Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-26T09:48:13.925Z Has data issue: false hasContentIssue false

A practical approach to Rietveld analysis. Comparison of some programs running on personal computers

Published online by Cambridge University Press:  10 January 2013

Lj. Karanović
Affiliation:
Crystallographic Laboratory, Faculty of Mining and Geology, University of Belgrade, Djušina 7, 11000 Belgrade, Yugoslavia
I. Petrović-Prelević
Affiliation:
Crystallographic Laboratory, Faculty of Mining and Geology, University of Belgrade, Djušina 7, 11000 Belgrade, Yugoslavia
D. Poleti
Affiliation:
Department of General and Inorganic Chemistry, Faculty of Technology and Metallurgy, University of Belgrade, P.O. Box 494, 11001 Belgrade, Yugoslavia

Abstract

Three computer programs for Rietveld analysis DBWS-9411, HILL-93.06, and FULLPROF-3.1 have been tested and compared using data for two samples of different complexity, spinel, and anglesite. The investigation shows that results are “program dependent.” The obtained R indices for spinel are in the ranges 10.60%–13.26%(Rwp) and 3.15%–5.25%(RB). Similarly, the ranges for anglesite are 9.76%–14.06%(Rwp) and 2.15%–5.06%(RB). Some atom and displacement parameters are significantly different, too. In attempt to define the standard procedure for Rietveld analysis, three parameters, n, BKPOS, and RLIM were examined. It was found that the most appropriate values for them are: n=8–10, BKPOS=90°, and RLIM=40° 2θ. Using Fourier filtering for background modeling, significantly lower R indices were obtained in comparison to polynomial and interpolated background. At the same time the great numbers of atom and cell parameters agree within ±3σmax and e.s.d.'s were identical or lower than those achieved for polynomial and interpolated background. It was found that the function given by Bérar and Baldinozzi (1993) (J. Appl. Crystallogr. 26, 128–129) much better described asymmetric peak profiles at low 2θ angles. This function and Fourier filtering were implemented only in FULLPROF, which has more possibilities and some advantages over the other two programs. In addition, the peak shift parameters (sample displacement and transparency) were tested. It was shown that under present circumstances these parameters do not have much effect on atom parameters and R indices. However, differences in unit cell parameters were considerable greater, most probably because of the large correlation between zero-point, lattice, and peak shift parameters.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1999

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

Andreev, Yu. G. (1994). “The use of the serial-correlations concept in the figure-of-merit function for powder diffraction profile fitting,” J. Appl. Crystallogr. 27, 288297.CrossRefGoogle Scholar
Baerlocher, Ch., and McCusker, L. B. (1994). “Practical aspects of powder diffraction data analysis,” Stud. Surf. Sci. Catal. 85, 391428.CrossRefGoogle Scholar
Bérar, J. F., and Baldinozzi, G. (1993). “Modeling of line-shape asymmetry in powder diffraction,” J. Appl. Crystallogr. 26, 128129.CrossRefGoogle Scholar
Bérar, J. F., and Lelann, P. (1991). “E.S.D.'s and estimated probable error obtained in Rietveld refinements with local correlations,” J. Appl. Crystallogr. 24, 15.CrossRefGoogle Scholar
David, W. I. F. (1986). “Powder diffraction peak shapes. Parametrization of the pseudo-Voigt as a Voigt function,” J. Appl. Crystallogr. 19, 6364.CrossRefGoogle Scholar
David, W. I. F., and Matthewman, J. C. (1985). “Profile refinement of powder diffraction patterns using the Voigt function,” J. Appl. Crystallogr. 18, 461466.CrossRefGoogle Scholar
Dollase, W. A. (1986). “Correction of intensities for preferred orientation in powder diffractometry: Application of the March model,” J. Appl. Crystallogr. 19, 267272.CrossRefGoogle Scholar
Gorter, S., and Smith, D. K. (1995). “World directory of powder diffraction Programs,” Report to the IUCr Commission on Powder Diffraction, Published by the Program Exchange Bank, Release 2.2, pp. 24–26.Google Scholar
Hill, R. J. (1992). “International Union of Crystallography, Commission on Powder Diffraction, Rietveld Refinement Round Robin. I. Analysis of standard X-ray and neutron data for PbSO 4,” J. Appl. Crystallogr. 25, 589610.CrossRefGoogle Scholar
Hill, R. J., and Cranswick, L. M. D. (1994). “International Union of Crystallography, Commission on Powder Diffraction, Rietveld Refinement Round Robin. II. Analysis of monoclinic ZrO 2,” J. Appl. Crystallogr. 27, 802844.CrossRefGoogle Scholar
Hill, R. J., and Fischer, R. X. (1990). “Profile agreement indices in Rietveld and pattern-fitting analysis,” J. Appl. Crystallogr. 23, 462468.CrossRefGoogle Scholar
Hill, R. J., and Howard, C. J. (1986). “A computer program for Rietveld analysis of fixed-wavelength X-ray and neutron powder diffraction patterns,” Australian Atomic Energy Commission (now ANSTO) Report No. M112, Lucas Heights Research Laboratories, New South Wales, Australia.Google Scholar
Howard, C. J. (1982). “The approximation of asymmetric neutron powder diffraction peaks by sums of Gaussians,” J. Appl. Crystallogr. 15, 615620.CrossRefGoogle Scholar
Howard, S. A., and Preston, K. D. (1989).Modern Powder Diffraction, edited by D. L. Bish, and J. E. Post(Rev. Mineral. 20, 217275).Google Scholar
Ilic, A., Antic, B., Poleti, D., Rodic, D., Petrović-Prelević, I., and Karanović, Lj. (1996). “Cation distribution and magnetic properties of ternary Zn 2.33−xCo xSb 0.67O 4 spinels,” J. Phys.: Condens. Matter 8, 23172325.Google Scholar
Jansen, E., Schäfer, W., and Will, G. (1994). “R values in analysis of powder diffraction data using Rietveld refinement,” J. Appl. Crystallogr. 27, 492496.CrossRefGoogle Scholar
Kogan, V. A., and Kupriyanov, M. F. (1992). “X-ray powder diffraction line profiles by Fourier synthesis,” J. Appl. Crystallogr. 25, 1625.CrossRefGoogle Scholar
Larson, A. C., and Von Dreele, R. B. (1986). “GSAS. Generalized Structure Analysis System,” Manual LAUR 86-748, Los Alamos National Laboratory, Los Alamos, New Mexico.Google Scholar
Lengauer, C. L. (1993). “HILL. A computer program for Rietveld Analysis. PC-Implementation of LHPM8, Version 93.06,” University of Vienna, Institute of Mineralogy and Crystallography, Vienna, Austria.Google Scholar
Miyake, M., Minato, I., Morikawa, H., and Iwai, S. (1978). “Crystal structures and sulphate force constants of barite, celestite, and anglesite,” Am. Mineral. 63, 506510.Google Scholar
Poleti, D., Vasović, D., Karanović, Lj., and Branković, Z. (1994). “Synthesis and characterization of ternary zinc-antimony-transition metal spinels,” J. Solid State Chem. 112, 3944.CrossRefGoogle Scholar
Prince, E. (1981). “Comparison of profile and integrated-intensity methods in powder refinement,” J. Appl. Crystallogr. 14, 157159.CrossRefGoogle Scholar
Richardson, J. W. (1993). The Rietveld Method, edited by R. A. Young, IUCr Monograph on Crystallography No. 5 (IUCr/Oxford University Press, Oxford), pp. 102–110.Google Scholar
Rietveld, H. M. (1966). “A method for including the line profiles of neutron powder diffraction peaks in the determination of crystal structures,” Abstracts of the Seventh Congress and Symposium of the IUCr, Moscow 1966, Acta Crystallogr. 21,A228.Google Scholar
Rietveld, H. M. (1969a). “A profile refinement method for nuclear and magnetic structures,” J. Appl. Crystallogr. 2, 6571.CrossRefGoogle Scholar
Rietveld, H. M. (1969b). “An ALGOL program for refinement of nuclear and magnetic structures by the profile method,” Reactor Centrum Nederland Research Report No. 104, RCN, Petten, The Netherlands.Google Scholar
Rodríguez-Carvajal, J. (1990). “FULLPROF: A program for Rietveld refinement and pattern matching analysis,” Powder Diffraction, Abstracts of the Satellite Meeting of the XVth Congress of the IUCr, Toulouse, France.Google Scholar
Rodríguez-Carvajal, J. (1995). User's Guide to Program FULLPROF, Version 3.0.0 Apr95-LLB-JRC (Laboratoire Léon Brillouin, CEA-CNRS, Centre d’Etudes de Saclay, Gif sur Yvette, France).Google Scholar
Rodríguez-Carvajal, J. (1997). Introduction to the Program FULLPROF (Laboratoire Léon Brillouin, CEA-CNRS, Saclay, France).Google Scholar
Sakata, M., and Cooper, M. J. (1979). “An analysis of the Rietveld profile refinement method,” J. Appl. Crystallogr. 12, 554563.CrossRefGoogle Scholar
Scott, H. G. (1983). “The estimation of standard deviations in powder diffraction Rietveld refinements,” J. Appl. Crystallogr. 16, 159163.CrossRefGoogle Scholar
Toraya, H. (1985). “The effects of profile-function truncation in X-ray powder-pattern fitting,” J. Appl. Crystallogr. 18, 351358.CrossRefGoogle Scholar
Toraya, H. (1990). “Array-type universal profile function for powder pattern fitting,” J. Appl. Crystallogr. 23, 485491.CrossRefGoogle Scholar
Wiles, D. B., Sakhtivel, A., and Young, R. A. (1988). Users Guide to Program DBW3.2Sfor Rietveld Analysis of X-ray and Neutron Powder Diffraction Patterns (Version 8804) (School of Physics, Georgia Institute of Technology, Atlanta, GA).Google Scholar
Wiles, D. B., and Young, R. A. (1981). “A new computer program for Rietveld analysis of X-ray powder diffraction patterns,” J. Appl. Crystallogr. 14, 149151.CrossRefGoogle Scholar
Young, R. A. (1980).Structural Analysis from X-ray Powder Diffraction Patterns with the Rietveld Method, in Accuracy in Powder Diffraction, edited by Block and Hubbard(NBS Spec. Publ. 567, 143163).Google Scholar
Young, R. A. (1993). The Rietveld Method, edited by R. A. Young, IUCr Monograph on Crystallography, 5 (IUCr/Oxford University Press, Oxford), pp. 1–38.Google Scholar
Young, R. A., Sakthivel, A., Moss, T. S., and Paiva-Santos, C. O. (1994). User's Guide to Program DBWS-9411 (School of Physics, Georgia Institute of Technology, Atlanta, GA).Google Scholar
Young, R. A., and Wiles, D. B. (1982). “Profile shape functions in Rietveld refinements,” J. Appl. Crystallogr. 15, 430438.CrossRefGoogle Scholar