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. 521-522

Section 5.3.3.4.2. Development of methods based on an asymmetric arrangement and their applications

E. Gałdeckaa

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

5.3.3.4.2. Development of methods based on an asymmetric arrangement and their applications

| top | pdf |

Although the Bond (1960[link]) method, based on a symmetric arrangement presented in §5.3.3.4.3[link], makes possible higher accuracy than that obtained by means of a standard diffractometer, an asymmetric arrangement proves to be more suitable for certain tasks connected with lattice-parameter measurement, because of its greater simplicity. The more detailed arguments for the use of such a device result from some disadvantages of the Bond method, discussed in §5.3.3.4.3.4[link].

One of the earliest and most often cited methods of lattice-parameter determination by means of the counter single-crystal diffractometer (in an asymmetric arrangement) is that of Smakula & Kalnajs (1955[link]). The authors reported unit-cell determinations of eight cubic crystals. The systematic errors due to seven factors were analysed according to the formulae derived by Wilson (1950[link]) and Eastabrook (1952[link]) for powder samples, and valid also for single crystals. The lattice parameters computed for various diffraction angles were plotted versus [\cos^2\theta]; extrapolation to [2\theta=180]° gave the lattice parameters corrected for systematic errors. Accuracy of 4 parts in 105, limited by the uncertainty of the X-ray wavelength, and precision of 1 part in 106 were achieved.

A more complete list of factors causing broadening and asymmetry of the diffraction profile, and so affecting statistical and systematic errors of lattice-parameter determination, has been given by Kheiker & Zevin (1963[link], Tables IV, IVa, and IVb). Since the systematic errors due to the factors causing asymmetry (specimen transparency, axial divergence, flat specimen) are, as a rule, dependent on the Bragg angle and proportional to [\cos\theta], [\cos^2\theta], [\cot\theta] or [\cot^2\theta], they can be removed or reduced – as in the method of Smakula & Kalnajs (1955[link]) – by means of extrapolation to [\theta=90]°. The problem has also been discussed by Wilson (1963[link], 1980[link]) in the case of powder diffractometry [cf. §5.3.3.3.1(i)[link]]. When comparing the considerations of Kheiker & Zevin and Wilson [the list of references concerning the subject given by Kheiker & Zevin (1963[link]) is, with few exceptions, contained in that given by Wilson (1963[link])], it will be noticed that some differences in the formulae result from differences in the geometry of the measurement rather than from the different nature of the samples (single crystal, powder).

As in the photographic methods, the accurate recording of the angular separation between [K\alpha] and [K\beta] diffraction lines can be the basis for lattice-parameter measurements with a diffractometer (Popović, 1971[link]). The method allows one to reduce the error in the zero setting of the [2\theta] scale and the error due to incorrect positioning of the sample on the diffractometer, since the angular separations are independent of the zero positions of the [2\theta] and [\omega] scales.

An example of a contemporary method of lattice-parameter determination is given by Berger (1984[link]). As has been mentioned in §5.3.3.4.1[link], the characteristic feature of the device is the Soller slits, which limit the divergence of both primary and diffraction beams and, at the same time, eliminate errors due to eccentricity and absorption. On the other hand, systematic errors due to refraction, vertical inclination, vertical divergence, and Soller-slit inaccuracy, as well as asymmetry of profiles and crystal imperfection, have to be analysed.

Since, in this case, the angle between the incident and the reflected beam is measured, the inclinations of both beams must be considered. As a result of the analysis [analogous to that of Burke & Tomkeieff, 1969[link]; referred to in §5.3.3.4.3.2(4)[link]], the following expression for the angular correction [\Delta\theta_t] (to be added to the measured value of [\theta]) is obtained: [\Delta\theta_t={\alpha\gamma \over2\sin2\theta} + {\alpha^2+\gamma^2 \over 4\tan 2\theta}, \eqno (5.3.3.21)]where α and γ are the vertical inclinations of the incident and reflected beams, respectively. The correction for vertical divergence is presented in §5.3.3.4.3.2(3)[link].

The Soller-slit method, the accuracy and precision of which are comparable to those obtained with the Bond method, is suitable both for imperfect crystals, since only a single diffracting position of the sample is required, and for perfect samples, when an exactly defined irradiated area is required. It is applicable to absolute and to relative measurements. Examples are given by Berger, Rosner & Schikora (1989[link]), who worked out a method of absolute lattice-parameter determination of superlattices; by Berger, Lehmann & Schenk (1985[link]), who determined lattice-parameter variations in PbTe single crystals; and by Berger (1993[link]), who examined point defects in II–VI compounds.

An original method, based on determining the Bragg angle from a two-dimensional map of the intensity distribution (around the reciprocal-lattice point) of high-angle reflections as a function of angular positions of both the specimen and the counter, was described by Kobayashi, Yamada & Nakamura (1963[link]) and Kobayashi, Mizutani & Schmidt (1970[link]). A finely collimated X-ray beam, with a half-width less than 3′, was used for this purpose. The accuracy of the counter setting was ±0.1°, the scanning step [\Delta\theta=0.01]°. Systematic errors depending on the depth of penetration and eccentricity of the specimen were reported, and were corrected both experimentally (manifold measurements of the same planes for different diffraction ranges, and rotation of the crystal around its axis by 180°) and by means of extrapolation. The correction for refraction was introduced separately. The method was used in studies of the antiparallel 180° domains in the ferroelectric barium titanate, which were combined with optical studies.

The determination of variations in the cell parameter of GaAs as a function of homogeneity, effects of heat treatments, and surface defects has been presented by Pierron & McNeely (1969[link]). Using a conventional diffractometer, they obtained a precision of 3 parts in 106 and an accuracy better than 2 parts in 106. The systematic errors were removed both by means of suitable corrections (Lorentz–polarization factor and refraction) and by extrapolation.

A study of the thermal expansion of α-LiIO3 over a wide range of temperatures (between 20 and 520 K) in the vicinity of the phase transition has been reported by Abrahams et al. (1983[link]). Lattice-parameter changes were examined by means of a standard diffractometer (CAD-4); absolute values at separate points were measured by the use of a Bond-system diffractometer.

An apparatus for the measurement of uniaxial stress based on a four-circle diffractometer has been presented by d'Amour et al. (1982[link]). The stress, produced by turning a differential screw, can be measured in situ, i.e. without removing the apparatus from the diffractometer. An example of lattice-parameter measurement of Si stressed along [111] is given, in which the stress parameter ζ is calculated from intensity changes of the chosen 600 reflection.

References

First citation Abrahams, S. C., Liminga, R., Marsh, P., Schrey, F., Albertsson, J., Svensson, C. & Kvick, Å. (1983). Thermal expansivity of α-LiIO3 between 20 and 520 K. J. Appl. Cryst. 16, 453–457.Google Scholar
First citation d'Amour, H., Denner, W., Schulz, H. & Cardona, M. (1982). A uniaxial stress apparatus for single-crystal X-ray diffraction on a four-circle diffractometer: Application to silicon and diamond. J. Appl. Cryst. 15, 148–153.Google Scholar
First citation Berger, H. (1984). A method for precision lattice-parameter measurement of single crystals. J. Appl. Cryst. 17, 451–455.Google Scholar
First citation Berger, H. (1993). X-ray diffraction studies on point defects in II–VI compounds. Cryst. Res. Technol. 28, 795–801.Google Scholar
First citation Berger, H., Lehmann, A. & Schenk, M. (1985). Lattice parameter variations in PbTe single crystals. Cryst. Res. Technol. 20, 579–581.Google Scholar
First citation Berger, H., Rosner, B. & Schikora, D. (1989). Lattice parameter determination of superlattices. Cryst. Res. Technol. 24, 437–441.Google Scholar
First citation Bond, W. L. (1960). Precision lattice constant determination. Acta Cryst. 13, 814–818.Google Scholar
First citation Burke, J. & Tomkeieff, M. V. (1969). Errors in the Bond method of lattice parameter determinations. Further considerations. J. Appl. Cryst. 2, 247–248.Google Scholar
First citation Eastabrook, J. N. (1952). Effect of vertical divergence on the displacement and breadth of X-ray powder diffraction lines. Br. J. Appl. Phys. 3, 349–352.Google Scholar
First citation Kheiker, D. M. & Zevin, L. S. (1963). Rentgenowskaya diffraktometriya (X-ray diffractometry), Chap. 4. Moscow: Fizmatgiz.Google Scholar
First citation Kobayashi, J., Mizutani, I. & Schmidt, H. (1970). X-ray study on the lattice strains of ferroelectric iron iodine boracite Fe3B7O13. Phys. Rev. B, 1, 3801–3808.Google Scholar
First citation Kobayashi, J., Yamada, N. & Nakamura, T. (1963). Origin of the visibility of the antiparallel 180° domains in barium titanate. Phys. Rev. Lett. 11, 410–414.Google Scholar
First citation Pierron, E. D. & McNeely, J. B. (1969). Precise cell parameters of semiconductor crystals and their applications. Adv. X-ray Anal. 12, 343–353.Google Scholar
First citation Popović, S. (1971). An X-ray diffraction method for lattice parameter measurements from corresponding Kα and Kβ reflexions. J. Appl. Cryst. 4, 240–241.Google Scholar
First citation Smakula, A. & Kalnajs, J. (1955). Precision determination of lattice constants with a Geiger-counter X-ray diffractometer. Phys. Rev. 99, 1737–1743.Google Scholar
First citation Wilson, A. J. C. (1950). Geiger-counter X-ray spectrometer – influence of size and absorption coefficient of specimen on position and shape of powder diffraction maxima. J. Sci. Instrum. 27, 321–325.Google Scholar
First citation Wilson, A. J. C. (1963). Mathematical theory of X-ray powder diffractometry. Philips Technical Library. Eindhoven: Centrex Publishing Company.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








































to end of page
to top of page