International
Tables for
Crystallography
Volume C
Mathematical, physical and chemical tables
Edited by E. Prince

International Tables for Crystallography (2006). Vol. C. ch. 5.2, p. 493

Section 5.2.3.1. Aberrations

W. Parrish,a A. J. C. Wilsonb and J. I. Langfordc

a IBM Almaden Research Center, San Jose, CA, USA,bSt John's College, Cambridge CB2 1TP, England, and cSchool of Physics & Astronomy, University of Birmingham, Birmingham B15 2TT, England

5.2.3.1. Aberrations

| top | pdf |

The systematic errors are generally called aberrations; they differ from random errors in that in principle they can be calculated for any particular experimental arrangement and the observations corrected for them, leaving only the random error. In practice, of course, the calculation may be difficult. Aberrations can be divided broadly into two classes: (i) geometrical and (ii) physical. The geometrical aberrations are those that depend on the dimensions of the source, specimen, and detector (or of the slits that limit their effective dimensions). In angle-dispersive techniques, the physical aberrations depend on the intensity distribution in the range of wavelengths used, and in both angle- and energy-dispersive techniques they depend on the response characteristics of the detector and associated circuits.

The aberrations shift and distort the diffraction maxima. The study of their effects can be divided into four stages, corresponding to four levels of mathematical difficulty, and the stage to which it is necessary to carry the calculation depends on the purpose in view and the identifiable feature (Subsection 5.2.2.1[link]) of the wavelength distribution that it is intended to adopt as a measure of the position of the line profile. The three usual features are:

  • (i) the centroid (centre of gravity, mean, average) of the wavelength distribution;

  • (ii) the peak (mode, maximum); and

  • (iii) the best overall fit between the observed and the synthesized line profile.

The first of these, the centroid, requires only the first stage of the calculation for the geometrical aberrations and the first and second for the physical; the second, the peak, logically requires all four stages, but approximations can be obtained at the second stage; and the third, the best overall fit, requires all four stages.

The first stage of the calculation is the determination of the effect of the aberration on the centroid of the diffraction maximum, and ordinarily this gives rise to no insurmountable difficulty (Spencer, 1931[link], 1935[link], 1937[link], 1939[link], 1941[link], 1949[link]; Wilson, 1950[link]; Ladell, Parrish & Taylor, 1959[link]; Pike & Wilson, 1959[link]). It is all that is required for the correction of centroid positions for geometrical aberrations, which should be strictly additive. There is some limitation for physical aberrations (Edwards & Toman, 1970[link]; Wilson, 1970b[link]).

The second stage is the calculation of the mean-square broadening (variance). This can be used to obtain a reasonable approximation to the correction of peak positions over a wide range of Bragg angle (Wilson, 1961[link]; Gale, 1963[link], 1968[link]). To this approximation, the position of the observed peak is given by [(2\theta)_{\rm obs}=(2\theta)_{\rm true} + \langle\Delta(2\theta)\rangle + WI'''/2I'', \eqno (5.2.3.1)]where [\langle \Delta(2\theta)\rangle] is the centroid and W the variance of the geometrical aberrations and I′′ and I′′′ are second and third derivatives of the observed line profile evaluated at its maximum. The physical aberrations of the centroid depend on the variance of the part of the wavelength distribution used in determining the centroid (Wilson, 1958[link], 1963[link]; Wilson & Delf, 1961[link]). Those of the peak depend on the ratio of the peak intensity I to its second derivative I′′ (Wilson, 1961[link], 1963[link], 1965c[link]).

Often an aberration can be expressed in the form [\Delta(2\theta)= KF(2\theta), \eqno (5.2.3.2)]where the function F gives the angular variation of the aberration and K depends only on dimensions etc. that are fixed for a particular experiment but whose actual measurement is too difficult or tedious. The constant K can then be treated along with the lattice parameters as an adjustable parameter in least-squares refinement (analytical extrapolation; see Subsection 5.2.3.2[link]).

The third stage is the calculation of the line profile corresponding to each geometrical aberration. These aberration profiles can be combined by convolution (folding), either directly or by Fourier methods, and, in the fourth stage, the combined aberration profile can be convoluted with the emission profile of the X-ray source (or the emission profile as trimmed by a monochromator, pulse-height analyser, filter etc.) and with the diffraction profile corresponding to the state of strain, crystallite size, etc. of the specimen. This calculation of the composite line profile would be a necessary preliminary to an exact use of peak positions or of overall-profile fitting in lattice-parameter determination.

Such calculations were proposed many years ago (for example, by Alexander, 1948[link], 1950[link], 1953[link], 1954[link]), and have been used by Beu and co-workers (see Section 5.2.9[link]), and also by Boom and Smits (Boom & Smits, 1965[link]; Boom, 1966[link]). With the development of more powerful computer methods, such calculations can now be carried out routinely (e.g. Cheary & Coelho, 1992[link], 1994[link]; Kogan & Kupriyanov, 1992[link]; Timmers, Delhez, Tuinstra & Peerdeman, 1992[link]). However, not all the relevant instrumental parameters can in general be determined with sufficient accuracy and overall instrumental line profiles are normally obtained by means of a suitable standard material, for which sample broadening is negligible (Section 5.2.12[link]). There is, however, a related common approach, empirical rather than fundamental, based on the proposal of Rietveld (1967[link], 1969[link]). Its use in structure determination is treated in detail in Chapter 8.6[link] , and its use in lattice-parameter determination in Section 5.2.6[link]. There seems to be no detailed published study of the accuracy attainable for lattice parameters, but the estimated standard deviations quoted (see, for example, Young, 1988[link]) are comparable with those obtained for simpler structures giving resolved reflections.

References

First citation Alexander, L. (1948). Geometrical factors affecting the contours of X-ray spectrometer maxima. I. Factors causing asymmetry. J. Appl. Phys. 19, 1068–1071.Google Scholar
First citation Alexander, L. (1950). Geometrical factors affecting the contours of X-ray spectrometer maxima. II. Factors causing broadening. J. Appl. Phys. 21, 126–136.Google Scholar
First citation Alexander, L. (1953). The effect of vertical divergence on X-ray powder diffraction lines. Br. J. Appl. Phys. 4, 92–93.Google Scholar
First citation Alexander, L. (1954). The synthesis of X-ray spectrometer line profiles with application to crystallite size measurements. J. Appl. Phys. 25, 155–161.Google Scholar
First citation Boom, G. (1966). Accurate lattice parameters and the LPC method. Groningen: van Denderen.Google Scholar
First citation Boom, G. & Smits, D. W. (1965). Calculation of Debye–Scherrer diffraction line profiles and their applications in precision determination of lattice parameters. Proc. K. Ned. Akad. Wet. Ser. B, 68, 46–52.Google Scholar
First citation Cheary, R. W. & Coelho, A. (1992). A fundamental parameters approach to X-ray line-profile fitting. J. Appl. Cryst. 25, 109–121.Google Scholar
First citation Cheary, R. W. & Coelho, A. (1994). Synthesizing and fitting linear position-sensitive detector step-scanned line profiles. J. Appl. Cryst. 27, 673–681.Google Scholar
First citation Edwards, H. J. & Toman, K. (1970). The additivity of variances in powder diffraction profile analysis. J. Appl. Cryst. 3, 165–171.Google Scholar
First citation Gale, B. (1963). The positions of Debye diffraction line peaks. Br. J. Appl. Phys. 14, 357–364.Google Scholar
First citation Gale, B. (1968). The aberrations of a focusing X-ray diffraction instrument: second-order theory. Br. J. Appl. Phys. (J. Phys. D), 1, 393–408.Google Scholar
First citation Kogan, V. A. & Kupriyanov, M. F. (1992). X-ray powder diffraction line profiles by Fourier synthesis. J. Appl. Cryst. 25, 16–25.Google Scholar
First citation Ladell, J., Parrish, W. & Taylor, J. (1959). Interpretation of diffractometer line profiles. Acta Cryst. 12, 561–567.Google Scholar
First citation Pike, E. R. & Wilson, A. J. C. (1959). Counter diffractometer – the theory of the use of centroids of diffraction profiles for high accuracy in the measurement of diffraction angles. Br. J. Appl. Phys. 10, 57–68.Google Scholar
First citation Rietveld, H. M. (1967). Line profiles of neutron powder diffraction peaks for structure refinement. Acta Cryst. 22, 151–152.Google Scholar
First citation Rietveld, H. M. (1969). A profile refinement method for nuclear and magnetic structures. J. Appl. Cryst. 2, 65–71.Google Scholar
First citation Spencer, R. C. (1931). Additional theory of the double X-ray spectrometer. Phys. Rev. 38, 618–629.Google Scholar
First citation Spencer, R. C. (1935). The effect of the spectrometer on the width of spectral lines. Phys. Rev. 48, 473.Google Scholar
First citation Spencer, R. C. (1937). A theorem on the effect of vertical divergence. Phys. Rev. 52, 761.Google Scholar
First citation Spencer, R. C. (1939). The correction of experimental curves for the resolving power of the apparatus. Phys. Rev. 55, 239.Google Scholar
First citation Spencer, R. C. (1941). Optimum design of physical apparatus. Phys. Rev. 60, 172.Google Scholar
First citation Spencer, R. C. (1949). Discussion of ``Geometrical factors affecting X-ray spectrometer maxima''. J. Appl. Phys. 20, 413–414.Google Scholar
First citation Timmers, J., Delhez, R., Tuinstra, F. & Peerdeman, F. (1992). X-ray tracing; a tool for improved accuracy in powder diffractometry. Accuracy in powder diffraction II. NIST Spec. Publ. No. 846, edited by E. Prince & J. K. Stalick, p. 217.Google Scholar
First citation Wilson, A. J. C. (1950). Geiger-counter X-ray spectrometer – influence of the size and absorption coefficient of the specimen on position and shape of powder diffraction maxima. J. Sci. Instrum. 27, 321–325.Google Scholar
First citation Wilson, A. J. C. (1958). Effect of absorption on mean wavelength of X-ray emission lines. Proc. Phys. Soc. London, 72, 924–925.Google Scholar
First citation Wilson, A. J. C. (1961). A note on peak displacements in X-ray diffractometry. Proc. Phys. Soc. London, 78, 249–255.Google Scholar
First citation Wilson, A. J. C. (1963). Mathematical theory of X-ray powder diffractometry. Eindhoven: Centrex.Google Scholar
First citation Wilson, A. J. C. (1965c). Röntgenstrahlpulverdiffractometrie. Mathematische Theorie. Eindhoven: Centrex.Google Scholar
First citation Wilson, A. J. C. (1970b). Limitations on the additivity of moments in line-profile analysis. J. Appl. Phys. 3, 71–73.Google Scholar
First citation Wilson, A. J. C. & Delf, B. W. (1961). Effects of variations in the quantum counting efficiency of detectors on the mean wavelengths of emission lines. Proc. Phys. Soc. London, 78, 1256–1258.Google Scholar
First citation Young, R. A. (1988). Pressing the limits of Rietveld refinement. Aust. J. Phys. 41, 294–310.Google Scholar








































to end of page
to top of page