Bader, R.F.W. (1990) Atoms in Molecules: A Quantum Theory. Oxford University Press, UK, 458 pp.Google Scholar
Bonazzi, P., Bindi, L., Medenbach, O., Pagano, R., Lampronti, G.I. and Menchetti, S. (2007) Olmiite, CaMn[SiO3(OH)](OH), the Mn-dominant analogue of poldervaartite, a new mineral species from Kalahari manganese fields (Republic of South Africa). Mineralogical Magazine, 71, 193201.10.1180/minmag.2007.071.2.193CrossRefGoogle Scholar
Brese, N.E. and O'Keeffe, M.O. (1991) Bond-valence parameters for solids. Acta Crystallographica, B47, 192197.10.1107/S0108768190011041CrossRefGoogle 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, 244247.10.1107/S0108768185002063CrossRefGoogle Scholar
Buerger, M.J. (1956) The determination of the crystal structure of pectolite, Ca2NaHSi3O9. Zeitschrift für Kristallographie, 108, 248262.CrossRefGoogle Scholar
Cairncross, B. and Beukes, N.J. (2013) The Kalahari Manganese Field: The Adventure Continues. Assore. Struik Nature, Cape Town, South Africa, 384 pp.Google Scholar
Capitani, G.C. and Mellini, M. (2000) The johannsenite-hedenbergite complete solid solution: clinopyroxenes from the Campiglia Marittima skarn. European Journal of Mineralogy, 12, 12151227.10.1127/0935-1221/2000/0012-1215CrossRefGoogle Scholar
Carr, G.R., Phillips, E.R. and Williams, P.R. (1976) An occurrence of eudialyte and manganoan pectolite in a phonolite dyke from south-eastern Queensland. Mineralogical Magazine, 40, 853856.CrossRefGoogle Scholar
Chester, A.H. (1896) A Dictionary of the Names of Minerals Including Their History and Etymology. First Edition. John Wiley & Sons, New York, 320 pp.Google Scholar
Czank, M. and Bissert, G. (1993) The crystal structure of Li2Mg2[Si4O11], a loop-branched dreier single chain silicate. Zeitschrift für Kristallographie, 204, 129142.10.1524/zkri.1993.204.Part-1.129CrossRefGoogle Scholar
Dai, Y., Harlow, G.E. and McGhie, A.R. (1993) Poldervaartite, Ca(Ca0.5Mn0.5)(SiO3OH)(OH), a new acid nesosilicate from the Kalahari manganese field, South Africa: Crystal structure and description. American Mineralogist, 78, 10821087.Google Scholar
Dana, E.S. (1892) The System of Mineralogy. 6th Edition. John Wiley & Sons, New York, 1134 pp.Google Scholar
Downs, R.T. (2000) Analysis of harmonic displacement factors. Pp. 6188 in: High-Temperature and High-Pressure Crystal Chemistry (Hazen, R.M. and Downs, R.T., editors). Reviews in Mineralogy & Geochemistry, 41. Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA.CrossRefGoogle Scholar
Downs, R.T., Andalman, A. and Hudacsko, M. (1996) The coordination numbers of Na and K atoms in low albite and microcline as determined from a procrystal electron-density distribution. American Mineralogist, 81, 13441349.10.2138/am-1996-11-1206CrossRefGoogle Scholar
Gutzmer, J. and Beukes, N.J. (1996) Mineral paragenesis of the Kalahari manganese field, South Africa. Ore Geology Reviews, 11, 405428.10.1016/S0169-1368(96)00011-XCrossRefGoogle Scholar
Hintze, C. (1897) Handbuch der Mineralogie. Zweiter Band. Silicate und Titanate. Verlag von Veit & Comp., Leipzig, Germany, 1841 pp.Google Scholar
Jacobsen, S.D., Smyth, J.R., Swope, R.J. and Sheldon, R.I. (2000) Two proton positions in the very strong hydrogen bond of serandite, NaMn2[Si3O8(OH)]. American Mineralogist, 85, 745752.10.2138/am-2000-5-613CrossRefGoogle Scholar
Kolitsch, U., Blass, G., Jahn, S., Cámara, F., Von Bezing, L., Wartha, R.R., Tremmel, G., Sturla, M., Cerea, P., Skebo, M. and Ciriotti, M.E. (2016) Aris. Mineralogy of the Famous Alkaline Phonolite. Associazione Micromineralogia Italiana, Cremona, Italy, 95 pp.Google Scholar
Lacroix, A. (1931) Les pegmatites de la syénite sodalitique de l'île Rouma (archipel de Los, Guinée française). Description d'un nouveau minéral (sérandite) qu'elles renferment. Comptes Rendus Hebdomadaires des Séances de l'Académie des Sciences, 192, 189194.Google Scholar
Lafuente, B., Downs, R.T., Yang, H. and Stone, N. (2015) The power of databases: the RRUFF project. Pp. 130 in: Highlights in Mineralogical Crystallography (Armbruster, T. and Danisi, R.M., editors). De Gruyter, Berlin, Germany.Google Scholar
Libowitzky, E. (1999) Correlation of O–H stretching frequencies and O–H⋅⋅⋅O hydrogen bond lengths in minerals. Monatshefte fur Chemie, 130, 10471059.10.1007/BF03354882CrossRefGoogle Scholar
Liebau, F. (1980) The role of cationic hydrogen in pyroxenoid crystal chemistry. American Mineralogist, 65, 981985.Google Scholar
Mellini, M. and Merlino, S. (1982) The crystal structure of cascandite, CaScSi3O8(OH). American Mineralogist, 67, 604609.Google Scholar
Mellini, M., Merlino, S., Orlandi, P. and Rinaldi, R. (1982) Cascandite and jervisite, two new scandium silicates from Baveno, Italy. American Mineralogist, 67, 599603.Google Scholar
Mitchell, R.H., Welch, M.D., Kampf, A.R., Chakhmouradian, A.K. and Spratt, J. (2015) Barrydawsonite-(Y), Na1.5CaY0.5Si3O9H: a new pyroxenoid of the pectolite–serandite group. Mineralogical Magazine, 79, 671686.10.1180/minmag.2015.079.3.12CrossRefGoogle Scholar
Nagase, T., Hori, H., Kitamine, M., Nagashima, M., Abduriyim, A. and Kuribayashi, T. (2012) Tanohataite, LiMn2Si3O8(OH): a new mineral from the Tanohata mine, Iwate Prefecture, Japan. Journal of Mineralogical and Petrological Sciences, 107, 149154.10.2465/jmps.111130CrossRefGoogle Scholar
Nagashima, M., Armbruster, T., Kolitsch, U. and Pettke, T. (2014 a) The relation between Li ↔ Na substitution in five-periodic single-chain silicates nambulite and marsturite: a single crystal X-ray study. American Mineralogist, 99, 14621470.CrossRefGoogle Scholar
Nagashima, M., Mitani, K. and Akasada, M. (2014 b) Structural variation of babingtonite depending on cation distribution at the octahedral sites. Mineralogy and Petrology, 108, 287301.10.1007/s00710-013-0297-zCrossRefGoogle Scholar
Ohashi, Y. and Finger, L.W. (1978) The role of octahedral cations in pyroxenoid crystal chemistry. I. Bustamite, wollastonite, and the pectolite–schizolite–serandite series. American Mineralogist, 63, 274288.Google Scholar
Peacock, M.A. (1935) On pectolite. Zeitschrift für Kristallographie, 90, 97111.10.1524/zkri.1935.90.1.97CrossRefGoogle Scholar
Peacor, D.R., Essene, E.J. and Gaines, A.M. (1987) Petrologic and crystal-chemical implications of cation order-disorder in kutnahorite [CaMn(CO3)2]. American Mineralogist, 72, 319328.Google Scholar
Pouchou, J.L. and Pichoir, F. (1984) Un nouveau modèle de calcul pour la microanalyse quantitative par spectrométrie de rayons X. Partie I: Application a l'analyse d’énchantillons homogènes. La Recherche Aérospatiale, 3, 167192.Google Scholar
Prewitt, C.T. (1967) Refinement of the structure of pectolite, Ca2NaHSi3O9. Zeitschrift für Kristallographie, 125, 298316.CrossRefGoogle Scholar
Robinson, K., Gibbs, G.V. and Ribbe, P.H. (1971) Quadratic elongation: A quantitative measure of distortion in coordination polyhedra. Science, 172, 567570.10.1126/science.172.3983.567CrossRefGoogle ScholarPubMed
Schaller, W.T. (1955) The pectolite–schizolite–serandite series. American Mineralogist, 40, 10221031.Google Scholar
Sheldrick, G.M. (2008) A short history of SHELX. Acta Crystallographica, A64, 112122.10.1107/S0108767307043930CrossRefGoogle Scholar
Takéuchi, Y. and Kudoh, Y. (1977) Hydrogen bonding and cation ordering in Magnet Cove pectolite. Zeitschrift für Kristallographie, 146, 281292.10.1524/zkri.1977.146.4-6.281CrossRefGoogle Scholar
Takéuchi, Y., Kudoh, Y. and Yamanaka, T. (1976) Crystal chemistry of the serandite–pectolite series and related minerals. American Mineralogist, 61, 229237.Google Scholar
Thompson, R.M. and Downs, R.T. (2004) Model pyroxenes II: Structural variation as a function of tetrahedral rotation. American Mineralogist, 89, 614628.10.2138/am-2004-0416CrossRefGoogle Scholar
von Kobell, F. (1828) Ueber den Pektolith. Archiv für die gesammte Naturlehre, 13, 385393.Google Scholar
Warren, B.E. and Biscoe, J. (1931) The crystal structure of the monoclinic pyroxenes. Zeitschrift für Kristallographie, 80, 391401.10.1524/zkri.1931.80.1.391CrossRefGoogle Scholar
Williams, J.F. (1891) Manganopektolith, ein neues Pektolith-ähnliches Mineral von Magnet Cove, Arkansas. Zeitschrift für Kristallographie, 18, 386389.10.1524/zkri.1891.18.1.386CrossRefGoogle Scholar
Williams, E.R. and Weller, M.T. (2014) A variable-temperature neutron diffraction study of serandite: A Mn-silicate framework with a very strong, two-proton site, hydrogen bond. American Mineralogist, 99, 17551760.CrossRefGoogle Scholar
Winther, C. (1901) Schizolite, a new mineral. Meddelelser om Grønland, 24, 196203.Google Scholar
Yang, H., Downs, R.T. and Yang, Y.W. (2011) Lithiomarsturite, LiCa2Mn2Si5O14(OH). Acta Crystallographica, E67, i73.Google Scholar
Supplementary material: File

Origlieri et al. supplementary material 1

Origlieri et al. supplementary material

Download Origlieri et al. supplementary material 1(File)
File 37.4 KB