Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T10:26:25.489Z Has data issue: false hasContentIssue false

Schorlomite and morimotoite: what's in a name?

Published online by Cambridge University Press:  16 June 2014

Sytle M. Antao*
Affiliation:
Department of Geoscience, University of Calgary, Calgary, Alberta T2N 1N4, Canada
*
a)Author to whom correspondence should be addressed. Electronic mail: antao@ucalgary.ca

Abstract

The crystal structure of an isotropic, single phase, Ti-rich schorlomite garnet, ideally Ca3Ti4+2(Fe3+2Si)O12, from Magnet Cove, Arkansas was refined with the Rietveld method, space group $Ia\overline 3 d$, and monochromatic synchrotron high-resolution powder X-ray diffraction data. Electron-microprobe analysis gave an average composition {Ca2.92Na0.04Mn2+0.03Mg0.01}Σ3[Ti1.04Fe3+0.46Mg0.18Fe2+0.16Zr4+0.13V3+0.04]Σ2(Si2.21Fe3+0.71Al0.07)Σ3O12, and corresponds to the general garnet formula of [8]X3[6]Y2[4]Z3[4]O12. Schorlomite is the dominant component, but the composition contains significant amounts (>15 mol.%) of andradite, Ca3(Fe3+2)Si3O12, morimotoite, Ca3(Ti4+Fe2+)Si3O12, and morimotoite-(Mg), Ca3(Ti4+Mg)Si3O12. The crystal structure model was refined well using isotropic and anisotropic displacement parameters. Using isotropic displacement parameters, the χ2 and R (F2) Rietveld refinement values are 1.148 and 0.0742, respectively. The unit-cell parameter, a = 12.18599(1) (Å), is large for a natural schorlomite for which complete structural data are available. The bond distances are average <Ca–O> = 2.4461, Ti–O = 2.0085(5), Si–O = 1.7022(5) Å, and site occupancy factors (sofs) for Ca(sof) = 0.963(1), Ti(sof) = 1.045(1), and Si(sof) = 1.150(2). Comparison of schorlomite data from the type locality in Magnet Cove with morimotoite from Ice River, Canada and the type locality in Japan show that they are quite similar and cast doubts as to morimotoite being different from schorlomite.

Type
Technical Articles
Copyright
Copyright © International Centre for Diffraction Data 2014 

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

Antao, S. M. (2013a). “Three cubic phases intergrown in a birefringent andradite–grossular garnet and their implications,” Phys. Chem. Miner. 40, 705716.Google Scholar
Antao, S. M. (2013b). “Can birefringent near end-member grossular be non-cubic? New evidence from synchrotron diffraction,” Can. Mineral. 51, 771784.Google Scholar
Antao, S. M. (2013c). “The mystery of birefringent garnet: is the symmetry lower than cubic?,” Powder Diffr. 28, 265272.Google Scholar
Antao, S. M. (2014). “Crystal structure of morimotoite from Ice River, Canada,” Powder Diffr. doi: 10.1017/S0885715614000414.Google Scholar
Antao, S. M. and Klincker, A. M. (2013a). “Origin of birefringence in andradite from Arizona, Madagascar, and Iran,” Phys. Chem. Miner. 40, 575586.CrossRefGoogle Scholar
Antao, S. M. and Klincker, A. M. (2013b). “Crystal structure of a birefringent andradite–grossular from Crowsnest Pass, Alberta, Canada,” Powder Diffr. 29, 2027.Google Scholar
Antao, S. M. and Round, S. A. (2014). “Crystal chemistry of birefringent spessartine,” Powder Diffr. doi: 10.1017/S0885715614000062.CrossRefGoogle Scholar
Antao, S. M., Hassan, I., Wang, J., Lee, P. L., and Toby, B. H. (2008). “State-of-the-art high-resolution powder X-ray diffraction (HRPXRD) illustrated with Rietveld structure refinement of quartz, sodalite, tremolite, and meionite,” Can. Mineral. 46, 15011509.Google Scholar
Armbruster, T., Birrer, J., Libowitzky, E., and Beran, A. (1998). “Crystal chemistry of Ti-bearing andradites,” Eur. J. Mineral. 10, 907921.Google Scholar
Chakhmouradian, A. R. and McCammon, C. A. (2005). “Schorlomite: a discussion of the crystal chemistry, formula, and inter-species boundaries,” Phys. Chem. Miner. 32, 277289.CrossRefGoogle Scholar
Deer, W. A., Howie, R. A., and Zussman, J. (1982). Rock-Forming Minerals: Orthosilicates (Longman Group Limited, New York), vol 1A, p. 919.Google Scholar
Flohr, M. J. K. and Ross, M. (1989). “Alkaline igneous rocks of Magnet Cove, Arkansas: metasomatized ijolite xenoliths from Diamond Jo quarry,” Am. Mineral. 74, 113131.Google Scholar
Grew, E. S., Locock, A. J., Mills, S. J., Galuskina, I. O., Galuskin, E. V., and Hålenius, U. (2013). “Nomenclature of the garnet supergroup,” Am. Mineral. 98, 785811.Google Scholar
Henmi, C., Kusachi, I., and Henmi, K. (1995). “Morimotoite, Ca3TiFe2+Si3O12, a new titanian garnet from Fuka, Okayama Prefecture, Japan,” Mineral. Mag. 59, 115120.CrossRefGoogle Scholar
Larson, A. C. and Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). (Report, LAUR 86–748). Los Alamos National Laboratory.Google Scholar
Lee, P. L., Shu, D., Ramanathan, M., Preissner, C., Wang, J., Beno, M. A., Von Dreele, R. B., Ribaud, L., Kurtz, C., Antao, S. M., Jiao, X., and Toby, B. H. (2008). “A twelve-analyzer detector system for high-resolution powder diffraction,” J. Synchrotron Radiat. 15, 427432.Google Scholar
Locock, A. J. (2008). “An excel spreadsheet to recast analyses of garnet into end-member components, and a synopsis of the crystal chemistry of natural silicate garnets,” Comput. Geosci. 34, 17691780.CrossRefGoogle Scholar
Locock, A., Luth, R. W., Cavell, R. G., Smith, D. G. W., and Duke, M. J. M. (1995). “Spectroscopy of the cation distribution in the schorlomite species of garnet,” Am. Mineral. 80, 2738.Google Scholar
Novak, G. A. and Gibbs, G. V. (1971). “The crystal chemistry of the silicate garnets,” Am. Mineral. 56, 17691780.Google Scholar
Peterson, R. C., Locock, A. J., and Luth, R. W. (1995). “Positional disorder of oxygen in garnet: the crystal-structure refinement of schorlomite,” Can. Mineral. 33, 627631.Google Scholar
Rass, I. T. (1997). “Morimotoite, a new titanian garnet? – discussion,” Mineral Mag. 61, 728730.Google Scholar
Rietveld, H. M. (1969). “A profile refinement method for nuclear and magnetic structures,” J. Appl. Crystallogr. 2, 6571.Google Scholar
Shannon, R. D. (1976). “Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides,” Acta Crystallogr. A32, 751767.Google Scholar
Toby, B. H. (2001). “EXPGUI, a graphical user interface for GSAS,” J. Appl. Crystallogr. 34, 210213.Google Scholar
Wang, J., Toby, B. H., Lee, P. L., Ribaud, L., Antao, S. M., Kurtz, C., Ramanathan, M., Von Dreele, R. B., and Beno, M. A. (2008). “A dedicated powder diffraction beamline at the advanced photon source: commissioning and early operational results,” Rev. Sci. Instrum. 79, 085105.Google Scholar
Weber, H. P., Virgo, D., and Huggins, F. E. (1975). “A neutron-diffraction and 57Fe Mössbauer study of a synthetic Ti-rich garnet,” Carnegie Inst. Wash Year Book 74, 575579.Google Scholar
Wills, A. S. and Brown, I. D. (1999). VaList. (CEA, France). This is a freely available computer program.Google Scholar