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. 505-508

Section 5.3.1. Introduction

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. Introduction

| top | pdf |

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.

5.3.1.2. Introduction to single-crystal methods

| top | pdf |

The essential feature of single-crystal methods is the necessity for the very accurate setting of the diffracting planes of the crystal in relation to the axes or planes of the instrument; this may be achieved manually or automatically. The fully automated four-circle diffractometer permits measurements to be made with an arbitrarily oriented single crystal, since its original position in relation to the axes of the device can be determined by means of a computer (calculation of the orientation matrix); the crystal can then be automatically displaced into each required diffracting position. Misalignment of the crystal and/or of element(s) of the device (the collimator, for example) may be a source of serious error (Burke & Tomkeieff, 1968[link], 1969[link]; Halliwell, 1970[link]; Walder & Burke, 1971[link]; Filscher & Unangst, 1980[link]; Larson, 1974[link]). On the other hand, a well defined single-crystal setting allows the unequivocal indexing of recorded reflections (moving-film methods, counter diffractometer) and the accurate determination of Bragg angles. In the particular case of large, specially cut and set, single crystals and a suitable measurement geometry, it is possible to avoid some sources of systematic error (Bond, 1960[link]).

Single-crystal methods are realized by a great variety of techniques.

  • (i) Usually, a single diffraction phenomenon, which occurs when only one set of planes is in position to diffract the incident X-ray beam at a given moment, is applied in lattice-parameter measurements. A separate group of so-called multiple-diffraction methods is formed by methods in which two or more sets of planes simultaneously fulfil the diffraction condition [equations (5.3.1.1)[link] and (5.3.1.2)[link]]. Some photographic divergent-beam methods (Lonsdale, 1947[link]; Heise, 1962[link]; Morris, 1968[link]; Isherwood & Wallace, 1971[link]; Lang & Pang, 1995[link]) and methods with counter recording in which a collimated beam is used (Renninger, 1937[link]; Post, 1975[link]) belong to this group.

  • (ii) In traditional methods, the unit-cell dimensions are determined in relation to the wavelength of a given X-radiation. Their accuracy and precision [equations (5.3.1.3)[link] and (5.3.1.4)[link]] are limited to those determined by the accuracy and precision of the wavelength determination and the width [w_\lambda] of the wavelength distribution [see equations (5.3.1.7)[link] and (5.3.1.9)[link]], so that the accuracy cannot exceed the limit of about 1 part in 106. To surmount these difficulties, new methods have been introduced. Combined X-ray and optical interferometry (applied simultaneously) permits the determination both of the lattice parameter and of the X-ray wavelength in terms of the visible wavelength standard so that interplanar distances can be `non-dispersively' (independently of the optical and X-ray wavelength values and their dispersions) measured in metric units with an accuracy of 1 part in 107. These methods (Deslattes, 1969[link]; Deslattes & Henins, 1973[link]; Becker et al., 1981[link]; see also Section 4.4.2[link] ) are so-called non-dispersive methods. Various pseudo-non-dispersive methods, in which the width of the diffraction profile has been reduced owing to the use of two- or three-crystal spectrometers (Godwod, Kowalczyk & Szmid, 1974[link]; Hart & Lloyd, 1975[link]; Buschert, 1965[link]) and/or multiple-beam techniques (Hart, 1969[link]; Larson, 1974[link]; Kishino, 1973[link]; Ando, Bailey & Hart, 1978[link]; Buschert, Pace, Inzaghi & Merlini, 1980[link]; Buschert, Meyer, Stuckey Kauffman & Gotwals, 1983[link]; Häusermann & Hart, 1990[link]), are a very suitable tool for high-precision (up to 1 part in 109) differential measurements. In multiple-diffraction methods, an interesting type of `n-crystal spectrometer' is generated within the specimen (Post, 1975[link]), so that the resulting diffraction profiles are also narrow.

  • (iii) In traditional methods, usually only one single crystal (the specimen) is used to determine a given lattice parameter, while, in the non-dispersive and pseudo-non-dispersive methods, two, three or more single crystals are required, which play the role of the monochromator and/or the reference crystal.

  • (iv) In the majority of methods, a single X-ray beam is used, which is usually well collimated. In some methods, however, in which the sample remains stationary, a highly divergent beam is applied to satisfy the Bragg law for various sets of planes. These are the Kossel (1936[link]) method and the divergent-beam techniques developed by Lonsdale (1947[link]). The original conception of multiple-beam measurement was introduced by Hart (1969[link]). The beams may come from two sources (as in the original paper) or may be separated from only one source (Kishino, 1973[link]).

  • (v) Most frequently, only one wavelength of characteristic X-radiation is used. Sometimes, a monochromator is applied – in particular, in two- or three-crystal spectrometers [see, for example, Fewster (1989[link]) and Obaidur (2002[link])]. White X-radiation may also be used in the Laue method, recently introduced for absolute measurements of the unit-cell dimensions (Carr, Cruickshank & Harding, 1992[link]), or in connection with the two-crystal spectrometer (Okazaki & Kawaminami (Okazaki & Kawaminami, 1973a[link], b[link]; Okazaki & Ohama, 1979[link]; Ohama, Sakashita & Okazaki, 1979[link]; Okazaki & Soejima, 2001[link]), to allow measurements at extremely large [\theta] angles, or in energy-dispersive diffractometers (Buras, Olsen, Gerward, Will & Hinze, 1977[link]), which make short exposures possible.

    In the second case, the measurement is based on a principle different from that of traditional methods: a continuous incident X-ray spectrum and a fixed Bragg angle [2\theta_0] are used. The relation between the interplanar spacing [d_H] and the energy [E_H] of the scattered photons is given by [E_H\,d_H\sin\theta= \textstyle{1\over2}hc=6.199\hbox{ (keV}\kern2pt{\rm \AA}), \eqno (5.3.1.10)]where H denotes hkl.

    The resolved [K\alpha_{1,2}] doublet is often used in various methods (photographic moving-crystal methods as well as divergent-beam diffractometers) to base the measurements on two independent constant values (Main & Woolfson, 1963[link]; Polcarová & Zůra, 1977[link]; Schwartzenberger, 1959[link]; Mackay, 1966[link]; Isherwood & Wallace, 1971[link]; Spooner & Wilson, 1973[link]; Heise, 1962[link]) or to obtain two independent X-ray beams from a single X-ray source (the multiple-beam method proposed by Kishino, 1973[link]). Sometimes, the Kβ line is also applied (Popović, 1971[link]; Kishino, 1973[link]).

  • (vi) More and more frequently, new sources of radiation are introduced instead of traditional laboratory Bremsstrahlung sources. In the case of methods using a divergent beam, the excitation of the characteristic X-rays may be performed both by primary X-rays (Lonsdale, 1947[link]) and by electron bombardment (Kossel, 1936[link]; Gielen, Yakowitz, Ganow & Ogilvie, 1965[link]; Ullrich & Schulze, 1972[link]), or by proton irradiation (Geist & Ascheron, 1984[link]). Also, a Mössbauer source (because of its short wavelength) (Bearden, Marzolf & Thomsen, 1968[link]) and synchrotron radiation (Buras et al., 1977[link]; Ando, Hagashi, Usuda, Yasuami & Kawata, 1989[link]) may be used (see also §5.3.3.9[link] below). The latter is considered to be an ideal X-ray source because of the short exposure required.

When making a choice of the method, the aim of the measurement, the required accuracy and/or precision as well as the laboratory equipment available should be taken into account.

  • (i) In the case when unit-cell parameters of a standard crystal are to be determined, the highest accuracy is needed. This special task is sometimes realized by unique, sophisticated and time-consuming methods (Baker, George, Bellamy & Causer, 1968[link]; Deslattes & Henins, 1973[link]; Becker, Seyfried & Siegert, 1982[link]; Härtwig & Grosswig, 1989[link]): the sample has to be of high quality (pure, defect-free, suitably prepared) and should have a small thermal-expansion coefficient. A detailed error analysis is required. Such measurement is often reduced to only one parameter, since high-symmetry crystals are generally used as standard.

  • (ii) The second special problem is to determine all the lattice parameters of a lower-symmetry system with high, though not necessarily the highest, accuracy.

    The most suitable tool for this task is the automated four-circle diffractometer, which permits a proper setting of the crystal for each possible hkl reflection. One crystal mounting is then sufficient to determine the unit cell with rather high (10 parts in 106) accuracy. The measurements can also be performed using a two-circle diffractometer (Clegg & Sheldrick, 1984[link]); these are, however, more troublesome since, in this case, only two rotations are motor-driven while the two remaining angles must be set by hand.

    When the diffractometer is to be used, preliminary measurements with photographic methods are advisable. A single rotation photograph allows one to determine one lattice parameter only, a single moving-film photograph makes it possible to obtain a two-dimensional picture of the reciprocal cell, and a suitable combination of two photographic techniques can be the basis for the determination of all the lattice parameters from a single mounting of the crystal (Buerger, 1942[link]; Hulme, 1966[link]; Hebert, 1978[link]; Wölfel, 1971[link]) with moderate accuracy (not better than 1 part in 104).

    The above counter and photographic methods are suitable for small, preferably spherical, crystals. In the case of one-crystal spectrometers, which give better accuracy (from 10 to 1 parts in 106), in particular when the Bond (1960[link]) arrangement is used and when all necessary corrections are taken into account, the preferable form of the sample is a large, flat slab, the surface of which is parallel to the planes of interest. Usually, several samples from one single crystal are needed in order to determine the unit cell of a lower-symmetry system (Cooper, 1962[link]). It is possible, however, to determine coplanar lattice parameters using a single sample, in one crystal mounting, when reflections from different crystal planes are taken into consideration (Luger, 1980[link], Section 4.2.2; Grosswig, Jäckel, Kittner, Dietrich & Schellenberger 1985[link]), and a special combination of reflection and transmission geometries enables one to determine all the lattice parameters from a single sample (Lisoivan, 1974[link], 1981[link], 1982[link]).

    Multiple-diffraction methods, both photographic and with counter recording, can provide a large number of reflections from one crystal mounting. In spite of this, these methods are as yet applied to mainly cubic lattices and exceptionally to other (orthogonal) lattices, because the interpretation of multiple-diffraction patterns is rather complicated (Chang, 1984[link]; and references therein).

    High-precision multiple-crystal spectrometers are suitable for comparison measurements in which differences in interplanar spacings rather than absolute values of the spacings are determined.

  • (iii) When the lattice-parameter changes caused by changes of environmental conditions are to be determined, high precision and sensitivity are more important than high accuracy.

    In the case of temperature dependence, an additional low- and/or high-temperature attachment is necessary for the precise establishment and control of the temperature (Baker, George, Bellamy & Causer, 1968[link]; Łukaszewicz, Kucharczyk, Malinowski & Pietraszko, 1978[link]; Okazaki & Ohama, 1979[link]; Okada, 1982[link]; Soejima, Tomonaga, Onitsuka & Okazaki, 1991[link]), so that the basic instrument should be relatively simple (Glazer, 1972[link]; Berger, 1984[link]; Clegg & Sheldrick, 1984[link]). Since measurements at many temperatures are then performed, the problem is to obtain the desired precision in as short a time as possible by using automatic control (Baker, George, Bellamy & Causer, 1968[link]), special strategy of measurement (Barns, 1972[link]; Urbanowicz, 1981a[link]), and special sources of radiation [synchrotron radiation used by Buras et al. (1977[link]) and Ando, Hagashi, Usuda, Yasuami & Kawata (1989[link])]. Analogous problems appear when the effects of other factors such as pressure (Mauer, Hubbard, Piermarini & Block, 1975[link]; d'Amour, Denner, Schulz & Cardona, 1982[link]; Leszczyński, Podlasin & Suski, 1993[link]), or electric field (Kobayashi, Yamada & Nakamura, 1963[link]) are examined.

  • (iv) To detect small differences of lattice parameter between the sample and the standard or between two points of the same specimen, the highest precision is required. To improve resolution in traditional methods, a finely collimated X-ray beam (Kobayashi, Yamada & Nakamura, 1963[link]) and cameras with large radius (Kobayashi, Yamada & Azumi, 1968[link]) are required. Really high precision, which reaches 1 part in 109, can be obtained with multiple-crystal (pseudo-non-dispersive) techniques (Hart, 1969[link]; Buschert, Meyer, Stuckey Kauffman & Gotwals, 1983[link]).

In the present review, all the methods are classified with respect to the measurement technique, in particular into photographic and counter-diffractometer techniques. Moreover, the methods will be described in approximately chronological sequence,1 i.e. from the earliest and simplest rotating-crystal method to the latest more-complex non-dispersive techniques, and at the same time from those of poor accuracy and precision to those attaining the highest precision and/or accuracy. In each of the methods realizing a given technique, first the absolute and then the relative methods will be described.

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 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 Ando, M., Bailey, D. & Hart, M. (1978). A simple Bragg-spacing comparator. Acta Cryst. A34, 484–489.Google Scholar
First citation Ando, M., Hagashi, Y., Usuda, K., Yasuami, S. & Kawata, H. (1989). A precision Bond method with SR. Rev. Sci. Instrum. 60, 2410–2413.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 Baker, T. W., George, J. D., Bellamy, B. A. & Causer, R. (1968). Fully automated high-precision X-ray diffraction. Adv. X-ray Anal. 11, 359–375.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 Bearden, J. A., Marzolf, J. G. & Thomsen, J. S. (1968). Crystal diffraction profiles for monochromatic radiation. Acta Cryst. A24, 295–301.Google Scholar
First citation Becker, P., Dorenwendt, K., Ebeling, G., Lauer, R., Lucas, W., Probst, R., Rademacher, H.-J., Reim, G., Seyfried, P. & Siegert, H. (1981). Absolute measurement of the (220) lattice plane spacing in silicon crystal. Phys. Rev. Lett. 46, 1540–1543.Google Scholar
First citation Becker, P., Seyfried, P. & Siegert, H. (1982). The lattice parameter of highly pure silicon single crystals. Z. Phys. B, 48, 17–21.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 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 Buras, B., Olsen, J. S., Gerward, L., Will, G. & Hinze, E. (1977). X-ray energy-dispersive diffractometry using synchrotron radiation. J. Appl. Cryst. 10, 431–438.Google Scholar
First citation Burke, J. & Tomkeieff, M. V. (1968). Specimen and beam tilt errors in Bond's method of lattice parameter determination. Acta Cryst. A24, 683–685.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 Buschert, R. C. (1965). X-ray lattice parameter and linewidth studies in silicon. Bull. Am. Phys. Soc. 10, 125.Google Scholar
First citation Buschert, R. C., Meyer, A. J., Stuckey Kauffman, D. & Gotwals, J. K. (1983). A double-source double-crystal X-ray spectrometer for high-sensitivity lattice-parameter difference measurements. J. Appl. Cryst. 16, 599–605.Google Scholar
First citation Buschert, R. C., Pace, S., Inzaghi, D. & Merlini, A. E. (1980). A high-sensitivity double-beam triple-crystal spectrometer for lattice parameter and topographic measurements. J. Appl. Cryst. 13, 207–210.Google Scholar
First citation Carr, P. D., Cruickshank, D. W. J. & Harding, M. M. (1992). The determination of unit-cell parameters from Laue diffraction patterns using their gnomonic projections. J. Appl. Cryst. 25, 294–308.Google Scholar
First citation Chang, S.-L. (1984). Multiple diffraction of X-rays in crystals, Chap. 7.2 in particular. Berlin: Springer-Verlag.Google Scholar
First citation Clegg, W. & Sheldrick, G. M. (1984). The refinement of unit cell parameters from two-circle diffractometer measurements. Z. Kristallogr. 167, 23–27.Google Scholar
First citation Cooper, A. S. (1962). Precise lattice constants of germanium, aluminium, gallium arsenide, uranium, sulphur, quartz and sapphire. Acta Cryst. 15, 578–582.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 Deslattes, R. D. (1969). Optical and X-ray interferometry of a silicon lattice spacing. Appl. Phys. Lett. 15, 386–388.Google Scholar
First citation Deslattes, R. D. & Henins, A. (1973). X-ray to visible wavelength ratios. Phys. Rev. Lett. 31, 972–975.Google Scholar
First citation Fewster, P. F. (1989). A high-resolution multiple-crystal multiple-reflection diffractometer. J. Appl. Cryst. 22, 64–69.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 Filscher, G. & Unangst, D. (1980). Bond-method for precision lattice constant determination. Dependence of lattice constant error on sample adjustment and collimator tilt. Krist. Tech. 15, 955–960.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 Geist, V. & Ascheron, C. (1984). The proton-induced Kossel effect and its application to crystallographic studies. Cryst. Res. Technol. 19, 1231–1244.Google Scholar
First citation Gielen, P., Yakowitz, H., Ganow, D. & Ogilvie, R. E. (1965). Evaluation of Kossel microdiffraction procedures: the cubic case. J. Appl. Phys. 36, 773–782.Google Scholar
First citation Glazer, A. M. (1972). A technique for the automatic recording of phase transitions in single crystals. J. Appl. Cryst. 5, 420–423.Google Scholar
First citation Godwod, K., Kowalczyk, R. & Szmid, Z. (1974). Application of a precise double X-ray spectrometer for accurate lattice parameter determination. Phys. Status Solidi A, 21, 227–234.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 Grosswig, S., Jäckel, K.-H., Kittner, R., Dietrich, B. & Schellenberger, U. (1985). Determination of the coplanar geometric lattice parameters of monocrystals with high precision. Cryst. Res. Technol. 20, 1093–1100.Google Scholar
First citation Halliwell, M. A. G. (1970). Measurement of specimen tilt and beam tilt in the Bond method. J. Appl. Cryst. 3, 418–419.Google Scholar
First citation Hart, M. (1969). High precision lattice parameter measurements by multiple Bragg reflexion diffractometry. Proc. R. Soc. London Ser. A, 309, 281–296.Google Scholar
First citation Hart, M. (1981). Bragg angle measurement and mapping. J. Cryst. Growth, 55, 409–427.Google Scholar
First citation Hart, M. & Lloyd, K. H. (1975). Measurement of strain and lattice parameter in epitaxic layers. J. Appl. Cryst. 8, 42–44.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 Häusermann, D. & Hart, M. (1990). A fast high-accuracy lattice-parameter comparator. J. Appl. Cryst. 23, 63–69.Google Scholar
First citation Hebert, H. (1978). A simple method for obtaining triclinic cell parameters from Weissenberg photographs from one crystal setting. Acta Cryst. A34, 946–949.Google Scholar
First citation Heise, H. (1962). Precision determination of the lattice constant by the Kossel line technique. J. Appl. Phys. 33, 938–941.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 Hulme, R. (1966). Triclinic cell parameters from one crystal setting. Acta Cryst. 21, 898–900.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 Kishino, S. (1973). Improved techniques of lattice parameter measurements using two X-ray beams. Adv. X-ray Anal. 16, 367–378.Google Scholar
First citation Kobayashi, J., Yamada, N. & Azumi, I. (1968). An X-ray method for accurate determination of lattice strain of crystals. Rev. Sci. Instrum. 39, 1647–1650.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 Kossel, W. (1936). Messungen am vollständigen Reflexsystem eines Kristallgitters. Ann. Phys. (Leipzig), 26, 533–553.Google Scholar
First citation Lang, A. R. & Pang, G. (1995). A possible new route to precise lattice-parameter measurement of perfect crystals using the divergent-X-ray-beam method. J. Appl. Cryst. 28, 61–64.Google Scholar
First citation Larson, B. C. (1974). High-precision measurements of lattice parameter changes in neutron-irradiated copper. J. Appl. Phys. 45, 514–518.Google Scholar
First citation Leszczyński, M., Podlasin, S. & Suski, T. (1993). A 109 Pa high-pressure cell for X-ray and optical measurements. J. Appl. Cryst. 26, 1–4.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 Lisoivan, V. I. (1981). Experimental refinement of the angles between unit-cell axes. (In Russian.) Kristallografiya, 26, 458–463.Google Scholar
First citation Lisoivan, V. I. (1982). Measurements of unit-cell parameters on one-crystal spectrometer. (In Russian.) Novosibirsk: Nauka.Google Scholar
First citation Lonsdale, K. (1947). Divergent-beam X-ray photography of crystals. Proc. R. Soc. London Ser. A, 240, 219–250.Google Scholar
First citation Luger, P. (1980). Modern X-ray analysis of single crystals. In particular, Chap. 4 and Section 4.2.2. Berlin: de Gruyter.Google Scholar
First citation Łukaszewicz, K., Kucharczyk, D., Malinowski, M. & Pietraszko, A. (1978). New model of the Bond diffractometer for precise determination of lattice parameters and thermal expansion of single crystals. Krist. Tech. 13, 561–567.Google Scholar
First citation Mackay, K. J. H. (1966). Proceedings of the IVth Congress on X-ray Optics and Microanalysis, pp. 544–554. Paris: Hermann.Google Scholar
First citation Main, P. & Woolfson, M. M. (1963). Accurate lattice parameters from Weissenberg photographs. Acta Cryst. 16, 731–733.Google Scholar
First citation Mauer, F. A., Hubbard, C. R., Piermarini, G. J. & Block, S. (1975). Measurement of anisotropic compressibilities by a single crystal diffractometer method. Adv. X-ray Anal. 18, 437–453.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 Morris, W. G. (1968). Crystal orientation and lattice parameters from Kossel lines. J. Appl. Phys. 39, 1813–1823.Google Scholar
First citation Obaidur, R. M. (2002). Energy-selective (+,+) monolithic monochromator and relative lattice-spacing measurement of Si wafers with synchrotron radiation. J. Synchrotron Rad. 9, 28–35.Google Scholar
First citation Ohama, N., Sakashita, H. & Okazaki, A. (1979). Improvement of high-angle double-crystal X-ray diffractometry (HADOX) for measuring temperature dependence of lattice constants. II. Practice. J. Appl. Cryst. 12, 455–459.Google Scholar
First citation Okada, Y. (1982). A high-temperature attachment for precise measurement of lattice parameters by Bond's method between room temperature and 1500 K. J. Phys. E, 15, 1060–1063.Google Scholar
First citation Okazaki, A. & Kawaminami, M. (1973a). Accurate measurement of lattice constants in a wide range of temperature: use of white X-ray and double-crystal diffractometry. Jpn. J. Appl. Phys. 12, 783–789.Google Scholar
First citation Okazaki, A. & Kawaminami, M. (1973b). Lattice constant of strontium titanate at low temperatures. Mater. Res. Bull. 8, 545–550.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 Okazaki, A. & Soejima, Y. (2001). Ultra-high-angle double-crystal X-ray diffractometry (U-HADOX) for determining a change in the lattice spacing: theory. Acta Cryst. A57, 708–712.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 Polcarová, M. & Zůra, J. (1977). A method for the determination of lattice parameters on single crystals. Czech. J. Phys. B27, 322–331.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 Post, B. (1975). Accurate lattice constants from multiple diffraction measurements. I. Geometry, techniques and systematic errors. J. Appl. Cryst. 8, 452–456.Google Scholar
First citation Renninger, M. (1937). `Umweganregung', eine bisher unbeachtete Wechselwirkungserscheinung bei Raumgitterinterferezen. Z. Phys. 106, 141–176.Google Scholar
First citation Schwartzenberger, D. R. (1959). Philos. Mag. 4, 1242–1246.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 Soejima, Y., Tomonaga, N., Onitsuka, H. & Okazaki, A. (1991). Two-dimensional intensity distribution in high-angle double-crystal X-ray diffractometry (HADOX). Z. Kristallogr. 195, 161–168.Google Scholar
First citation Spooner, F. J. & Wilson, C. G. (1973). The measurement of single-crystal lattice parameters using a double-diffraction technique. J. Appl. Cryst. 6, 132–135.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 Ullrich, H.-J. & Schulze, G. E. R. (1972). Röntgenographische Mikrobeugungsuntersuchungen an kristallinen Festkörpern mittels Gitterquelleninterferenzen (Kossel-Linien) und Weitwinkelinterferenzen (Pseudo-Kossel-Linien). Krist. Tech. 7, 207–220.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 Walder, V. & Burke, J. (1971). The elimination of specimen and beam tilt errors in the Bond method of precision lattice parameter determinations. J. Appl. Cryst. 4, 337–339.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
First citation Wölfel, E. R. (1971). A new film instrument for the exploration of reciprocal space. J. Appl. Cryst. 4, 297–302.Google Scholar








































to end of page
to top of page