Hostname: page-component-7479d7b7d-qlrfm Total loading time: 0 Render date: 2024-07-11T08:55:37.314Z Has data issue: false hasContentIssue false

Site occupancy in richterite-winchite from Libby, Montana, USA, by FTIR spectroscopy

Published online by Cambridge University Press:  05 July 2018

G. Iezzi*
Affiliation:
Dipartimento Scienze della Terra, Università G. d’Annunzio, via Dei Vestini 30, I-66013 Chieti, Italy
G. Della Ventura
Affiliation:
Dipartimento Scienze Geologiche, Università Roma Tre, Italy
F. Bellatreccia
Affiliation:
Dipartimento Scienze Geologiche, Università Roma Tre, Italy
S. Lo Mastro
Affiliation:
Dipartimento Scienze Geologiche, Università Roma Tre, Italy
B. R. Bandli
Affiliation:
MVA Scientific Consultants, Duluth, Georgia, USA
M. E. Gunter
Affiliation:
Geological Sciences, University of Idaho, Moscow, Idaho, USA
*

Abstract

Three natural amphibole samples collected from the former vermiculite mine near Libby, Montana. USA, have been analysed by Rietveld X-ray powder diffraction (XRPD) refinement and Fourier transform infrared spectroscopy (FTIR) in the OH-stretching region. The same materials have been analysed previously by electron microprobe analysis (EMPA), Mössbauer spectroscopy and structure refinement (SREF) single crystal X-ray diffraction (SC-XRD), which revealed that these amphiboles have a crystal chemical formula very close to an intermediate composition between winchite and richterite, i.e. AA0.5BNaCaCMg4.5M3+T0.5Si8O22(OH)2 (A = Na and/or K; M3+ = Fe3+ and/or Al). The Rietveld analysis showed the powder samples used for the experiments here to be composed only of amphibole. This in turn allowed us to use FTIR OH-stretching data to derive cation ordering on these powder samples. The three FTIR spectra are quite similar and up to four components can be fitted to the patterns. The two lower-frequency components (labelled A and B) can be attributed to a local O(3)-H dipole surrounded by M(1)M(3)Mg3 and M(1)M(3)Mg2Fe2+; (respectively), an empty A site and rSi8 environments; on the contrary, the higher-frequency C and D bands indicate the presence of an occupied A site. The FTIR OH-stretching data alone allow us to calculate the site occupancy of the A, M(1)–M(3) and T sites with confidence, as compared with previously published data. By contrast M(4)- and M(2)-site occupancies are more difficult to evaluate. This study takes advantage of the large database of well characterized synthetic amphiboles, built over the last two decades. The comparison of vibrational spectroscopy data with micro-chemical and crystallographic data reported in this study demonstrate that the FTIR OH-stretching method alone is a valuable and rapid method to derive or at least sensibly constrain site occupancy for natural amphiboles. A much more detailed cation site occupancy can be obtained by combining micro-chemical and FTIR OH-stretching data.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2007

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

Bandli, B.R., Gunter, M.E., Twamley, B., Foit, F.F., Jr. and Cornelius, S.B. (2003) Optical, compositional, morphological, and X-ray data on eleven particles of amphibole from Libby, Montana, U.S.A. The Canadian Mineralogist , 41, 1241–1253.CrossRefGoogle Scholar
Burns, R.G. and Strens, R.G.J. (1966) Infrared study of the hydroxyl bonds in clinoamphiboles. Science, 153, 890–892.CrossRefGoogle Scholar
Deer, W.A., Howie, R.A. and Zussman, J. (1999) Rock-Forming Minerals, Double-Chain Silicates . Geological Society, London, 692 pp.Google Scholar
Delia Ventura, G. (1992) Recent developments in the synthesis and characterization of amphiboles. Synthesis and crystal chemistry of richterite. Trends in Mineralogy , 1, 153–192.Google Scholar
Delia Ventura, G., Robert, J.-L. and Hawthorne, F.C. (1996) Infrared spectroscopy of synthetic (Ni, Mg, Co)-potassium-richterite. Pp. 55–63 in: Mineral Spectroscopy: a Tribute to Roger G. Burns (Dyar, M.D. McCammon, C. and Schaefer, M.W., editors). Special Publication, No. 5, The Geochemical Society, Washington, D.C. Google Scholar
Delia Ventura, G., Robert, J.-L., Raudsepp, M., Hawthorne, F.C. and Welch, M.D. (1997) Site occupancies in synthetic monoclinic amphiboles: Rietveld structure refinement and infrared spectroscopy of (nickel, magnesium, cobalfj-richterite. American Mineralogist, 82, 291–301.Google Scholar
Delia Ventura, G., Hawthorne, F.C, Robert, J.-L., Del Bove, F., Welch, M.D. and Raudsepp, M. (1999) Short-range order of cations in synthetic amphiboles along the richterite-pargasite join. European Journal of Mineralogy, 11, 79–94.Google Scholar
Delia Ventura, G., Hawthorne, F.C, Robert, J.-L. and Iezzi, G. (2003) Synthesis and infrared spectroscopy of amphiboles along the tremolite-pargasite join. European Journal of Mineralogy, 15, 341–347.Google Scholar
Delia Ventura, G., Iezzi, G., Redhammer, G., Hawthorne, F.C Scaillet, B. and Novembre, D. (2005a) Synthesis and crystal-chemistry of amphiboles along the magnesioriebeckite–magnesio-arfvedsonite series as a function of fOl . American Mineralogist , 90, 1375–1383.Google Scholar
Delia Ventura, G., Redhammer, G., Iezzi, G., Hawthorne, F.C Papin, A. and Robert, J.-L. (20056) A Mössbauer and FTIR study of synthetic amphiboles along the magnesioriebeckite–ferri-clinoholmquistite join. Physics and Chemistry of Minerals, 32, 103–113.Google Scholar
Driscall, J., Jenkins, D.M., Dyar, M.D. and Bozhilov, K.N. (2005) Cation ordering in synthetic low-calcium actinolite. American Mineralogist, 90, 900–911.CrossRefGoogle Scholar
Gottschalk, M. and Andrut, M. (1998) Structural and chemical characterization of synthetic (Na, K)-richterite solid solutions by EMP, HRTEM, XRD and OH-valence vibrational spectroscopy. Physics and Chemistry of Minerals , 25, 101–111.CrossRefGoogle Scholar
Gottschalk, M., Andrut, M., and Melzer, S. (1999) The determination of the cummingtonite content of synthetic tremolite. European Journal of Mineralogy, 11, 967–982.CrossRefGoogle Scholar
Gunter, M.E., Dyar, M.D., Twamley, B., Foit, F.F. Jr., and Cornelius, S.B. (2003) Composition, Fe3+/ZFe, and crystal structure of non-asbestiform and asbesti-form amphiboles from Libby, Montana U.S.A. American Mineralogist, 88, 1970–1978.CrossRefGoogle Scholar
Hawthorne, F.C. (1983) The crystal chemistry of the amphiboles. The Canadian Mineralogist, 21, 173–480.Google Scholar
Hawthorne, F.C, Delia Ventura, G., Robert, J.-L., Welch, M.D., Raudsepp, M. and Jenkins, D. (1997) A Rietveld and infrared study of synthetic amphiboles along the potassium-richterite-tremolite join. American Mineralogist, 82, 708–716.CrossRefGoogle Scholar
Hawthorne, F.C, Welch, M.D., Delia Ventura, G., Shuangxi, Liu, Robert, J.-L. and Jenkins, D.M. (2000) Short-range order in synthetic aluminous tremolites: an infrared and triple-quantum MAS NMR study. American Mineralogist, 85, 1716–1724.CrossRefGoogle Scholar
Hawthorne, F.C, Delia Ventura, G., Oberti, R., Robert, J.-L. and Iezzi, G. (2005) Short-range order in minerals: amphiboles. The Canadian Mineralogist, 43, 1895–1920.CrossRefGoogle Scholar
Iezzi, G., Delia Ventura, G., Pedrazzi, G., Robert, J.-L. and Oberti, R. (2003a) Synthesis and characterisation of ferri-clinoferroholmqui stite, Li2(Fe +3Fe +2)Si8O22(OH)2. European Journal of Mineralogy , 15, 321–327.CrossRefGoogle Scholar
Iezzi, G., Delia Ventura, G., Cámara, F., Pedrazzi, G. and Robert, J.-L. (20036) BNa- BLi solid-solution in A-site vacant amphiboles: synthesis and cation ordering along the ferri-clinoferroholmquisti-te–riebeckite join. American Mineralogist, 88, 955–961.CrossRefGoogle Scholar
Iezzi, G., Cámara, F., Delia Ventura, G., Oberti, R., Pedrazzi, G. and Robert, J.-L. (2004a) Synthesis, crystal structure and crystal-chemistry of ferri-clinoholmquistite, Li2Mg2Fe3Si8O22(OH)2. Physics and Chemistry of Minerals , 31, 375–385.CrossRefGoogle Scholar
Iezzi, G., Delia Ventura, G., Oberti, R., Cámara, F. and Holtz, F. (20046) Synthesis and crystal-chemistry of Na(NaMg)Mg5Si8O22(OH)2, a P2xlm amphibole. American Mineralogist, 89, 640–646.Google Scholar
Iezzi, G., Delia Ventura, G., Hawthorne, F.C., Pedrazzi, G., Robert, J.-L. and Novembre, D. (2005a) The (Mg-Fe +) substitution in ferri-clinoholmquistite, Li2(Mg, Fe2+)Fe3+ 2Si8O22(OH)2 . European Journal of Mineralogy , 17, 733–740.CrossRefGoogle Scholar
Iezzi, G., Gatta, G.D., Kockelmann, W., Delia Ventura, G., Rinaldi, R., Schafer, W., Piccinini, M. and Gaillard, F. (20056) how-T neutron diffraction and synchrotron-radiation IR study of synthetic amphibole Na(NaMg)Mg5Si8O22(OH)2 . American Mineralogist, 90, 695–700.CrossRefGoogle Scholar
Iezzi, G., Tribaudino, M., Delia Ventura, G., Nestola, F. and Bellatreccia, F. (2005c) High-r phase transition of synthetic ANaB(LiMg)cMg5Si8O22(OH)2 . Physics and Chemistry of Minerals , 32, 515–523.CrossRefGoogle Scholar
Iezzi, G., Delia Ventura, G. and Tribaudino, M. (2006a) Synthetic P21/m amphiboles in the system Li2O-Na2O-MgO-SiO-H2O (LNMSH). American Mineralogist , 91, 425–429.CrossRefGoogle Scholar
Iezzi, G., Liu, Z. and Delia Ventura, G. (2006b) Synchrotron infrared spectroscopy of synthetic Na(NaMg)Mg5Si8O22(OH)2 up to 30 GPa: insight on a new high-pressure amphibole polymorph. American Mineralogist, 91, 479–482.CrossRefGoogle Scholar
Ishida, K. and Hawthorne, F.C. (2001) Assignment of infrared OH-stretching bands in manganoan magne-sio-arfedsonite and richterite through heat-treatments. American Mineralogist, 86, 965–972.CrossRefGoogle Scholar
Ishida, K., Hawthorne, F.C. and Ando, Y. (2002) Fine structure of infrared OH-stretching bands in natural and heat-treated amphiboles of the tremolite-ferro-actinolite series. American Mineralogist, 87, 891–898.CrossRefGoogle Scholar
Jenkins, D.M., Bozhilov, K.N. and Ishida, K. (2003) Infrared and TEM characterisation of amphiboles synthesized near the tremolite-pargasite join in the ternary system tremolite-pargasite-cummingtonite. American Mineralogist, 88, 1104–1114.CrossRefGoogle Scholar
Larson, A.C. and Von Dreele, R.B. (1997) GSAS: General Structure Analysis System . Document LAUR 86–748, Los Alamos National Laboratory, New Mexico, USA.Google Scholar
Leake, B.E., Woolley, A.R., Birch, W.D., Burke, E.A.J., Ferraris, G., Grice, ID., Hawthorne, F.C, Kisch, H.J., Krivovichev, V.G., Schumacher, J.C., Stephenson, N.C.N. and Whittaker, EJ.W. (2004) Nomenclature of amphiboles: additions and revisions to the International Mineralogical Association's amphibole nomenclature. The Canadian Mineralogist, 41, 1355–1362.Google Scholar
Meeker, G.P., Bern, A.M., Brownfield, I.K., Lowers, H.A., Sutley, S.J., Hoefen, T.M. and Vance, J.S. (2003) The composition and morphology of amphibole from the Rainy Creek Complex, near Libby, Montana. American Mineralogist, 88, 1955–1969.CrossRefGoogle Scholar
Melzer, S., Gottschalk, M., Andrut, M. and Heinrich, W. (2000) Crystal chemistry of K-richterite-richterite-tremolite solid solutions: a SEM, EMP, XRD, HRTEM and IR study. European Journal of Mineralogy, 12, 273–291.CrossRefGoogle Scholar
Najorka, J. and Gottschalk, M. (2003) Crystal chemistry of tremolite-tschermakite solid solutions. Physics and Chemistry of Minerals, 30, 108–124.CrossRefGoogle Scholar
Robert, J.-L., Delia Ventura, G. and Hawthorne, F.C. (1999) Near-infrared study of short-range disorder of OH and F in monoclinic amphiboles. American Mineralogist, 84, 86–91.CrossRefGoogle Scholar
Strens, R.S.J. (1974): The common chain, ribbon and ring silicates. Pp. 305–330 in: The Infrared Spectra of Minerals (Farmer, V.C., editor). Monograph, 4, Mineralogical Society, London.Google Scholar
Toby, B.H. (2001) EXPGUI, a Graphical User Interface for GSAS. Journal of Applied Crystallography, 34, 210–213.CrossRefGoogle Scholar
Wylie, A.G. and Verkouteren, J.R. (2000) Amphibole asbestos from Libby, Montana: Aspects of nomen-clature. American Mineralogist, 85, 1540–1542.CrossRefGoogle Scholar