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

International Tables for Crystallography (2006). Vol. C. ch. 2.3, pp. 60-61

Section 2.3.3.1.1. Preferred orientation

W. Parrisha and J. I. Langfordb

a IBM Almaden Research Center, San Jose, CA, USA, and bSchool of Physics & Astronomy, University of Birmingham, Birmingham B15 2TT, England

2.3.3.1.1. Preferred orientation

| top | pdf |

Preferred orientation changes the relative intensities from those obtained with a randomly oriented powder sample. It occurs in materials that have good cleavage or a morphology that is platy, acicular or any special shape in which the particles tend to orient themselves in specimen preparation. The micas and clay minerals are examples of materials that exhibit very strong preferred orientation. When they are prepared as reflection specimens, the basal reflections dominate the pattern. It is common in prepared thin films where preferred orientation occurs frequently or may be purposely induced to enhance certain optical, electrical, or magnetic properties for electronic devices. By comparison of the relative intensities with the random powder pattern, the degree of preferred orientation can be observed.

Powder reflections take place from crystallites oriented in different ways in the instrument geometries as shown in Fig. 2.3.1.2[link]. In reflection specimen geometry with θ–2θ scanning, reflections can occur only from lattice planes parallel to the surface and in the transmission mode they must be normal to the surface. In the Seemann–Bohlin and fixed specimen with 2θ scanning methods, the orientation varies from parallel to about 45° inclination to the surface. The effect of preferred orientation can be seen in diffraction patterns obtained by using the same specimen in the different geometries.

The effect is illustrated in Fig. 2.3.3.1[link] for m-chlorobenzoic acid, C7H5ClO2, with reflection and transmission patterns and the pattern calculated from the crystal structure. The degree of preferred orientation is shown by comparing the peak intensities of four reflections in the three patterns:[\halign{&\quad #\hfil&\quad #\hfil&\quad #\hfil&\quad #\hfil\cr\hbox{({\it hkl})} &(120) &(200) &(040) &(121)\cr \hbox{Reflection} &\hfil 9.8 &\hfil 0.6 &\hfil 1.6 &\hfil 2.5\cr \hbox{Transmission} &\hfil 5.2 &\hfil 0.5 &\hfil 0.7 &\hfil 9.3\cr \hbox{Calculated} &\hfil 3.0 &\hfil 6.6 &\hfil 4.0 & \hfil 9.1\hfil \cr}]

[Figure 2.3.3.1]

Figure 2.3.3.1| top | pdf |

Differences in relative intensities due to preferred orientation as seen in synchrotron X-ray patterns of m-chlorobenzoic acid obtained with a specimen in reflection and transmission compared with calculated pattern. Peaks marked × are impurities, O absent in experimental patterns.

Care is required to make certain the differences are not caused by a few fortuitously oriented large particles.

Various methods have been used to minimize preferred orientation in the specimen preparation (Calvert, Sirianni, Gainsford & Hubbard, 1983[link]; Smith & Barrett, 1979[link]; Jenkins et al., 1986[link]; Bish & Reynolds, 1989[link]). These include using small particles, loading the powder from the back or side of the specimen holder, and cutting shallow grooves to roughen the surface. The powder has also been sifted directly on the surface of a microscope slide or single-crystal plate that has been wetted with the binder or petroleum jelly. Another method is to mix the powder with an inert amorphous powder such as Lindemann glass or rice starch, or add gum arabic, which after setting can be reground to obtain irregular particles. Any additive reduces the intensity and the peak-to-background ratio of the pattern. A promising method that requires a considerable amount of powder is to mix it with a binder and to use spray drying to encapsulate the particles into small spheres which are then used to prepare the specimen (Smith, Snyder & Brownell, 1979[link]).

Preferred orientation would not cause a serious problem in routine identification providing the reference standard had a similar preferred orientation and both patterns were obtained with the same diffractometer geometry. However, when accurate values of the relative intensities are required, as in crystal-structure refinement and quantitative analysis, it may be the major factor limiting the precision.

In practice, it is very difficult to prepare specimens that have a completely random orientation. Even materials that do not have good cleavage or special morphological forms, such as quartz and silicon, show small deviations from a completely random orientation. These show up as errors in the structure refinement and a correction factor is required.

An empirical correction factor determined by the acute angle [\varphi] between the preferred-orientation plane and the diffracting plane (hkl) [I ({\rm corr.})= I (hkl) P (hkl) \varphi \eqno (2.3.3.1)]can be used (Will et al., 1988[link]). Three functions have been used to represent [P(hkl)\varphi] and the term GP is the variable refined: [P(hkl) \varphi = \exp (- {\rm GP} \varphi ^2) \eqno (2.3.3.2)](Rietveld, 1969[link]) for transmission specimens; [P (hkl) \varphi = \exp [{\rm GP} (\pi / 2 - \varphi ^2)] \eqno (2.3.3.3)]for reflection specimens; and [P (hkl) \varphi = ({\rm GP} ^2 \cos^2 \varphi + \sin ^2 \varphi / {\rm GP}) ^{-3/2} \eqno (2.3.3.4)](Dollase, 1986[link]).

These functions are quite similar for small amounts of non-randomness. The preferred-orientation plane is selected by trial and error. For example, a modified fast routine of the powder least-squares refinement program with only seven cycles of refinement on each plane for the first dozen allowed Miller indices can be used to find the plane that gives the lowest R(Bragg) value as shown in Table 2.3.3.1[link]. All three functions improve the R(Bragg) value as shown in Table 2.3.3.2[link] but the evidence is not conclusive as to which is the best. More research is required in this area. Several specimens made of the same material may show different preferred-orientation planes, and in some cases the preferred-orientation plane never occurred in the crystal morphology. A more complicated method examines the polar-axis density distribution using a cubic harmonic expansion to describe the crystallite orientation of a rotating sample (Järvinen, Merisalo, Pesonen & Inkinen 1970[link]; Ahtee, Nurmela, Suortti & Järvinen, 1989[link]; Järvinen, 1993[link]).

Table 2.3.3.1| top | pdf |
Preferred-orientation data for silicon

hklR(Bragg) (%)GP
1111.86−0.11
2202.02−0.11
3112.010.17
4000.86−0.15
3311.73−0.19
4222.430.04
5111.360.19
5312.44−0.08
4421.69−0.19
6201.250.29
5332.40−0.04
Selected preferred orientation plane.

Table 2.3.3.2| top | pdf |
R(Bragg) values obtained with different preferred-orientation formulae

 R(Bragg)
SiSiO2Mg2GeO4
No corrections3.502.5712.5
Gaussian1.651.605.71
Exponential0.751.835.30
March/Dollase0.751.644.87
Preferred-orientation plane100211100

References

First citation Ahtee, M., Nurmela, M., Suortti, P. & Järvinen, M. (1989). Correction for preferred orientation in Rietveld refinement. J. Appl. Cryst. 22, 261–268.Google Scholar
First citation Bish, D. L. & Reynolds, R. C. (1989). Sample preparation for X-ray diffraction. Modern powder diffraction, edited by D. L. Bish & J. E. Post, Chap. 4. Washington: Mineralogical Society of America.Google Scholar
First citation Calvert, L. D., Sirianni, A. F., Gainsford, G. J. & Hubbard, C. R. (1983). A comparison of methods for reducing preferred orientation. Adv. X-ray Anal. 26, 105–110.Google Scholar
First citation Dollase, W. A. (1986). Correction of intensities for preferred orientation in powder diffractometry: application of the March model. J. Appl. Cryst. 19, 267–272.Google Scholar
First citation Järvinen, M. (1993). Application of symmetrized harmonics expansion to correction of the preferred orientation effect. J. Appl. Cryst. 26, 525–531.Google Scholar
First citation Järvinen, M., Merisalo, M., Pesonen, A. & Inkinen, O. (1970). Correction of integrated X-ray intensities for preferred orientation in cubic powders. J. Appl. Cryst. 3, 313–318.Google Scholar
First citation Jenkins, R., Fawcett, T. G., Smith, D. K., Visser, J. W., Morris, M. C. & Frevel, L. K. (1986). International Centre for Diffraction Data. Sample preparation methods in X-ray powder diffraction. Powder Diffr. 1, 51–63.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 Smith, D. K. & Barrett, C. S. (1979). Special handling problems in X-ray diffractometry. Adv. X-ray Anal. 22, 1–12Google Scholar
First citation Smith, S. T., Snyder, R. L. & Brownell, W. E. (1979). Minimization of preferred orientation in powders by spray drying. Adv. X-ray Anal. 22, 77–88.Google Scholar
First citation Will, G., Bellotto, M., Parrish, W. & Hart, M. (1988). Crystal structures of quartz and magnesium germanate by profile analysis of synchrotron-radiation high-resolution powder data. J. Appl. Cryst. 21, 182–191.Google Scholar








































to end of page
to top of page