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

International Tables for Crystallography (2006). Vol. C. ch. 4.2, pp. 251-258

Section 4.2.6.3.3. Comparison of theory with experiment

D. C. Creaghb

4.2.6.3.3. Comparison of theory with experiment

| top | pdf |

In this section, discussion will be focused on (i) the scattering of photons having energies considerably greater than that of the K-absorption edge of the atom from which they are scattered, and (ii) scattering of photons having energies in the neighbourhood of the K-absorption edge of the atom from which they are scattered.

4.2.6.3.3.1. Measurements in the high-energy limit [(\omega/\omega_\kappa\rightarrow0)]

| top | pdf |

In this case, there is some possibility of testing the validity of the relativistic dipole and relativistic multipole theories since, in the high-energy limit, the value of [f'(\omega,0)] must approach a value related to the total self energy of the atom [(E_{\rm tot}/mc^2)]. That there is an atomic number dependent systematic error in the relativistic dipole approach has been demonstrated by Creagh (1984[link]). The question of whether the relativistic multipole approach yields a result in better accord with the experimental data is answered in Table 4.2.6.4[link], where a comparison of values of [f'(\omega,0)] is made for three theoretical data sets (this work; Cromer & Liberman, 1981[link]; Wagenfeld, 1975[link]) with a number of experimental results. These include the `direct' measurements using X-ray interferometers (Cusatis & Hart, 1975[link]; Creagh, 1984[link]), the Kramers–Kronig integration of X-ray attenuation data (Gerward et al., 1979[link]), and the angle-of-the-prism data of Deutsch & Hart (1984b[link]). Also included in the table are `indirect' measurements: those of Price et al. (1978[link]), based on Pendellösung measurements, and those of Grimvall & Persson (1969[link]). These latter data estimate [f'(\omega,{\bf g}_{hkl})] and not [f'(\omega,0)]. Table 4.2.6.4[link] details values of the real part of the dispersion correction for LiF, Si, Al and Ge for the characteristic wavelengths Ag [K\alpha_1], Mo [K\alpha_1] and Cu [K\alpha_1]. Of the atomic species listed, the first three are approaching the high-energy limit at Ag [K\alpha_1], whilst for germanium the K-shell absorption edge lies between Mo [K\alpha_1] and Ag [K\alpha_1].

Table 4.2.6.4| top | pdf |
Comparison of measurements of the real part of the dispersion correction for LiF, Si, Al and Ge for characteristic wavelengths Ag Kα1, Mo Kα1 and Cu Kα1 with theoretical predictions; the experimental accuracy claimed for the experiments is shown thus: (10) = 10% error

SampleReference[f'(\omega,0)]
Cu [K\alpha_1]Mo [K\alpha_1]Ag [K\alpha_1]
LiFTheory   
 This work0.0750.0170.010
 Cromer & Liberman (1981[link])0.0680.0140.006
 Wagenfeld (1975[link])0.0800.0230.015
Experiment   
 Creagh (1984[link])0.085 (5)0.020 (10)0.014 (10)
 Deutsch & Hart (1984b)0.0217 (1)0.0133 (1)
SiTheory   
 This work0.2540.8170.052
 Cromer & Liberman (1981[link])0.2420.0710.042
 Wagenfeld (1975[link])0.2820.1010.071
Experiment   
 Cusatis & Hart (1975[link])0.0863 (2)0.0568 (2)
 Price et al. (1978[link])0.085 (7)0.047 (7)
 Gerward et al. (1979[link])0.244 (7)0.099 (7)0.070 (7)
 Creagh (1984[link])0.236 (5)0.091 (5)0.060 (5)
 Deutsch & Hart (1984b[link])0.0847 (1)0.0537 (1)
AlTheory   
 This work0.2130.06450.041
 Cromer & Liberman (1981[link])0.2030.04860.020
 Wagenfeld (1975[link])0.2350.0760.553
Experiment   
 Creagh (1985[link])0.065 (20)0.044 (20)
 Takama et al. (1982[link])0.20 (5)0.07 (5)0.035 (10)
GeTheory   
 This work−1.0890.1550.302
 Cromer & Liberman (1981[link])−1.1670.0620.197
 Wagenfeld (1975[link])−1.80−0.080.14
Experiment   
 Gerward et al. (1979[link])−1.040.300.43
 Grimvall & Persson (1969[link])−1.790.080.27

The high-energy-limit case is considered first: both the relativistic dipole and relativistic multipole theories underestimate [f'(\omega,0)] for LiF whereas the non-relativistic theory overestimates [f'(\omega,0)] when compared with the experimental data. For silicon, however, the relativistic multipole yields values in good agreement with experiment. Further, the values derived from the work of Takama et al. (1982[link]), who used a Pendellösung technique to measure the atomic form factor of aluminium are in reasonable agreement with the relativistic multipole approach. Also, some relatively imprecise measurements by Creagh (1985[link]) are in better accordance with the relativistic multipole values than with the relativistic dipole values.

Further from the high-energy limit (smaller values of [\omega/\omega_\kappa)], the relativistic multipole approach appears to give better agreement with theory. It must be reported here that measurements by Katoh et al. (1985a[link]) for lithium fluoride at a wavelength of 0.77366 Å yielded a value of 0.018 in good agreement with the relativistic multipole value 0.017.

At still smaller values of [(\omega/\omega_\kappa)], the non-relativistic theory yields values considerably at variance with the experimental data, except for the case of LiF using Cu [K\alpha_1] radiation. The relativistic multipole approach seems, in general, to be a little better than the relativistic approach, although agreement between experiment and theory is not at all good for germanium. Neither of the experiments cited here, however, has claims to high accuracy.

In Table 4.2.6.5[link], a comparison is made of measurements of [f''(\omega,0)] derived from the results of the IUCr X-ray Attenuation Project (Creagh & Hubbell, 1987[link], 1990[link]) with a number of theoretical predictions. The measurements were made on carbon, silicon and copper specimens at the characteristic wavelengths Cu [K\alpha_1], Mo [K\alpha_1] and Ag [K\alpha_1]. The principal conclusion that can be drawn from perusal of Table 4.2.6.5[link] is that only minor, non-systematic differences exist between the predictions of the several relativistic approaches and the experimental results. In contrast, the non-relativistic theory fails for higher values of atomic number.

Table 4.2.6.5| top | pdf |
Comparison of measurements of f′(ω, 0) for C, Si and Cu for characteristic wavelengths Ag Kα1, Mo Kα1 and Cu Kα1 with theoretical predictions; the measurements are from the IUCr X-ray Attenuation Project Report (Creagh & Hubbell, 1987[link], 1990[link]), corrected for the effects of Compton, Laue–Bragg, and small-angle scattering

SampleReference[f'(\omega,0)]
Cu [K\alpha_1]Mo [K\alpha_1]Ag [K\alpha_1]
6CTheory   
 This work0.00910.00160.0009
 Cromer & Liberman (1981[link])0.00910.00160.0009
 Wagenfeld (1975[link])
 Scofield (1973[link])0.00930.00160.0009
 Storm & Israel (1970[link])0.00900.00160.0009
Experiment   
 IUCr Project0.00930.00160.0009
14SiTheory   
 This work0.3300.0700.043
 Cromer & Liberman (1981[link])0.3300.07040.0431
 Wagenfeld (1975[link])0.3300.0710.044
 Scofield (1973[link])0.3320.07020.0431
 Storm & Israel (1970[link])0.3310.06980.0429
Experiment   
 IUCr Project0.3320.06960.0429
29CuTheory   
 This work0.5881.2650.826
 Cromer & Liberman (1981[link])0.5891.2650.826
 Scofield (1973[link])0.5861.2560.826
Experiment   
 IUCr Project0.5881.2670.826

4.2.6.3.3.2. Measurements in the vicinity of an absorption edge

| top | pdf |

The advent of the synchrotron-radiation source as a routine experimental tool and the deep interest that many crystallographers have in both XAFS and the anomalous-scattering determinations of crystal structures have stimulated considerable interest in the determination of the dispersion corrections in the neighbourhood of absorption edges. In this region, the inter­action of the ejected photoelectron with electrons belonging to neighbouring atoms causes the modulations that are referred to as XAFS. Both [f''(\omega,0)] (which is directly proportional to the X-ray scattering cross section) and [f'(\omega,0)] [which is linked to [f''(\omega,0)] through the Kramers–Kronig integral] exhibit these modulations. It is at this point that one must realize that the theoretical tabulations are for the interactions of photons with isolated atoms. At best, a comparison of theory and experiment can show that they follow the same trend.

Measurements have been made in the neighbourhood of the absorption edges of a variety of atoms using the `direct' techniques interferometry, Kramers–Kronig, refraction of a prism and critical-angle techniques, and by the `indirect' refinement techniques. In Table 4.2.6.6[link], a comparison is made of experimental values taken at or near the absorption edges of copper, nickel and niobium with theoretical predictions. These have not been adjusted for any energy window that might be thought to exist in any particular experimental configuration. The theoretical values for niobium have been calculated at the energy at which the experimentalists claimed the experiment was conducted.

Table 4.2.6.6| top | pdf |
Comparison of [f'(\omega_A,0)] for copper, nickel, zirconium, and niobium for theoretical and experimental data sets; in this table: BR [\equiv] Bragg reflection; IN [\equiv] interferometer; KK [\equiv] Kramers–Kronig; CA = critical angle; and REF = reflectivity; measurements have been made for the K-absorption edges of copper and nickel and near the K-absorption edges of zirconium and niobium; claimed experimental errors are not worse than 5%

ReferenceMethod[f'(\omega_A,0)]
CuNiNbZr
Experiment     
  Freund (1975[link])BR−8.2   
  Begum, Hart, Lea & Siddons (1986[link])IN−7.84−7.66  
  Bonse & Materlik (1972[link])IN −8.1  
  Bonse, Hartmann-Lotsch & Lotsch (1983a[link])IN−8.3   
  Hart & Siddons (1981[link])IN−9.3−9.2−4.396−6.670
  Kawamura & Fukimachi (1978; cited in Bonse & Hartmann-Lotsch, 1984[link])KK −7.9  
  Dreier et al. (1984[link])KK−8.2−7.8 −7.83
 IN−8.3−8.1  
  Bonse & Hartmann-Lotsch (1984[link])KK−8.3−7.7  
  Fukamachi et al. (1978; cited in Bonse & Hartmann-Lotsch, 1984[link])KK−8.8   
 CA−10.0   
      
  Bonse & Henning (1986[link])IN  −7.37; −7.73 
 KK  −7.21; −7.62 
  Stanglmeier, Lengeler, Weber, Gobel & Schuster (1992[link])REF−8.5−8.1  
  Creagh (1990[link], 1993[link])REF−8.2−7.7 −6.8
Theory     
  Cromer & Liberman (1981[link]) −13.50−9.45−4.20; −7.39−6.207
  This work −9.5−9.40−4.04; −7.23−6.056
  Averaged values (5 eV window) −9.0−7.53−8.18−6.04

Despite the considerable experimental difficulties and the wide variety of experimental apparatus, there appears to be close agreement between the experimental data for each type of atom. There appears to be, however, for both copper and nickel, a large discrepancy between the theoretical values and the experimental values. It must be remembered that the experimental values are averages of the value of [f'(\omega,0)], the average being taken over the range of photon energies that pass through the device when it is set to a particular energy value. Furthermore, the exact position of the wavelength chosen may be in doubt in absolute terms, especially when synchrotron-radiation sources are used. Therefore, to be able to make a more realistic comparison between theory and experiment, the theoretical data gained using the relativistic multipole approach (this work) were averaged over a rectangular energy window of 5 eV width in the region containing the absorption edge. The rectangular shape arises because of the shape of the reflectivity curve and 5 eV was chosen as a result of (i) analysis of the characteristics of the interferometers used by Bonse et al. and Hart et al., and (ii) a statement concerning the experimental bandpass of the interferometer used by Bonse & Henning (1986[link]). It must also be borne in mind that mechanical vibrations and thermal fluctuations can broaden the energy window and that 5 eV is not an overestimate of the width of this window. Note that for elements with atomic numbers less than 40 the experimental width is greater than the line width.

For the Bonse & Henning (1986[link]) data, two values are listed for each experiment. Their experiment demonstrates the effect the state of polarization of the incoming photon has on the value of [f'(\omega,0)]. Similar X-ray dichroism has been shown for sodium bromate by Templeton & Templeton (1985b[link]) and Chapuis et al. (1985[link]). The theoretical values are for averaged polarization in the incident photon beam. Another important feature is the difference of 0.16 electrons between the Kramers–Kronig and the interferometer values. Bonse & Henning (1986[link]) did not add the relativistic correction term to their Kramers–Kronig values. Inclusion of this term would have reduced the quoted values by 0.20, bringing the two data sets into close agreement with one another.

Katoh et al. (1985b[link]) have made measurements spanning the K-absorption edge of germanium using the deviation by a prism method, and these data have been shown to be in excellent agreement with the theory on which these tables are based (Creagh, 1993[link]). In contrast, the theoretical approach of Pratt, Kissel & Bergstrom (1994[link]) does not agree so well, especially near to, and at higher photon energies, than the K-edge energy. Also, Chapuis et al. (1985[link]) have measured the dispersion corrections for holmium in [HoNa(edta)]·8H2O for the characteristic emission lines Cu [K\alpha_1], Cu [K\alpha_2], Cu Kβ, and Mo [K\alpha_1] using a refinement technique. Their results are in reasonable agreement with the relativistic multipole theory, e.g. for [f'(\omega,{\boldDelta})] at the wavelength of Cu [K\alpha_1] experiment gives −(16.0 ± 0.2) whereas the relative multipole approach yields −15.0. For Cu [K\alpha_2], experiment yields −(13.9 ± 0.3) and theory gives −13.67. The discrepancy between theory and experiment may well be explained by the oxidation state of the holmium ion, which is in the form Ho3+. The oxidation state of an atom affects both the position of the absorption edge and the magnitude of the relativistic correction. Both of these will have a large influence on the value of [f'(\omega,{\boldDelta})] in the neighbourhood of the absorption edge, Another problem that may be of some significance is the natural width of the absorption edge, about 60 eV. What is remarkable is the extent of the agreement between theory and experiment given the nature of the experiment. In these experiments, the intensities of many reflections (usually nearly 1000) are analysed and compared. Such a procedure can be followed only if there is no dependence of [f'(\omega,{\boldDelta})] on [{\boldDelta}].

It had often been thought that the dispersion corrections should exhibit some functional dependence on scattering angle. Indeed, some texts ascribe to these corrections the same functional dependence on angle of scattering as the form factor. A fundamental dependence was also predicted theoretically on the basis of non-relativistic quantum mechanics (Wagenfeld, 1975[link]). This prediction is not supported by modern approaches using relativistic quantum mechanics [see, for example, Kissel et al. (1980[link])]. Reference to Tables 4.2.6.4[link] and 4.2.6.6[link] shows that the agreement between experimental values derived from diffraction experiments and those derived from `direct' experiments is excellent. They are also in excellent agreement with the recent calculations, using relativistic quantum mechanics, so that it may be inferred that there is indeed no functional dependence of the dispersion corrections on scattering angle. Moreover, Suortti, Hastings & Cox (1985[link]) have recently demonstrated that [f'(\omega,{\boldDelta})] was independent of [{\boldDelta}] in a powder-diffraction experiment using a nickel specimen.

4.2.6.3.3.3. Accuracy in the tables of dispersion corrections

| top | pdf |

Experimentalists must be aware of two potential sources of error in the values of [f'(\omega,0)] listed in Table 4.2.6.5[link]. One is computational, arising from the error in calculating the relativistic correction. Stibius-Jensen (1980[link]) has suggested that this error may be as large as [\pm0.25(E_{\rm tot}/mc^2)]. This means, for example, that the real part of the dispersion correction [f'(\omega,0)] for lead at the wavelength of 0.55936 Å is −(1.168 ± 0.146). The effect of this error is to shift the dispersion curve vertically without distorting its shape. Note, however, that the direction of the shift is either up or down for all atoms: the effect of multipole cancellation and retardation will be in the same direction for all atoms.

The second possible source of error occurs because the position of the absorption edge varies somewhat depending on the oxidation state of the scattering atom. This has the effect of displacing the dispersion curve laterally. Large discrepancies may occur for those regions in which the dispersion corrections are varying rapidly with photon energy, i.e. near absorption edges.

It must also be borne in mind that in the neighbourhood of an absorption edge polarization effects may occur. The tables are valid only for average polarization.

4.2.6.3.3.4. Towards a tensor formalism

| top | pdf |

The question of how best to describe the interaction of X-rays with crystalline materials is quite difficult to answer. In the form factor formalism, the atoms are supposed to scatter as though they are isolated atoms situated at fixed positions in the unit cell. In the vast majority of cases, the polarization on scattering is not detected, and only the scattered intensities are measured. From the scattered intensities, the distribution of the electron density within the unit cell is calculated, and the difference between the form-factor model and that calculated from the intensities is taken as a measure of the nature and location of chemical bonds between atoms in the unit cell. This is the zeroth-order approximation to a solution, but it is in fact the only way crystal structures are solved ab initio.

The existence of chemical bonding imposes additional restrictions on the symmetry of lattices, and, if the associated influence this has on the complexity of energy levels is taken into account, significant changes in the scattering factors may occur in the neighbourhood of the absorption edges of the atoms comprising the crystal structure. The magnitudes of the dispersion corrections are sensitive to the chemical state, particularly oxidation state, and phenomena similar to those observed in the XAFS case (Section 4.2.4[link]) are observed.

The XAFS interaction arising from the presence of neighbouring atoms is proportional to [f''(\omega,0)] and therefore is related to [f'(\omega,0)] through the Kramers–Kronig integral. It is not surprising that these modulations are observed in diffracted intensities in those X-ray diffraction experiments where the photon energy is scanned through the absorption edge of an atomic species in the crystal lattice. Studies of this type are referred to as diffraction absorption fine structure (DAFS) experiments. A recent review of work performed using counter techniques has been given by Sorenson (1994[link]). Creagh & Cookson (1995[link]) have described the use of imaging-plate techniques to study the structure and site symmetry using the DAFS technique. This technique has the ability to discriminate between different lattice sites in the unit cell occupied by an atomic species. XAFS cannot make this discrimination. The DAFS modulations are small perturbations to the diffracted intensities. They are, however, significantly larger than the tensor effects described in the following paragraphs.

In the case where the excited state lacks high symmetry and is oriented by crystal bonding, the scattering can no longer be described by a scalar scattering factor but must be described by a symmetric second-rank tensor. The consequences of this have been described by Templeton (1994[link]). It follows therefore that material media can be optically active in the X-ray region. Hart (1994[link]) has used his unique polarizing X-ray optical devices to study, for example, Faraday rotation in such materials as iron, in the region of the iron K-absorption edge, and cobalt(III) bromide monohydrate in the region of the cobalt K-absorption edge.

The theory of anisotropy in anomalous scattering has been treated extensively by Kirfel (1994[link]), and Morgenroth, Kirfel & Fischer (1994[link]) have extended this to the description of kinematic diffraction intensities in lattices containing anisotropic anomalous scatterers. Their treatment was developed for space groups up to orthorhombic symmetry.

All the preceding treatments apply to scattering in the neighbourhood of an absorption edge, and to a fairly restricted class of crystals for which the local site symmetry of the electron density of states in the excited state is very different from the apparent crystal symmetry.

These approaches seek to treat the scattering from the crystal as though the scattering from each atomic position can be described by a symmetric second-rank tensor whose properties are determined by the point-group symmetries of those sites. Clearly, this procedure cannot be followed unless the structure has been solved by the usual method. The tensor approach can then be used to explain apparent deficiencies in that model such as the existence of `forbidden' reflections, birefringence, and circular dichroism.

Scattering of X-rays from the electron spins in anti-ferromagnetically ordered materials can also be described by imposing a tensor description on the form factor (Blume, 1994[link]). The tensor in this case is a fourth-rank tensor, and the strength of the interaction, even for the favourable case of resonance scattering, is several orders of magnitude lower in intensity than the polarization effects. Nevertheless, studies have been made on holmium and uranium arsenide, and significant magnetic Bragg scattering has been observed.

All the cases cited above represent exciting, state-of-the-art, scientific studies. However, none of the work will assist in the solution of crystal structures directly. Researchers should avoid the temptation, in the first instance, to ascribe anything but a scalar value to the form factor.

4.2.6.3.3.5. Summary

| top | pdf |

For the imaginary part of the dispersion correction [f''(\omega,{\boldDelta})], the following observations can be made.

  • (i) Measurements of the linear absorption coefficient [\mu_l] from which [f'(\omega,0)] is deduced should follow the recommendations set out in Subsection 4.2.3.2[link].

  • (ii) There is no rational basis for preferring one set of relativistic calculations of atomic scattering cross sections over another, as Creagh & Hubbell (1987[link], 1990[link]) and Kissel et al. (1980[link]) have shown.

  • (iii) The total scattering cross section for an ensemble of atoms is not simply the sum of the individual scattering cross sections in the neighbourhood of an absorption edge and therefore [f'(\omega,0)] will fluctuate as [\omega\rightarrow\omega_\kappa].

  • (iv) There is no dependence of [f''(\omega,{\boldDelta})] and [{\boldDelta}].

For the real part of the dispersion correction [f'(\omega,{\boldDelta})], the following observations can be made.

  • (i) The relativistic multipole values listed here tend to accord better with experiment than the non-relativistic and relativistic dipole values.

  • (ii) There is no dependence of [f'(\omega,{\boldDelta})] on [{\boldDelta}].

  • (iii) The theoretical tables are calculated for averaged polarizations.

  • (iv) Experimentalists wishing to compare their data with theoretical predictions should take account of the energy bandpass of their system when determining the appropriate theoretical value. They should also be aware of the fact that the position of the absorption edge depends on the oxidation state of the scattering atom, and that there is an inaccuracy in the tables of [f'(\omega,0)] of either [+0.20(E_{\rm tot}/mc^2)] or [-0.10(E_{\rm tot}/mc^2)].

References

First citation Begum, R., Hart, M., Lea, K. R. & Siddons, D. P. (1986). Direct measurements of the complex X-ray atomic scattering factors for elements by X-ray interferometry at the Daresbury synchrotron radiation source. Acta Cryst. A42, 456–463.Google Scholar
First citation Blume, M. (1994). Magnetic effects in anomalous dispersion. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 495–512. Amsterdam: North Holland.Google Scholar
First citation Bonse, U. & Hartmann-Lotsch, I. (1984). Kramers–Kronig correlation of measured f′(E) and f ′′(E) values. Nucl. Instrum. Methods, 222, 185–188.Google Scholar
First citation Bonse, U., Hartmann-Lotsch, I. & Lotsch, H. (1983a). Interferometric measurement of absorption μ(E) and dispersion f′(E) at the K edge of copper. EXAFS and near edge structure, edited by A. Bianconi, I. Incoccia & A. S. Stipcich, pp. 376–377. Berlin: Springer.Google Scholar
First citation Bonse, U. & Henning, A. (1986). Measurement of polarization isotropy of the anomalous forward scattering amplitude at the niobium K-edge in LiNbO3. Nucl. Instrum. Methods, A246, 814–816.Google Scholar
First citation Bonse, U. & Materlik, G. (1972). Dispersionskorrektur für Nickel nahe der K-Absorptionskante. Z. Phys. 253, 232–239.Google Scholar
First citation Chapuis, G., Templeton, D. H. & Templeton, L. K. (1985). Solving crystal structures using several wavelengths from conventional sources. Anomalous scattering by holmium. Acta Cryst. A41, 274–278.Google Scholar
First citation Creagh, D. C. (1984). The real part of the forward scattering factor for aluminium. Unpublished.Google Scholar
First citation Creagh, D. C. (1985). Theoretical and experimental techniques for the determination of X-ray anomalous dispersion corrections. Aust. J. Phys. 38, 371–404.Google Scholar
First citation Creagh, D. C. (1990). Tables of X-ray absorption corrections and dispersion corrections: the new versus the old. Nucl. Instrum. Methods, A295, 417–434.Google Scholar
First citation Creagh, D. C. (1993). f′: its present and its future. Ind. J. Phys. 67B, 511–525.Google Scholar
First citation Creagh, D. C. & Cookson, D. J. (1995). Diffraction anomalous fine structure study of basic zinc sulphate and basic zinc sulphonate. In Photon Factory Activity Report 1994, edited by K. Nasu. National Laboratory for High Energy Physics, Tsukuba 93-0305, Japan.Google Scholar
First citation Creagh, D. C. & Hubbell, J. H. (1987). Problems associated with the measurement of X-ray attenuation coefficients. I. Silicon. Acta Cryst. A43, 102–112.Google Scholar
First citation Creagh, D. C. & Hubbell, J. H. (1990). Problems associated with the measurement of X-ray attenuation coefficients. II. Carbon. Acta Cryst. A46, 402–408.Google Scholar
First citation Cromer, D. T. & Liberman, D. A. (1981). Anomalous dispersion calculations near to and on the long-wavelength side of an absorption edge. Acta Cryst. A37, 267–268.Google Scholar
First citation Cusatis, C. & Hart, M. (1975). Dispersion correction measurements by X-ray interferometry. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 57–68. Copenhagen: Munksgaard.Google Scholar
First citation Deutsch, M. & Hart, M. (1984b). X-ray refractive-index measurement in silicon and lithium fluoride. Phys. Rev. B, 30, 640–642.Google Scholar
First citation Dreier, P., Rabe, P., Malzfeldt, W. & Niemann, W. (1984). Anomalous X-ray scattering factors calculated from experimental absorption factor. J. Phys. C, 17, 3123–3136.Google Scholar
First citation Freund, A. (1975). Anomalous scattering of X-rays in copper. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 69–86. Copenhagen: Munksgaard.Google Scholar
First citation Gerward, L., Thuesen, G., Stibius-Jensen, M. & Alstrup, I. (1979). X-ray anomalous scattering factors for silicon and germanium. Acta Cryst. A35, 852–857.Google Scholar
First citation Grimvall, G. & Persson, E. (1969). Absorption of X-rays in germanium. Acta Cryst. A25, 417–422.Google Scholar
First citation Hart, M. (1994). Polarizing X-ray optics and anomalous dispersion in chiral systems. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 103–118. Amsterdam: North Holland.Google Scholar
First citation Hart, M. & Siddons, D. P. (1981). Measurements of anomalous dispersion corrections made from X-ray interferometers. Proc. R. Soc. London Ser. A, 376, 465–482.Google Scholar
First citation Katoh, H., Shimakura, H., Ogawa, T., Hattori, S., Kobayashi, Y., Umezawa, K., Ishikawa, T. & Ishida, K. (1985a). Measurement of X-ray refractive index by multiple reflection diffractometer. Activity Report, KEK, Tsukuba, Japan.Google Scholar
First citation Katoh, H., Shimakura, H., Ogawa, T., Hattori, S., Kobayashi, Y., Umezawa, K., Ishikawa, T. & Ishida, K. (1985b). Measurement of the X-ray anomalous scattering of the germanium K-edge with synchrotron radiation. J. Phys. Soc. Jpn, 54, 881–884.Google Scholar
First citation Kirfel, A. (1994). Anisotropy of anomalous scattering in single crystals. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 231–256. Amsterdam: North Holland.Google Scholar
First citation Kissel, L., Pratt, R. H. & Roy, S. C. (1980). Rayleigh scattering by neutral atoms, 100 eV to 10 MeV. Phys. Rev. A, 22, 1970–2004.Google Scholar
First citation Morgenroth, W., Kirfel, A. & Fischer, K. (1994). Computing kinematic diffraction intensities with anomalous scatterers – `forbidden' axial reflections in space groups up to orthorhombic symmetry. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 257–264. Amsterdam: North Holland.Google Scholar
First citation Pratt, R. H., Kissel, L. & Bergstrom, P. M. Jr (1994). New relativistic S-matrix results for scattering – beyond the anomalous factors/beyond impulse approximation. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 9–34. Amsterdam: North Holland.Google Scholar
First citation Price, P. F., Maslen, E. N. & Mair, S. L. (1978). Electron-density studies. III. A re-evaluation of the electron distribution in crystalline silicon. Acta Cryst. A34, 183–193.Google Scholar
First citation Scofield, J. H. (1973). Theoretical photoionization cross sections from 1 to 1500 keV. Report UCRL-51326. Lawrence Livermore National Laboratory, Livermore, CA, USA.Google Scholar
First citation Sorenson, L. B., Cross, J. O., Newville, M., Ravel, B., Rehr, J. J., Stragier, H., Bouldin, C. E. & Woicik, J. C. (1994). Diffraction anomalous fine structure: unifying X-ray diffraction and X-ray absorption with DAFS. In Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 389–420. Amsterdam: North Holland.Google Scholar
First citation Stanglmeier, F., Lengeler, B., Weber, W., Gobel, H. & Schuster, M. (1992). Determination of the dispersive correction f′(E) to the atomic form factor from X-ray reflection. Acta Cryst. A48, 626–639.Google Scholar
First citation Stibius-Jensen, M. (1980). Atomic X-ray scattering factors for forward scattering beyond the dipole approximation. Personal communication.Google Scholar
First citation Stuhrmann, H. (1982). Editor. Uses of synchrotron radiation in biology. Esp. Properties of synchrotron radiation, Chap. 1, by G. Materlik. London: Academic Press.Google Scholar
First citation Suortti, P., Hastings, J. B. & Cox, D. E. (1985). Powder diffraction with synchrotron radiation. II. Dispersion factors of Ni. Acta Cryst. A41, 417–420.Google Scholar
First citation Takama, T., Kobayashi, K. & Sato, S. (1982). Determination of the atomic scattering factor of aluminium by the Pendellösung beat measurement using white radiation. Trans. Jpn. Inst. Met. 23, 153–160.Google Scholar
First citation Takama, T. & Sato, S. (1982). Atomic scattering factors of copper determined by Pendellösung beat measurements using white radiation. Philos. Mag. 45, 615–626.Google Scholar
First citation Templeton, D. H. (1994). X-ray resonance, then and now. In Resonant anomalous X-ray scattering: theory and applications, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 1–7. Amsterdam: North Holland.Google Scholar
First citation Templeton, D. H. & Templeton, L. K. (1985b). Tensor X-ray optical properties of the bromate ion. Acta Cryst. A41, 133–142.Google Scholar
First citation Wagenfeld, H. (1975). Theoretical computations of X-ray dispersion corrections. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 12–23. Copenhagen: Munksgaard.Google Scholar








































to end of page
to top of page