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

International Tables for Crystallography (2006). Vol. C. ch. 5.3, pp. 505-506

Section 5.3.1.1. General remarks

E. Gałdeckaa

a Institute of Low Temperature and Structure Research, Polish Academy of Sciences, PO Box 937, 50-950 Wrocław 2, Poland

5.3.1.1. General remarks

| top | pdf |

The starting point for lattice-parameter measurements by X-ray diffraction methods and evaluation of their accuracy and precision is the Bragg law, combining diffraction conditions (the Bragg angle [\theta] and the wavelength λ) with the parameters of the lattice to be determined: [2 d\sin\theta=n\lambda, \eqno (5.3.1.1)]in which d is the interplanar spacing, being a function of direct-lattice parameters a, b, c, α, β, γ, and n is the order of interference. Before calculating the lattice parameters, corrections for refraction should be introduced to d values determined from (5.3.1.1)[link] [James (1967[link]); Isherwood & Wallace (1971[link]); Lisoivan (1974[link]); Hart (1981[link]); Hart, Parrish, Bellotto & Lim (1988[link]); cf. §5.3.3.4.3.2[link], paragraph (2) below].

Since only d values result directly from (5.3.1.1)[link] and the non-linear dependence of direct-lattice parameters on d is, in a general case, rather complicated (see, for example, Buerger, 1942[link], p. 103), it is convenient to introduce reciprocal-cell parameters ([a^*], [b^*], [c^*], [\alpha^*], [\beta^*], [\gamma^*]) and to write the Bragg law in the form: [\eqalignno{ 4\sin^2\theta/\lambda^2 &=n^2/d^2 \cr &=h^2a^{*2}+k^2b^{*2}+l^2c^{*2}+2hka^*b^*\cos\gamma^* \cr &\quad +2hla^*c^*\cos\beta^*+2klb^*c^*\cos\alpha^*, &(5.3.1.2)}]where h, k and l are the indices of reflection, and then those of the direct cell are calculated from suitable equations given elsewhere (Buerger, 1942[link], p. 361, Table 2). The minimum number of equations, and therefore number of measurements, necessary to obtain all the lattice parameters is equal to the number of parameters, but many more measurements are usually made, to make possible least-squares refinement to diminish the statistical error of the estimates. In some methods, extrapolation of the results is used to remove the [\theta]-dependent systematic errors (Wilson, 1980[link], Section 5, and references therein) and requires several measurements for various [\theta].

Measurements of lattice dimensions can be divided into absolute, in which lattice dimensions are determined under defined environmental conditions, and relative, in which, compared to a reference crystal, small changes of lattice parameters (resulting from changes of temperature, pressure, electric field, mechanical stress etc.) or differences in the cell dimensions of a given specimen (influenced by point defects, deviation from exact stoichiometry, irradiation damage or other factors) are examined.

In the particular case when the lattice parameter of the reference crystal has been very accurately determined, precise determination of the ratio of two lattice parameters enables one to obtain an accurate value of the specimen parameter (Baker & Hart, 1975[link]; Windisch & Becker, 1990[link]; Bowen & Tanner, 1995[link]).

Absolute methods can be characterized by the accuracy δd, defined as the difference between measured and real (unknown) interplanar spacings or, more frequently, by using the relative accuracy δd/d, defined by the formula obtained as a result of differentiation of the Bragg law [equation (5.3.1.1)[link]]: [\delta d/d=\delta \lambda/\lambda-\cot\theta\,\delta\theta, \eqno (5.3.1.3)]where δλ/λ is the relative accuracy of the wavelength determination in relation to the commonly accepted wavelength standard, and [\delta\theta] is the error in the Bragg angle determined.

The analogous criterion used for characterization of relative methods may be the precision, defined by the variance [\sigma^2(d)] [or its square root – the standard deviation, [\sigma(d)]] of the measured interplanar spacing d as the measure of repeatability of experimental results.

The relative precision of lattice-spacing determination can be presented in the form: [\sigma(d)/d=\cot\theta \sigma(\theta), \eqno (5.3.1.4)]where [\sigma(\theta)] is the standard deviation of the measured Bragg angle [\theta].

Another mathematical criterion proposed especially for relative methods is the sensitivity, defined (Okazaki & Ohama, 1979[link]) as the ratio [\delta\theta/\delta d], i.e. the change in the [\theta] value owing to the unit change in d.

The main task in unit-cell determination is the measurement of the Bragg angle. For a given [\theta] angle, the accuracy [\delta\theta] and precision [\sigma(\theta)] affect those of the lattice parameter [equations (5.3.1.3)[link] and (5.3.1.4)[link]]. To achieve the desired value of [\delta d/d], the accuracy [\delta \theta] must be no worse than resulting from the Bragg law (Bond, 1960[link]): [|\delta\theta|=(|\delta d|/d)\tan\theta. \eqno (5.3.1.5)]An analogous equation can be obtained for [\sigma(\theta)] as a function of [\sigma(d)/d]. The values [\delta\theta] and [\sigma(\theta)] depend not only on the measurement technique (X-ray source, device, geometry) and the crystal (its structure, perfection, shape, physical properties), but also on the processing of the experimental data.

The first two factors affect the measured profile [which will be denoted here – apart from the means of recording – by [h(\theta)]], being a convolution of several distributions (Alexander, 1948[link], 1950[link], 1954[link]; Alexander & Smith, 1962[link]; Härtwig & Grosswig, 1989[link]; Härtwig, Hölzer, Förster, Goetz, Wokulska & Wolf, 1994[link]) and the third permits calculations of the Bragg angle and the lattice parameters with an accuracy and a precision as high as possible in given conditions, i.e. for a given profile [h(\theta)].

In the general case, [h(\theta)] can be described as a convolution: [h(\theta)=h_\lambda(\theta)*h_A(\theta)*h_C(\theta), \eqno (5.3.1.6)]where [h_\lambda(\theta)] is an original profile due to wavelength distribution; [h_A(\theta)] is a distribution depending on various apparatus factors, such as tube-focus emissivity, collimator parameters, detector aperture; and [h_C(\theta)] is a function (the crystal profile) depending on the crystal, its perfection, mosaic structure, shape (flatness), and absorption coefficient.

The functions [h_A(\theta)] and [h_C(\theta)] are again convolutions of appropriate factors.

Each of the functions [h_\lambda(\theta)], [h_A(\theta)], and [h_C(\theta)] has its own shape and a finite width, which affect the shape and the width [\omega_h] of the resulting profile [h(\theta)].

Since [h_A(\theta)] and [h_C(\theta)] are ordinarily asymmetric, the profile [h(\theta)] is also asymmetric and may be considerably shifted in relation to the original one, [h_\lambda(\theta)], leading to systematic errors in lattice-parameter determination.

The finite precision [\sigma(d)/d], on the other hand, results from the fact that the two measured variables – the intensity h and the angle [\theta] – are random variables.

The half-width [\omega_\lambda] of [h_\lambda(\theta)] defines the minimum half-width of [h(\theta)] that it is possible to achieve with a given X-ray source: [\omega_h\gtrsim \omega_\lambda. \eqno (5.3.1.7)]It can be assumed from the Bragg law that: [\omega_\lambda=(w_\lambda/\lambda)\tan\theta, \eqno (5.3.1.8)]where [w_\lambda] is the half-width of the wavelength distribution. In commonly used X-ray sources, [w_\lambda/\lambda\approx 300\times10^{-6}].

Combination of (5.3.1.5)[link] and (5.3.1.8)[link], and with (5.3.1.7)[link] taken into consideration, gives an estimate of the ratio of the admissible error [\delta\theta] to the half-width of the measured profile: [{|\delta\theta|\over \omega_h}\le{|\delta d|/d\over w_\lambda/\lambda}. \eqno (5.3.1.9)]To obtain the highest possible accuracy and precision for a given experiment (given diffraction profile), mathematical methods of data analysis and processing and programming of the experiment are used (Bačkovský, 1965[link]; Wilson, 1965[link], 1967[link], 1968[link], 1969[link]; Barns, 1972[link]; Thomsen & Yap, 1968[link]; Segmüller, 1970[link]; Thomsen, 1974[link]; Urbanowicz, 1981a[link]; Grosswig, Jäckel & Kittner, 1986[link]; Gałdecka, 1993a[link], b[link]; Mendelssohn & Milledge, 1999[link]).

Measurements of lattice parameters can be realized both with powder samples and with single crystals. At the first stage of the development of X-ray diffraction methods, the highest precision was obtained with powder samples, which were easier to obtain and set, rather than with single crystals. The latter were considered to be more suitable in the case of lower-symmetry systems only. In the last 35 years, many single-crystal methods have been developed that allow the achievement of very high precision and accuracy and, at the same time, allow the investigation of different specific features characterizing single crystals only (defects and strains of a single-crystal sample, epitaxic layers).

Some elements are common to both powder and single-crystal methods: the application of the basic equations (5.3.1.1)[link] and (5.3.1.2)[link]; the use of the same formulae defining the precision and the accuracy [equations (5.3.1.4)[link] and (5.3.1.3)[link]] and – as a consequence – the tendency to use [\theta] values as large as possible; the means of evaluation of some systematic errors due to photographic cameras or to counter diffractometers (Parrish & Wilson, 1959[link]; Beu, 1967[link]; Wilson, 1980[link]); the methods of estimating statistical errors based on the analysis of the diffraction profile and some methods of increasing the accuracy (Straumanis & Ieviņš, 1940[link]). In other aspects, powder and single-crystal methods have developed separately, though some present-day high-resolution methods are not restricted to a particular crystalline form (Fewster & Andrew, 1995[link]). In special cases, the combination of X-ray powder diffraction and single-crystal Laue photography, reported by Davis & Johnson (1984[link]), can be useful for the determination of the unit-cell parameters.

Small but remarkable differences in lattice parameters determined by powder and single-crystal methods have been observed (Straumanis, Borgeaud & James, 1961[link]; Hubbard, Swanson & Mauer, 1975[link]; Wilson, 1980[link], Sections 6 and 7), which may result from imperfections introduced in the process of powdering or from uncorrected systematic errors (due to refractive-index correction, for example; cf. Hart, Parrish, Bellotto & Lim, 1988[link]). The first case was studied by Gamarnik (1990[link]) – both theoretically and experimentally. As shown by the author, the relative increase of lattice parameters in ultradispersed crystals of diamond in comparison with massive crystals was as high as [\Delta d/d=2.05\times10^{-3}\pm10^{-4}]. Analysis of results of lattice-parameter measurements of silicon single crystals and powders, performed by different authors (Fewster & Andrew, 1995[link], p. 455, Table 1), may lead to the opposite conclusion: The weighted-mean lattice parameter of silicon powder proved to be about 0.0002 Å smaller [(\Delta d/d\simeq -4\times 10^{-5})] than that of the bulk silicon.

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. (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 Alexander, L. E. & Smith, G. (1962). Single-crystal intensity measurements with the three-circle counter diffractometer. Acta Cryst. 15, 983–1004.Google Scholar
First citation Bačkovský, J. (1965). On the most accurate measurements of the wavelengths of X-ray spectral lines. Czech. J. Phys. B15, 752–759.Google Scholar
First citation Baker, J. F. C. & Hart, M. (1975). An absolute measurement of the lattice parameter of germanium using multiple-beam X-ray diffractometry. Acta Cryst. A31, 364–367.Google Scholar
First citation Barns, R. L. (1972). A strategy for rapid and accurate (p.p.m.) measurement of lattice parameters of single crystals by Bond's method. Adv. X-ray Anal. 15, 330–338.Google Scholar
First citation Beu, K. E. (1967). The precise and accurate determination of lattice parameters. Handbook of X-rays, edited by E. F. Kaelble, Chap. 10. New York: McGraw-Hill.Google Scholar
First citation Bond, W. L. (1960). Precision lattice constant determination. Acta Cryst. 13, 814–818.Google Scholar
First citation Bowen, D. K. & Tanner, B. K. (1995). A method for the accurate comparison of lattice parameters. J. Appl. Cryst. 28, 753–760.Google Scholar
First citation Buerger, M. J. (1942). X-ray crystallography. London: John Wiley.Google Scholar
First citation Davis, B. L. & Johnson, L. R. (1984). The true unit cell of ammonium hydrogen sulfate, (NH4)3H(SO4)2. J. Appl. Cryst. 17, 331–333.Google Scholar
First citation Fewster, P. F. & Andrew, N. L. (1995). Absolute lattice-parameter measurement. J. Appl. Cryst. 28, 451–458.Google Scholar
First citation Gałdecka, E. (1993a). Description and peak-position determination of a single X-ray diffraction profile for high-accuracy lattice-parameter measurements by the Bond method. I. An analysis of descriptions available. Acta Cryst. A49, 106–115.Google Scholar
First citation Gałdecka, E. (1993b). Description and peak-position determination of a single X-ray diffraction profile for high-accuracy lattice-parameter measurements by the Bond method. II. Testing and choice of description. Acta Cryst. A49, 116–126.Google Scholar
First citation Gamarnik, M. Ya. (1990). Size changes of lattice parameters in ultradisperse diamond and silicon. Phys. Status Solidi B, 161, 457–462.Google Scholar
First citation Grosswig, S., Jäckel, K.-H. & Kittner, R. (1986). Peak position determination of X-ray diffraction profiles in precision lattice parameter measurements according to the Bond-method with help of the polynomial approximation. Cryst. Res. Technol. 21, 133–139.Google Scholar
First citation Hart, M. (1981). Bragg angle measurement and mapping. J. Cryst. Growth, 55, 409–427.Google Scholar
First citation Hart, M., Parrish, W., Bellotto, M. & Lim, G. S. (1988). The refractive-index correction in powder diffraction. Acta Cryst. A44, 193–197.Google Scholar
First citation Härtwig, J. & Grosswig, S. (1989). Measurement of X-ray diffraction angles of perfect monocrystals with high accuracy using a single crystal diffractometer. Phys. Status Solidi A, 115, 369–382.Google Scholar
First citation Härtwig, J., Hölzer, G., Förster, E., Goetz, K., Wokulska, K. & Wolf, J. (1994). Remeasurement of characteristic X-ray emission lines and their application to line profile analysis and lattice parameter determination. Phys. Status Solidi A, 143, 23–34.Google Scholar
First citation Hubbard, C. R., Swanson, H. E. & Mauer, F. A. (1975). A silicon powder diffraction standard reference material. J. Appl. Cryst. 8, 45–48.Google Scholar
First citation Isherwood, B. J. & Wallace, C. A. (1971). The geometry of X-ray multiple diffraction in crystals. Acta Cryst. A27, 119–130.Google Scholar
First citation James, R. W. (1967). The optical principles of the diffraction of X-rays. London: Bell.Google Scholar
First citation Lisoivan, V. I. (1974). Local determination of all the lattice parameters of single crystals. (In Russian.) Appar. Methody Rentgenovskogo Anal. 14, 151–157.Google Scholar
First citation Mendelssohn, M. J. & Milledge, H. J. (1999). Divergent-beam technique used in a SEM to measure the cell parameters of isotopically distinct samples of LiF over the temperature range ~15–375 K. Acta Cryst. A55, 204–211.Google Scholar
First citation Okazaki, A. & Ohama, N. (1979). Improvement of high-angle double-crystal X-ray diffractometry (HADOX) for measuring temperature dependence of lattice constants. I. Theory. J. Appl. Cryst. 12, 450–454.Google Scholar
First citation Parrish, W. & Wilson, A. J. C. (1959). Precision measurements of lattice parameters of polycrystalline specimens. International tables for X-ray crystallography, Vol. II, Chap. 4.7, pp. 216–234. Birmingham: Kynoch Press.Google Scholar
First citation Segmüller, A. (1970). Automated lattice parameter determination on single crystals. Adv. X-ray Anal. 13, 455–467.Google Scholar
First citation Straumanis, M. E., Borgeaud, P. & James, W. J. (1961). Perfection of the lattice of dislocation-free silicon, studied by the lattice-constant and density method. J. Appl. Phys. 32, 1382–1384.Google Scholar
First citation Straumanis, M. & Ieviņš, A. (1940). Die Präzizionsbestimmung von Gitterkonstanten nach der asymmetrischen Methode. Berlin: Springer. [Reprinted by Edwards Brothers Inc., Ann Arbor, Michigan (1948).]Google Scholar
First citation Thomsen, J. S. (1974). High-precision X-ray spectroscopy. X-ray spectroscopy, edited by L. V. Azároff, pp. 26–132. New York: McGraw-Hill.Google Scholar
First citation Thomsen, J. S. & Yap, Y. (1968). Effect of statistical counting errors on wavelengths criteria for X-ray spectra. J. Res. Natl Bur. Stand. Sect. A, 72, 187–205.Google Scholar
First citation Urbanowicz, E. (1981a). The influence of in-plane collimation on the precision and accuracy of lattice-constant determination by the Bond method. I. A mathematical model. Statistical errors. Acta Cryst. A37, 364–368.Google Scholar
First citation Wilson, A. J. C. (1965). The location of peaks. Br. J. Appl. Phys. 16, 665–674.Google Scholar
First citation Wilson, A. J. C. (1967). Statistical variance of line-profile parameters. Measures of intensity, location and dispersion. Acta Cryst. 23, 888–898.Google Scholar
First citation Wilson, A. J. C. (1968). Statistical variance of line-profile parameters. Measures of intensity, location and dispersion: Corrigenda. Acta Cryst. A24, 478.Google Scholar
First citation Wilson, A. J. C. (1969). Statistical variance of line-profile parameters. Measures of intensity, location and dispersion: Addendum. Acta Cryst. A25, 584–585.Google Scholar
First citation Wilson, A. J. C. (1980). Accuracy in methods of lattice-parameter measurement. Natl Bur. Stand. (US) Spec. Publ. No. 567. Proceedings of Symposium on Accuracy in Powder Diffraction, NBS, Gaithersburg, MD, USA, 11–15 June 1979.Google Scholar
First citation Windisch, D. & Becker, P. (1990). Silicon lattice parameters as an absolute scale of length for high precision measurements of fundamental constants. Phys. Status Solidi A, 118, 379–388.Google Scholar








































to end of page
to top of page