Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-25T01:30:54.098Z Has data issue: false hasContentIssue false

Quantitative determination of chrysotile asbestos in bulk materials by combined Rietveld and RIR methods

Published online by Cambridge University Press:  10 January 2013

A. Gualtieri
Affiliation:
Università di Modena, Dipartimento di Scienze della Terra, I-41100 Modena, Italy
G. Artioli
Affiliation:
Università di Milano, Dipartimento di Scienze della Terra, I-20133 Milano, Italy

Abstract

Because of their potential to induce a number of pathological diseases and their widespread industrial usage in the past, the fibrous minerals forming asbestos have been the subject of a number of studies in the past. Although quantification of asbestos minerals by optical and electron microscopy (SEM, TEM) is a routine technique in the case of dispersed airborn fibers, the detection and the quantification of small amount of fibrous minerals like chrysotile in bulk materials such as building materials is exceedingly difficult. A method for the detection and evaluation of asbestos minerals in massive samples is described, based on a combination of Rietveld and RIR (Reference Intensity Ratio) methods. Lower detection limits are about 0.5-1.0 wt % for chrysotile, depending on powder pattern, counting statistics, and matrix absorption. The chrysotile wt % determined on powder diffraction profiles collected on a conventional instrument is precise to about 1.0 wt % absolute (relative error in the range 0-10%). The technique is of straightforward application. If compared with the commonly used microscopic or spectroscopic techniques, it is of much advantage from the point of view of time, and the results are more accurate and statistically significant of the bulk material. A model for the cylindrically disordered structure of fibrous chrysotile is especially developed for the simulation of the X-ray powder patterns, and it is proposed here.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1995

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

Armstrong, J.T. (1991). “Quantitative elemental analysis of individual microparticles with electron beam instruments,” in Electron probe Quantitation. edited by Heinrich, K.F.J. and Newbury, D.E. (Plenum, New York), pp. 261316.CrossRefGoogle Scholar
Artioli, G., Alberti, G., Cagossi, G., and Bellotto, M. (1991). “Quantitative determination of crystalline and amorphous components in clinoptiloliterich rocks by Rietveld analysis of X-ray powder diffraction profiles,” in Atti l ° Convegno Italiano di Scienza e Tecnologia delle Zeoliti, edited by Colella, C., (De Frede, Napoli), pp. 261270.Google Scholar
Baronnet, A., Belluso, E., and Ferraris, G. (1994a). “HRTEM study of carlosturanite: new polysomes and microstructural relationships with associated phases,” Proc. 16th General Meeting Int. Min. Ass. Pisa, Italy 4–9 September 1994.Google Scholar
Baronnet, A., Mellini, M., and Devouard, B. (1994b). “Sectors in polygonal serpentine. A model based on dislocations,” Phys. Chem. Minerals 21, 330343.CrossRefGoogle Scholar
Belluso, E., and Ferraris, G. (1991). “New data on balangeroite and carlosturanite from alpine serpentinites,” Eur. J. Mineral. 3, 559566.CrossRefGoogle Scholar
Bish, D.L., and Howard, S.A. (1988). “Quantitative phase analysis using the Rietveld method,” J. App. Cryst. 21, 8691.CrossRefGoogle Scholar
Bish, L.B., and Post, J.E. (1993). “Quantitative mineralogical analysis using the Rietveld full-pattern fitting method,” Am. Mineral 78, 932940.Google Scholar
Chisholm, J.E. (1991). “Geometrical constraints on the growth of sectors in polygonal serpentine,” J. Phys. D. 24, 199202.CrossRefGoogle Scholar
Chisholm, J.E. (1992). “The number of sectors in polygonal serpentine,” Can. Mineral. 30, 355365.Google Scholar
Deriu, A., Ferraris, G., and Belluso, E. (1994). “Fe Mössbauer study of the asbestiform silicates balangeroite and carlosturanite.” Phys. and Chem. Minerals 21, 222227.CrossRefGoogle Scholar
Devouard, B., and Baronnet, A. (1993). “Five fold symmetry in chrysotile,” Terra Abstr. Suppl. 1 Terra Nova 5, 351359.Google Scholar
Dollase, W.A. (1986). “Correction of intensities for preferred orientation in powder diffractometry: application for the March model,” J. Appl. Cryst. 19, 267272.CrossRefGoogle Scholar
Dunn, H.W., and Stewart, J.H. (1982). “Determination of chrysotile asbestos in building materials by X-ray diffractometry,” Anal. Chem. 54, 11221125.CrossRefGoogle Scholar
Effenberger, M., Mereiter, K., and Zeeman, J. (1981). “Crystal structure refinement of magnesite, calcite, rhodocrosite, siderite, smithsonite, and dolomite, with discussion of some aspects of the stereochemistry of calcite type carbonates,” Zeit. Krist. 156, 233243.Google Scholar
Gasden, J.A., Parker, J., Smith, W.L. (1970). “Determination of chrisotile in Airborne Asbestos Dusts by an Infrared Spectrometric Technique,” Atm. Env. 4, 667670.CrossRefGoogle Scholar
Griffiths, P.R. (1987). “Mid-infrared fourier transform spectrometry,” in Laboratory Methods in Vibrational Spectroscopy, edited by Willis, H.A., van der Maas, J.H., and Miller, R.G.J., (Wiley, New York), pp. 121143.Google Scholar
Guthrie, G.D. (1992). “Biological effects of inhaled minerals,” Am. Mineral 77, 225243.Google Scholar
Guthrie, G.D. (1993). “Mineral characterization in biological studies,” in Health Effects of Mineral Dusts, edited by Guthrie, G.D. and Mossman, B.T. (Minerological Society of America), Vol. 28, pp. 61137.CrossRefGoogle Scholar
Güven, N. (1971). “The crystal structure of 2m1 phengite and 2m1, muscovite,” Zeit. Krist. 134, 196212.Google Scholar
Hill, R.J. (1991). “Expanded use of the Rietveld method in studies of phase abundance in multiphase mixtures,” Powder Diffr. 6, 7477.CrossRefGoogle Scholar
Ishiżawa, N., Miyata, T., Marumo, F., and Iway, S. (1980). “A structural investigation of αAl203 at 2170 K,” Acta Cryst. B36, 228230.CrossRefGoogle Scholar
Kane, A.B. (1993). “Epidemiology and pathology of asbestos-related diseases,” in Health Effects of Mineral Dusts, edited by Guthrie, G.D. and Mossman, B.T. (Mineralogical Society of America), Vol. 28, pp. 347359.CrossRefGoogle Scholar
Kunze, G. (1961). “Antigorit. Strukturtheoretische Grundlangen und ihre praktisce bedeutung fúr die weiters serpentin-forschung,” Fortschr. Min. 39, 206324.Google Scholar
Larson, A.C., and Von Dreele, R.B. (1994). “GSAS. General Structure Analysis System,” Report LAUR 86-748, Los Alamos National Laboratory, NM.Google Scholar
Le Page, Y., and Donnay, G. (1976). “Refinement of the crystal structure of low-quartz,” Acta Cryst. B 32, 24762480.CrossRefGoogle Scholar
Luoma, G.A., Yee, L.K., and Rowland, R. (1982). “Determination of microgram amounts of asbestos in mixtures by infrared spectrometry,” Anal. Chem. 54, 21402142.CrossRefGoogle Scholar
Mangia, A. (1980). “A new approach to the problem of kaolinite interference in the determination of chrysotile asbestos by means of X-ray diffraction,” Anal. Chim. Acta 117, 337342.CrossRefGoogle Scholar
Mellini, M., Ferraris, G., and Compagnoni, R. (1985). “Carlosturanite: HRTEM evidence of a polysomatic series including serpentine,” Am. Mineral. 70, 773781.Google Scholar
Middleton, A.P. and Whittaker, J.W. (1976). “The structure of Povlen- type chrysotile,” Can. Mineral. 14, 301306.Google Scholar
Nakahira, M. (1952). “On the crystal structure of montmorillonite, with some references to other clay minerals,” J. Sci. Res. Inst. Tokyo, 46, 268287.Google Scholar
Puledda, S., and Marconi, A. (1988). “Determinazione quantitativa dell'asbesto con la diffrattometria a R.X. secondo il metodo del filtro d'argento. I. Applicazione al crisotilo,” Giorn. Ig. Ind., 13, 732.Google Scholar
Rietveld, H.M. (1969). “A profile refinement for nuclear and magnetic structures,” J. App. Cryst. 2, 6571.CrossRefGoogle Scholar
Rucklidge, J.C., and Zussman, J. (1965). “The crystal structure of the serpentine mineral, lizardite Mg3Si2O5(OH)4,” Acta Cryst. 19, 381389.CrossRefGoogle Scholar
Skinner, H.C.W., Ross, M., and Frondel, C. (1988). Asbestos and Other Fibrous Materials: Mineralogy, Crystal Chemistry and Health Effects, (Oxford U.P., New York).Google Scholar
Skinner, H.C.W. (1994). “Reactions of asbestos and other fibrous materials,” Proc. 16th General Meeting Int. Min. Ass. Pisa, Italy 4–9 September 1994.Google Scholar
Snyder, R.L. (1992). “The use of the reference intensity ratio in X-ray quantitative analysis,” Powder Diffr. 7, 186193.CrossRefGoogle Scholar
Snyder, R.L. and Bish, D.L. (1989). “Quantitative analysis,” in Modern Powder Diffraction, edited by Bish, D. L., and Post, J. E., (Mineralogy Society of America Reviews in Mineralogy), Vol. 20, pp. 101144.CrossRefGoogle Scholar
Suquet, H. (1989). “Effects of dry grinding and leaching on the crystal structure of chrysotile,” Clays and Clay Mineral. 37, 439445.CrossRefGoogle Scholar
Titulaer, M.K., van Mintelburg, J.C., Jansen, J.B., and Geus, J.W. (1993). “Characterization of tubular chrysotile by thermoporometry, nitrogen sorption, drifts and TEM,” Clays and Clay Mineral. 4, 496513.CrossRefGoogle Scholar
Veblen, D.R., and Wylie, A.G. (1993). “Mineralogy of amphiboles and 1:1 layer silicates,” in Health Effects of Mineral Dusts, edited by Guthrie, G.D. and Mossman, B.T. (Mineralogy Society of America), Vol. 28, pp. 61137.CrossRefGoogle Scholar
Wagner, J.C., Sleggs, C.A., and Marchand, P. (1960). “Diffuse pleural mesothelioma and asbestos exposure in the North Western Cape Province,” Br. J. Ind. Med. 17, 260271.Google ScholarPubMed
Wicks, F.J., and O'Hanley, D.S. (1988). “Serpentine minerals: structures and petrology,” in Hydrous Phyllosilicates, edited by Bailey, S.W. (Mineralogy Society of America), Vol. 19, pp. 91188.CrossRefGoogle Scholar
Wicks, F.J., and Whittaker, E.J.W. (1977). “Serpentine textures and serpentinization,” Can. Min. 15, 459488.Google Scholar
Wylie, A.G. (1993). “Modeling asbestos populations: a fractal approach,” Can. Min. 30, 437446.CrossRefGoogle Scholar
Whittaker, E.J.W. (1955). “A classification of cylindrical lattices,” Acta Cryst. 6, 571574.CrossRefGoogle Scholar
Whittaker, E.J.W. (1956a). “The structure of chrysotile. II. Clinochrysotile,” Acta Cryst. 9, 855861.CrossRefGoogle Scholar
Whittaker, E.J.W. (1956b). “The structure of chrysotile. III. Orthochrysotile,” Acta Cryst. 9, 862864.CrossRefGoogle Scholar
Whittaker, E.J.W. (1956c). “The structure of Chrysotile. IV. Para-Chrysotile,” Acta Cryst. 9, 865867.CrossRefGoogle Scholar
Whittaker, E.J.W. (1957). “The structure of chrysotile. V: Diffuse reflections and fiber texture,” Acta Cryst. 10, 149156.CrossRefGoogle Scholar
Whittaker, E.J.W., and Zussman, J. (1956). “The characterization of serpentine minerals by X-ray diffraction,” Mineral. Mag. 31, 107126.Google Scholar
Winter, J.K., Ghose, S., and Okamura, F.P. (1977). “A high temeprature study of the thermal expansion and the anisotropy of the sodium atom in low albite,” Am. Mineral. 62, 921931.Google Scholar
Yariv, S. and Heller-Kallai, L. (1975). “The relationship between the I.R. spectra of serpentines and their structures,” Clays and Clay Min. 23, 145152.CrossRefGoogle Scholar
Young, R.A. (1993). The Rietveld Method (Oxford U. P., New York).CrossRefGoogle Scholar