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. 191-258
https://doi.org/10.1107/97809553602060000592

Chapter 4.2. X-rays

U. W. Arndt,a D. C. Creagh,b R. D. Deslattes,c J. H. Hubbell,d P. Indelicato,e E. G. Kessler Jrf and E. Lindrothg

a MRC Laboratory of Molecular Biology, Hills Road, Cambridge CB2 2QH, England,bDivision of Health, Design, and Science, University of Canberra, Canberra, ACT 2601, Australia,cNational Institute of Standards and Technology, Gaithersburg, MD 20899, USA,dRoom C314, Radiation Physics Building, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA,eLaboratoire Kastler-Brossel, Case 74, Université Pierre et Marie Curie, 4 Place Jussieu, F-75252 Paris CEDEX 05, France,fAtomic Physics Division, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA, and gDepartment of Atomic Physics, Stockholm University, S-104 05 Stockholm, Sweden

The generation of X-rays is discussed in the first section of this chapter. X-rays are generated by (1) the bombardment of a target by electrons, (2) the decay of certain radio isotopes, (3) as part of the synchrotron-radiation spectrum and (4) in plasmas produced by bombarding targets with high-energy laser beams. The spectra produced by the first method consist of a line spectrum characteristic of the target material accompanied by a continuum of white radiation (Bremsstrahlung). The intensities and wavelengths of these components are discussed and the nomenclatures of the lines in the characteristic spectrum are compared. Synchrotron radiation and plasma generation of X-rays are discussed very briefly. X-ray wavelengths are discussed in the second section of the chapter. Tables of K- and L-series reference wavelengths and K- and L-emission lines and absorption edges are provided. X-ray absorption spectra are discussed in the third section of the chapter. A detailed discussion of the problems associated with the measurement of X-ray absorption and the requirements for the absolute measurement of the X-ray attenuation coefficients is followed by a description of the problems which are encountered in making measurements at X-ray absorption edges of atomic species in materials. The popular, and extremely valuable, experimental technique of X-ray absorption fine structure (XAFS) is discussed. This is a relative, rather than an absolute measurement of X-ray absorption, and the theoretical analysis of XAFS spectra depends on the deviation of the data points from a cubic spline fit to the data rather that the deviation from the extrapolation of the free-atom absorption. A brief description of the origin and use of X-ray absorption near edge structure (XANES) is given. In the fourth section of the chapter, X-ray absorption (or attenuation) coefficients are considered. A detailed description of the theoretical and experimental techniques for the determination of X-ray absorption (and attenuation coefficients) is given. The tables supercede the existing experimental and theoretical tables that were available up to 2001. The theoretical values given here are restricted to the characteristic radiations commonly available from laboratory X-ray sources (Ti Kβ to Ag Kα). Computational problems become significant for elements in the lanthanide and actinide series of the periodic table. These tables include a significant recalculation of the absorption and attenuation coefficients for the lanthanide series. Note that although many theory-based tables exist, few accurate experimental measurements have been made on atomic species with atomic numbers greater than 50, especially in the region of absorption edges. A comparison is given of the extent to which theoretical and experimental data agree. A recent tabulation discusses soft X-ray absorption in the XANES region. Filters and monochromators are discussed in the fifth section of the chapter. This section serves as an introduction to the use of filters and monochromators in experimental apparatus. It discusses, using synchrotron-radiation experimental configurations as examples, the use of X-ray reflectivity, X-ray refraction and X-ray Bragg (and Laue) or Fresnel scattering for the production of monochromatic or quasi- monochromatic (pink) beams from polychromatic sources. Amongst the techniques discussed are: bent and curved mirrors, capillaries, quasi-Bragg (multilayer) mirrors, multiple filters, crystal monochromators, polarization and polarization-producing systems, and focusing using a Bragg–Fresnel optical system. The development of monochromators for synchrotron-radiation research is a rapidly evolving field, but in practice all depend on either the processes of reflection, refraction or scattering, or combinations of these processes. In the final section of the chapter, X-ray dispersion corrections are considered. The various theoretical methods for the computation of the X-ray dispersion corrections and the experimental techniques used for their determination are described in some detail. The theoretical values given here are restricted to the characteristic radiations commonly available from laboratory X-ray sources (Ti Kβ to Ag Kα). An artificial distinction is made between the relativistic Dirac–Hartree–Fock–Slater (RDHFS) formalism used to calculate the data in these tables and the S-matrix method used by a number of theoreticians. Formally, the two techniques are equivalent: in terms of the computations the two approaches are different, and the outcomes are different. Comparison with experimental data shows that the RDHFS formalism gives a better fit to experimental results.

Keywords: absorption; anomalous dispersion; atomic form factor; atomic scattering factors; attenuation coefficients; Auger shifts; Bijvoet-pair techniques; black-body radiation in X-ray region; characteristic line spectrum; continuous spectrum; Dirac–Fock method; dispersion corrections; EXAFS; extended X-ray absorption fine structure; filters; Friedel-pair techniques; generation of X-rays; intensity; monochromators; normal attenuation; photo-effect data; QED corrections; Rayleigh scattering; scattering; synchrotron radiation; theoretical Rayleigh scattering data; theoretical photo-effect data; wavelengths; X-ray absorption coefficients; X-ray absorption spectra; X-ray attenuation coefficients; X-ray dispersion corrections; X-ray sources; X-ray tubes; X-ray wavelengths; XAFS; X-ray absorption near-edge structure; XANES.

4.2.1. Generation of X-rays

| top | pdf |
U. W. Arndta

X-rays are produced by the interaction of charged particles with an electromagnetic field. There are four sources of X-rays that are of interest to the crystallographer.

  • (1) The bombardment of a target by electrons produces a continuous (`white') X-ray spectrum, called Bremsstrahlung, which is accompanied by a number of discrete spectral lines characteristic of the target material. The high-vacuum, or Coolidge, X-ray tube is the most important X-ray source for crystallographic studies.

  • (2) The decay of natural or artificial radio isotopes is often accompanied by the emission of X-rays. Radioactive X-ray sources are often used for the calibration of X-ray detectors. Mössbauer sources have the narrowest known spectral bandwidth and are used in nuclear resonance scattering studies.

  • (3) Sources of synchrotron radiation produced by relativistic electrons in orbital motion are of growing importance.

  • (4) X-rays are also produced in plasmas generated by the bombardment of targets by high-energy laser beams, but to date the yield has been principally in the form of soft X-rays.

The classical text on the generation and properties of X-rays is that by Compton & Allison (1935[link]), which still summarizes much of the information required by crystallographers. There is a more recent comprehensive book by Dyson (1973[link]). X-ray physics has received a new impetus on the one hand through the development of X-ray microprobe analysis dealt with in a number of monographs (Reed, 1975[link]; Scott & Love, 1983[link]) and on the other hand through the increasing utilization of synchrotron-radiation sources (see Subsection 4.2.1.5[link]).

4.2.1.1. The characteristic line spectrum

| top | pdf |

Characteristic X-ray emission originates from the radiative decay of electronically highly excited states of matter. We are concerned mostly with excitation by electron bombardment of a target that results in the emission of spectral lines characteristic of the target elements. The electronic states occurring as initial and final states of a process involving the absorption of emission of X-rays are called X-ray levels. Levels involving the removal of one electron from the configuration of the neutral ground state are called normal X-ray levels or diagram levels.

Table 4.2.1.1[link] shows the relation between diagram levels and electron configurations. The notation used here is the IUPAC notation (Jenkins, Manne, Robin & Senemaud, 1991[link]), which uses arabic instead of the former roman subscripts for the levels. The IUPAC recommendations are to refer to X-ray lines by writing the initial and final levels separated by a hyphen, e.g. Cu K-L3 and to abandon the Siegbahn (1925[link]) notation, e.g. Cu Kα1, which is based on the relative intensities of the lines. The correspondence between the two notations is shown in Table 4.2.1.2[link]. Because this substitution has not yet become common practice, however, the Siegbahn notation is retained in Section 4.2.2[link], in which the wavelengths of the characteristic emission lines and absorption edges are discussed.

Table 4.2.1.1| top | pdf |
Correspondence between X-ray diagram levels and electron configurations; from Jenkins, Manne, Robin & Senemaud (1991[link]), courtesy of IUPAC

LevelElectron configurationLevelElectron configurationLevelElectron configuration
K[1s^{-1}]N1[4s^{-1}][O_1][5s^{-1}]
L1[2s^{-1}]N2[4p^{-1}_{1/2}][O_2][5p^{-1}_{1/2}]
L2[2p^{-1}_{1/2}]N3[4p^{-1}_{3/2}][O_3][5p^{-1}_{3/2}]
L3[2p^{-1}_{3/2}]N4[4d^{-1}_{3/2}][O_4][5d^{-1}_{3/2}]
M1[3s^{-1}]N5[4d^{-1}_{5/2}][O_5][5d^{-1}_{5/2}]
M2[3p^{-1}_{1/2}]N6[4f^{-1}_{5/2}][O_6][5f^{-1}_{5/2}]
M3[3p^{-1}_{3/2}]N7[4f^{-1}_{7/2}][O_7][5f^{-1}_{7/2}]
M4[3d^{-1}_{3/2}]    
M5[3d^{-1}_{5/2}]    

Table 4.2.1.2| top | pdf |
Correspondence between IUPAC and Siegbahn notations for X-ray diagram lines; from Jenkins, Manne, Robin & Senemaud (1991[link]), courtesy of IUPAC

SiegbahnIUPACSiegbahnIUPACSiegbahnIUPAC
[K\alpha_1][K\hbox{-}L_3][L\alpha_1][L_3\hbox{-}M_5][L\gamma_1][L_2\hbox{-}N_4]
[K\alpha_2][K\hbox{-}L_2][L\alpha_2][L_3\hbox{-}M_4][L\gamma_2][L_1\hbox{-}N_2]
[K\beta_1][K\hbox{-}M_3][L\beta_1][L_2\hbox{-}M_4][L\gamma_3][L_1\hbox{-}N_3]
[K\beta^1_2][K\hbox{-}N_3][L\beta_2][L_3\hbox{-}N_5][L\gamma_4][L_1\hbox{-}O_3]
[K\beta^{11}_2][K\hbox{-}N_2][L\beta_3][L_1\hbox{-}M_3][L\gamma'_4][L_1\hbox{-}O_2]
[K\beta_3][K\hbox{-}M_2][L\beta_4][L_1\hbox{-}M_2][L\gamma_5][L_2\hbox{-}N_1]
[K\beta^1_4][K\hbox{-}N_5][L\beta_5][L_3\hbox{-}O_{4,5}][L\gamma_6][L_2\hbox{-}O_4]
[K\beta^{11}_4][K\hbox{-}N_4][L\beta_6][L_3\hbox{-}N_1][L\gamma_8][L_2\hbox{-}O_1]
[K\beta_{4x}][K\hbox{-}N_4][L\beta_7][L_3\hbox{-}O_1][L\gamma'_8][L_2\hbox{-}N_{6(7)}]
[K\beta^1_5][K\hbox{-}M_5][L\beta'_7][L_3\hbox{-}N_{6,7}][L_\eta][L_2\hbox{-}M_1]
[K\beta^{11}_5][K\hbox{-}M_4][L\beta_9][L_1\hbox{-}M_5][Ll][L_3\hbox{-}M_1]
  [L\beta_{10}][L_1\hbox{-}M_4][Ls][L_3\hbox{-}M_3]
  [L\beta_{15}][L_3\hbox{-}N_4][Lt][L_3\hbox{-}M_2]
  [L\beta_{17}][L_2\hbox{-}M_3][Lu][L_3\hbox{-}N_{6,7}]
    [Lv][L_2\hbox{-}N_{6(7)}]

SiegbahnIUPAC
[M\alpha_1][M_5\hbox{-}N_7]
[M\alpha_2][M_5\hbox{-}N_6]
[M\beta][M_4\hbox{-}N_6]
[M\gamma][M_3\hbox{-}N_5]
[M\zeta][M_{4,5}\hbox{-}N_{2,3}]

In the case of unresolved lines, such as [K\hbox{-}L_2] and [K\hbox{-}L_3], the recommended IUPAC notation is [K\hbox{-}L_{2,3}].

4.2.1.1.1. The intensity of characteristic lines

| top | pdf |

The efficiency of the production of characteristic radiation has been calculated by a number of authors (see, for example, Dyson, 1973[link], Chap. 3). For a particular line, it depends on the fluorescence yield, that is the probability that the decay of an excited state leads to the emission of a photon, on the statistical weights of the X-ray levels involved, on the effects of the penetration and slowing down of the bombarding electrons in the target, on the fraction of electrons back-scattered out of the target, and on the contribution caused by fluorescent X-rays produced indirectly by the continuous spectrum. The emerging X-ray intensity is further affected by the partial absorption of the generated X-rays in the target.

Dyson (1973[link]) has also reviewed calculations and measurements made of the relative intensities of different lines in the K spectrum. The ratio of the [K\alpha_2] to [K\alpha_3] intensities is very close to 0.5 for Z between 23 and 48. The ratio of [K\beta_3] to [K\alpha_2] rises fairly linearly with Z from 0.2 at Z = 20 to 0.4 at Z = 80 and that of [K\beta_1] to [K\alpha_2] is near zero at Z = 29 and rises linearly with Z to about 0.1 at Z = 80. Relative intensities of lines in the L spectrum are given by Goldberg (1961[link]).

Green & Cosslett (1968[link]) have made extensive measurements of the efficiency of the production of characteristic radiation for a number of targets and for a range of electron accelerating voltages. Their results can be expressed empirically in the form [N_K/4\pi=N_0/4\pi(E_0-E_K-1){}^{1.63},\eqno (4.2.1.1)]where [N_K/4\pi] is the generated number of Kα photons per steradian per incident electron, N0 is a function of the atomic number of the target, E0 is the electron energy in keV and [E_K] is the excitation potential in keV. It should be noted that [N_K/4\pi] decreases with increasing Z.

For a copper target, this expression becomes [N_K/4\pi=1.8\times10^{-6}\,(E_0 - 8.9){}^{1.63}\eqno (4.2.1.2)]or [N'_K/4\pi=1.1\times10^{10}\,(E_0 - 8.9){}^{1.63},\eqno (4.2.1.3)]where [N'_K/4\pi] is the number of Kα photons per steradian per second per milliampere of tube current.

These expressions are probably accurate to within a factor of 2 up to values of [E_0/E_K] of about 10. Guo & Wu (1985[link]) found a linear relationship for the emerging number of photons with electron energy in the range [2\lt E_0/E_K\lt 5].

To obtain the number of photons that emerge from the target, the above expressions have to be corrected for absorption of the generated radiation in the target. The number of photons emerging at an angle [\varphi] to the surface, for normal electron incidence, is usually written [N_\varphi/4\pi=f(\chi)N/4\pi,\eqno (4.2.1.4)]where [\chi=(\mu/\rho)\hbox{ cosec }\varphi] (Castaing & Descamps, 1955[link]). Green (1963[link]) gives experimental values of the correction factor f(χ) for a series of targets over a range of electron energies. His curves for a copper target are given in Fig. 4.2.1.1[link] . It will be noticed that the correction factor increases with increasing electron energy since the effective depth of X-ray generation increases with voltage. As a result, curves of [N_\varphi] as a function of [E_0] have a broad maximum that is displaced towards lower voltages as [\varphi] decreases, as shown in the experimental curves for copper K radiation due to Metchnik & Tomlin (1963[link]) (Fig. 4.2.1.2[link] ). For very small take-off angles, therefore, X-ray tubes should be operated at lower than customary voltages. Note that the values in Fig. 4.2.1.2[link] agree to within ∼40% with those of Green & Cosslett. f(χ) at constant [E_0/E_K] increases with increasing Z, thus partly compensating for the decrease in [N_K], especially at small values of [\varphi]. A recent re-examination of the characteristic X-ray flux from Cr, Cu, Mo, Ag and W targets has been carried out by Honkimaki, Sleight & Suortti (1990[link]).

[Figure 4.2.1.1]

Figure 4.2.1.1| top | pdf |

f(χ) curves for Cu K-L3 at a series of different accelerating voltages (in kV). From Green (1963[link]).

[Figure 4.2.1.2]

Figure 4.2.1.2| top | pdf |

Experimental measurements of [N_\varphi] for Cu K-L3 as functions of the accelerating voltage for different take-off angles. From Metchnik & Tomlin (1963[link]).

4.2.1.2. The continuous spectrum

| top | pdf |

The shape of the continuous spectrum from a thick target is very simple: [I_\nu], the energy per unit frequency band in the spectrum, is given by the expression derived by Kramers (1923[link]): [I_\nu=AZ(\nu_0-\nu)+BZ^2,\eqno (4.2.1.5)]where Z is the atomic number of the target and A and B are constants independent of the applied voltage [E_0]. B/A is of the order of 0.0025 so that the term in [Z^2] can usually be neglected (Fig. 4.2.1.3[link] ). [\nu_0] is the maximum frequency in the spectrum, i.e. the Duane–Hunt limit at which the entire energy of the bombarding electrons is converted into the quantum energy of the emitted photon, where [H\nu_0=hc/\lambda_0=E_0.\eqno (4.2.1.6)]Using the latest adjusted values of the fundamental constants (Cohen & Taylor, 1987[link]): [\eqalign{hc&=1.23984244\pm0.00000037\times10^{-6}\ {\rm eV\ m}\cr&=12.3984244\pm0.0000037 {\rm\ keV}\ {\rm \AA}.}]Equation (4.2.1.5)[link] can be rewritten in a number of forms. If [{\rm d}N_E] is the number of photons of energy E per incident electron, [{\rm d}N_E=bZ(E_0/E-1)\, {\rm d}E,\eqno (4.2.1.7)]where [b\sim2\times10^{-9}] photons eV−1  electron−1, and is known as Kramer's constant.

[Figure 4.2.1.3]

Figure 4.2.1.3| top | pdf |

Intensity per unit frequency interval versus frequency in the continuous spectrum from a thick target at different accelerating voltages. From Kuhlenkampff & Schmidt (1943[link]).

From (4.2.1.7)[link], it follows that the total energy in the continuous spectrum per electron is [\textstyle\int\limits^{E_0}_0E\, {\rm d}N_E=bZE^2_0/2.\eqno (4.2.1.8)]Since the energy of the bombarding electron is [E_0], the efficiency of production of the continuous radiation is [\eta_c=bZE_0/2.\eqno (4.2.1.9)]Crystallographers are more accustomed to thinking of the spectrum in terms of wavelength. Equation (4.2.1.7)[link] can be transformed into [{\rm d}N_\lambda=hcbZ(1/\lambda^2-1/\lambda\lambda_0)\, {\rm d}\lambda,\eqno (4.2.1.10)]which has a maximum at [\lambda=2\lambda_0]. In practice, the emerging spectrum is modified by target absorption, which is greatest for the longer wavelengths and moves the maximum more nearly to [1.5\lambda_0].

It is of interest to compare the X-ray flux in a narrow wavelength band selected by an appropriate monochromator with the flux in a characteristic spectral line, in order to examine the practicability of XAFS (X-ray absorption fine-structure spectroscopy) or optimized anomalous-dispersion diffractometry experiments. For these purposes, the maximum permissible wavelength band is about 10−3 Å. From equation (4.2.1.10)[link], we see that, for a tungsten-target X-ray tube operated at 80 kV, [{\rm d}N_{\lambda}] is about 1.1 × 10−5 photons with the Kα energy electron−1 steradian−1 (10−3 δλ/λ)−1 for an X-ray wavelength in the neighbourhood of 1.5 Å. By comparison, from equation (4.2.1.2)[link], a copper-target tube operated at 40 kV produces about 5 × 10−4Kα photons electron−1 steradian−1. In spite of this shortcoming by a factor of about 45, laboratory XAFS experiments are sufficiently common to have merited at least one specialized conference (Stern, 1980[link]; see also Tohji, Udagawa, Kawasak & Masuda, 1983[link]; Sakurai, 1993[link]; Sakurai & Sakurai, 1994[link]).

The use of continuous radiation for diffraction experiments is complicated by the fact that the radiation is polarized. The degree of polarization may be defined as [p=(I_\|-I_\perp)/(I_\|+I_\perp),\eqno (4.2.1.11)]where [I_\|] and [I_\perp] are the intensities of radiation with the electric vector parallel and perpendicular to the plane containing the incident electrons and the direction of the emitted photons. For an angle of π/2 between the electrons and the emitted beam, p varies smoothly through the spectrum; it is negative for the softest radiation, approximately zero at [\nu/\nu_0\sim0.1] and reaches values between +0.7 and +0.9 near the Duane–Hunt limit (Kirkpatrick & Wiedmann, 1945[link]). Since practical use of white radiation is likely to be in the vicinity of [\nu/\nu_0\sim0.1], the effect is not a large one.

It should also be noted that the spatial distribution of the white spectrum, even after correction for absorption in the target, is not isotropic. The intensity has a maximum at about 50° to the electron beam and non-zero minima at 0 and 180° to that beam (Stephenson, 1957[link]).

4.2.1.3. X-ray tubes

| top | pdf |

The commonest source of X-rays is the high-vacuum, or Coolidge, X-ray tube, which may be either demountable and pumped continuously when in operation or permanently sealed after evacuation. The vacuum tube contains an electron gun that incorporates a thermionic cathode, which produces a well defined electron beam that is accelerated towards the anode or target, formerly often called the anticathode. In most X-ray tubes intended for crystallographic purposes, the anode is massive, i.e. its thickness is large compared with the range of the electrons; it is usually water-cooled and its surface is normal to the incident electron beam. Usually, it is desirable for the X-ray source to be small (between 25 μm and 1 mm square) and for the X-ray intensity from the tube to be the maximum possible for the amount of power that can be dissipated in the target. These objectives are best achieved by designing the electron gun to produce a line focus, that is the electron focus on the target face is approximately rectangular with the small dimension equal to the desired effective source size and the large dimension about 10 to 20 times larger. The focus is viewed at an angle between about 2 and 5° to the anode surface to produce an approximately square foreshortened effective source; and the X-ray windows are so positioned as to make these take-off angles possible. For some purposes, very fine line sources are required and windows may be provided to allow the focus to be viewed so as to foreshorten the line width. Higher power dissipation is possible in X-ray tubes in which the anode rotates: the line focus is now usually on the cylindrical surface of the anode with its long dimension parallel to the axis of rotation.

For focal-spot sizes down to about 100 μm, an electrostatic gun is adequate; this consists of a fine helical filament and a Wehnelt cathode, which produces a demagnified electron image of the filament on the anode. For most purposes, the Wehnelt cathode can be at the same potential as the filament but cleaner foci and adjustment of the focal spot size are possible when this electrode is negatively biased with respect to the filament. The filament is nearly always directly heated and made of tungsten. Lower filament temperatures, and smaller heating currents, could be achieved with activated heaters but the vacuum in high-power devices like X-ray tubes is rarely hard enough to permit their use since they are easily poisoned. However, Yao (1992[link]) has reported successful operation of a hot-pressed polycrystalline lanthanum hexaboride cathode in an otherwise unmodified RU-1000 rotating-target X-ray generator.

Very fine focus tubes, with foci in the range between 25 and 1 μm, require magnetic lenses. At one time, the all-electrostatic X-ray tube of Ehrenberg & Spear (1951[link]), which achieved foci between 20 and 80 μm, was very popular.

Sealed-off X-ray tubes for crystallographic use are nowadays made in the form of inserts containing a target of one of a range of standard metals to produce the desired characteristic radiation. A series of nominal focal-spot sizes, shown in Table 4.2.1.3[link], is commonly available. The insert is mounted inside a standard shield that is radiation- and shock-proof and that is fitted with X-ray shutters and filters and often also with a standardized track for mounting X-ray cameras. The water-cooled anode is normally at ground potential and the negative high voltage for the cathode, together with the filament supply, is brought in through a shielded shock-proof cable. The high voltage is nowadays generally of the constant-voltage type, that is, it is full-wave rectified and smoothed by means of solid-state rectifiers and capacitors housed in the high-voltage transformer tank, which also contains the filament transformer. The high tension and the tube current are frequently stabilized. Only the simplest X-ray generators now employ an alternating high tension that is rectified by the self-rectifying property of the X-ray tube itself.

Table 4.2.1.3| top | pdf |
Copper-target X-ray tubes and their loading

X-ray tubeAnode diameter (mm)Speedf1 × f2
(mm) (mm)
μLoading (kW)Recommended specific loading (kW mm−2)
r min−1mm s−1calc.recommended
Standard insert8 × 0.150.2951.00.80.67
8 × 0.40.3591.21.50.47
10 × 1.00.4251.82.00.20
12 × 2.00.4932.52.70.11
AEI-GX21896000280001 × 0.10.4251.41.212.0
   2 × 0.20.4253.953.28.0
   3 × 0.30.4257.35.25.8
   5 × 0.50.42515.615.06.0
AEI-GX1345745001080001 × 0.10.4252.72.727.0
Rigaku-RU200996000310001 × 0.10.4251.51.212.0
   2 × 0.20.4254.23.07.5
   3 × 0.30.4257.65.46.0
Rigaku-RU50040012502620010 × 0.50.35926.8306.0
Rigaku-RU100040025005245010 × 10.42560606.0
Rigaku-RU15002501000013100010 × 10.42596909.0
KFA-Jülich2501200015700014 × 1.40.4251731206.1

A demountable continuously pumped form of construction is nowadays adopted mainly for rotating-anode and other specialized X-ray tubes. The pumping system must be capable of maintaining a vacuum of better than 10−5 Torr: filament life is critically dependent upon the quality of the vacuum.

Rotating-anode tubes have been reviewed by Yoshimatsu & Kozaki (1977[link]). The first successful tube of this type that incorporated a vacuum shaft seal was described by Clay (1934[link]). Modern tubes mostly contain vacuum-oil-lubricated shaft seals of the type due to Wilson (1941[link]) and are based on, or are similar to, the rotating-anode tubes described by Taylor (1949[link], 1956[link]). In some tubes, successful use has been made of ferro-fluidic vacuum seals (see Bailey, 1978[link]). The main problems in the operation of rotating-anode tubes is the lifetime of the seals and of bearings that operate in vacuo. In successful tubes, e.g. those manufactured by Enraf, Rigaku-Denki, and Siemens, these lifetimes are about the same as the lifetime of the filament under good vacuum conditions, that is, of the order of 1000 h.

Phillips (1985[link]) has written a review article on stationary and rotating-anode X-ray tubes that contains many important practical details.

4.2.1.3.1. Power dissipation in the anode

| top | pdf |

The allowable power loading of X-ray tube targets is determined by the temperature of the target surface, which must remain below the melting point. Müller (1927[link], 1929[link], 1931[link]) first calculated the maximum loading both for stationary and for rotating anodes. His calculations were refined by Oosterkamp (1948a[link],b[link],c[link]) who considered, in particular, targets of finite thickness, and who also treated pulsed operation of the tube. For normal conditions, Oosterkamp's conclusions and those of Ishimura, Shiraiwa & Sawada (1957[link]) do not greatly differ from those of Müller, which are in adequate agreement with experimental observations.

For an elliptical focal spot with axes [f_1] and [f_2], Müller's formula for the maximum power dissipation on a stationary anode, assumed to be a water-cooled block of dimensions large compared with the focal-spot dimensions, can be written [W_{\rm stat}=2.063(T_M-T_0)Kf_1\mu(\,f_1,f_2),\eqno (4.2.1.12)]where K is the specific thermal conductivity of the target material in W mm−1, [T_M] is the maximum temperature at the centre of the focal spot on the target, that is, a temperature well below the melting point of the target material, and [T_0] is the temperature of the cold surface of the target, that is, of the cooling water. The function μ is shown in Fig. 4.2.1.4[link] . For copper, K is 400 W m−1 and, with [T_M-T_0] = 500 K, [W_{\rm stat}=425\mu f_1.\eqno (4.2.1.13)]For f2/f1 = 0.1, and μ = 0.425, this equation becomes [W_{\rm stat}=180 \, f_1.\eqno (4.2.1.14)]In these last two equations, f1 is in mm.

[Figure 4.2.1.4]

Figure 4.2.1.4| top | pdf |

The function μ in Müller's equation (equation 4.2.1.12[link]) as a function of the ratio of width to length of the focal spot.

For a rotating target, Müller found that the permissible power dissipation was given by [W_{\rm rot}=1.428\,K(T_M-T_0)\,f_1(\,f_2\rho Cv/2K)^{1/2},\eqno (4.2.1.15)]where f2 is the short dimension of the focus, assumed to be in the direction of motion of the target, v is the linear velocity, ρ is the density of the target material, and C is its specific heat.

For a copper target with f1 and f2 in mm and v in mm s−1, [W_{\rm rot}=26.4\,f_1(\,f_2v)^{1/2}.\eqno (4.2.1.16)]Equation (4.2.1.16)[link] shows that for very narrow focal spots rotating-anode tubes give useful improvements in permissible loading only if the surface speed is very high (see Table 4.2.1.3[link]). The reason is that with large foci on stationary anodes the isothermal surfaces in the target are planar; with fine foci, these surfaces become cylindrical and this already makes for very efficient cooling without the need for rotation. Rotating anodes are thus most useful for medium-size foci (200 to 500 μm) since for the larger focal spots it becomes very expensive to construct power supplies capable of supplying the permissible amount of power.

Table 4.2.1.3[link] shows the recommended loading for a number of commercially available X-ray tubes with copper targets, which will be seen to be in qualitative agreement with the calculations. Some of the discrepancy is due to the fact that the value of [K(T_M-T_0)] for the copper–chromium alloy targets used in actual X-ray tubes is appreciably lower than the value for pure copper used here. To a good approximation, the permissible loading for other targets can be derived by multiplying those in Table 4.2.1.3[link] by the factors shown in Table 4.2.1.4[link]. It is worth noting that the recommended loading of commercial stationary-target X-ray tubes has increased steadily in recent years. This is largely due to improvements in the water cooling of the back surface of the target by increasing the turbulence of the water and the effective surface area of the cooled surface.

Table 4.2.1.4| top | pdf |
Relative permissible loading for different target materials

CuCrFeCoMoAgW
1.00.90.60.91.21.01.2

In considering Table 4.2.1.3[link], it should be noted that the linear velocities of the highest-power X-ray-tube anode have already reached a speed that Yoshimatsu & Kozaki (1977[link]) consider the practical limit, which is set by the mechanical properties of engineering materials. It should also be noted that much higher specific loads can be achieved for true micro-focus tubes, e.g. 50 kW mm−2 for a 25 μm Ehrenberg & Spear tube and 1000 kW mm−2 for a tube with a 1 μm focus (Goldsztaub, 1947[link]; Cosslett & Nixon, 1951[link], 1960[link]).

Some tubes with focus spots of less than 10 μm utilize foil or needle targets. These targets and the heat dissipated in them have been discussed by Cosslett & Nixon (1960[link]). The dissipation is less than that in a massive target by a factor of about 3 for a foil and 10 for a needle, but, in view of the low absolute power, target movement and even water-cooling can be dispensed with.

4.2.1.4. Radioactive X-ray sources

| top | pdf |

Radioactive sources of X-rays are mainly of interest to crystallographers for the calibration of X-ray detectors where they have the great advantage of being completely stable with time, or at least of having an accurately known decay rate. For some purposes, spectral purity of the radiation is important; radionuclides that decay wholly by electron capture are particularly useful as they produce little or no β or other radiation. In this type of decay, the atomic number of the daughter nucleus is one less than that of the decaying isotope, and the emitted X-rays are characteristic of the daughter nucleus. In some cases, the probability of electron capture taking place from some shell other than the K shell is very small and most of the photons emitted are K photons. The number of photons emitted into a solid angle of 4π, uncorrected for absorption, is given by the strength of the source in Curies (1 Curie = 3.7 × 1010 disintegrations s−1), since each disintegration produces one photon. A list of these nuclei (after Dyson, 1973[link]) is given in Table 4.2.1.5[link].

Table 4.2.1.5| top | pdf |
Radionuclides decaying wholly by electron capture, and yielding little or no γ-radiation

NuclideHalf-lifeX-raysRemarks
Element[K{\alpha_1}] (keV)
37Ar35 dCl2.622
51Cr 27.8 dV4.952γ at 320 keV
55Fe2.6 aMn5.898
71Ge11.4 dGa9.251
103Pd17 dRh20.214Several γ's; all weak
109Cd453 dAg22.16γ at 88 keV
125I60 dTe27.47γ at 35.4 keV
131Cs10 dXe29.80
145Pm17.7 aNd37.36γ's at 67 and 72 keV
145Sm340 dPm38.65γ's at 61 keV; weak γ at 485 keV
179Ta600 dHf55.76
181W140 dTa57.52γ at 6.5 keV; weak γ's at 136, 153 keV
205Pb5 × 107 aTlL only (Lα1 = 10.27 keV)

Useful radioactive sources are also made by mixing a pure β-emitter with a target material. These sources produce a continuous spectrum in addition to the characteristic line spectrum. The nuclide most commonly used for this purpose is tritium which emits β particles with an energy up to 18 keV and which has a half-life of 12.4 a.

Radioactive X-ray sources have been reviewed by Dyson (1973[link]).

4.2.1.5. Synchrotron-radiation sources

| top | pdf |

The growing importance of synchrotron radiation is attested by a large number of monographs (Kunz, 1979[link]; Winick, 1980[link]; Stuhrmann, 1982[link]; Koch, 1983[link]) and review articles (Godwin, 1968[link]; Kulipanov & Skrinskii, 1977[link]; Lea, 1978[link]; Winick & Bienenstock, 1978[link]; Helliwell, 1984[link]; Buras, 1985[link]). Project studies for storage rings such as the European Synchrotron Radiation Facility, the ESRF (Farge & Duke, 1979[link]; Thompson & Poole, 1979[link]; Buras & Marr, 1979[link]; Buras & Tazzari, 1984[link]) are still worth consulting for the reasoning that lay behind the design; the ESRF has, in fact, achieved or even exceeded the design parameters (Laclare, 1994[link]).

A charged particle with energy E and mass m moving in a circular orbit of radius R at a constant speed v radiates a power P into a solid angle of 4π, where [P=2e^2c(v/c)^4(E/mc^2)^4/3R^2.\eqno (4.2.1.17)]The orbit of the particle can be maintained only if the energy lost in the form of electromagnetic radiation is constantly replenished. In an electron synchrotron or in a storage ring, the circulating particles are electrons or positrons maintained in a closed orbit by a magnetic field; their energy is supplied or restored by means of an oscillating radio-frequency (RF) electric field at one or more places in the orbit. In a synchrotron, designed for nuclear-physics experiments, the circulating particles are injected from a linear accelerator, accelerated up to full energy by the RF field and then deflected into a target with a cycle frequency of about 50 Hz. The synchrotron radiation is thus produced in the form of pulses of this frequency. A storage ring, on the other hand, is filled with electrons or positrons and after acceleration the particle energy is maintained by the RF field; the current ideally circulates for many hours and decays only as a result of collisions with remaining gas molecules. At present, only storage rings are used as sources of synchrotron radiation and many of these are dedicated entirely to the production of radiation: they are not used at all, or are used only for limited periods, for nuclear-physics collision experiments.

In equation (4.2.1.17)[link], we may substitute for the various constants and obtain for the radiated power [P=0.0885\,E^4I/R,\eqno (4.2.1.18)]where E is in GeV (109 eV), I is the circulating electron or positron current in milliamperes, and R is in metres. Thus, for example, at the Daresbury storage ring in England, R = 5.5 m and, for operation at 2 GeV and 200 mA, P = 51.5 kW. Storage rings with a total power of the order of 1 MW are planned.

For relativistic electrons, the electromagnetic radiation is compressed into a fan-shaped beam tangential to the orbit with a vertical opening angle [\psi\simeq mc^2/E], i.e. ~0.25 mrad for E = 2 GeV (Fig. 4.2.1.5[link] ). This fan rotates with circulating electrons: if the ring is filled with n bunches of electrons, a stationary observer will see n flashes of radiation every 2πR/c s, the duration of each flash being less than 1 ns.

[Figure 4.2.1.5]

Figure 4.2.1.5| top | pdf |

Synchrotron radiation emitted by a relativistic electron travelling in a curved trajectory. B is the magnetic field perpendicular to the plane of the electron orbit; ψ is the natural opening angle in the vertical plane; P is the direction of polarization. The slit S defines the length of the arc of angle Δθ from which the radiation is taken. From Buras & Tazzari (1984[link]); courtesy of ESRP.

The spectral distribution of synchrotron radiation extends from the infrared to the X-ray region; Schwinger (1949[link]) gives the instantaneous power radiated by a monoenergetic electron in a circular motion per unit wavelength interval as a function of wavelength (Winick, 1980[link]). An important parameter specifying the distribution is the critical wavelength [\lambda_c]: half the total power radiated, but only ∼9% of the total number of photons, is at [\lambda \lt \lambda_c] (Fig. 4.2.1.6[link] ). [\lambda_c] is given by [\lambda_c=4\pi R/3(E/mc^2){}^3,\eqno (4.2.1.19)]from which it follows that [\lambda_c] in Å can be expressed as [\lambda_c=18.64/(BE^2),\eqno (4.2.1.20)]where B (= 3.34 E/R) is the magnetic bending field in T, E is in GeV, and R is in metres.

[Figure 4.2.1.6]

Figure 4.2.1.6| top | pdf |

Synchrotron-radiation spectrum: percentage per unit wavelength interval (a) of power of total power and (b) of number of photons of total number of photons at wavelengths greater than λ versus λ/λc. Note that half the power but only 9% of the photons are radiated at wavelengths less than λc; courtesy of H. Winick.

Synchrotron radiation is highly polarized. In an ideal ring where all electrons are parallel to one another in a central orbit, the radiation in the orbital plane is linearly polarized with the electric vector lying in this plane. Outside the plane, the radiation is elliptically polarized.

In practice, the electron path in a storage ring is not a circle. The `ring' consists of an alternation of straight sections and bending magnets and beam lines are installed at these magnets. So-called insertion devices with a zero magnetic field integral, i.e. wigglers and undulators, may be inserted in the straight sections (Fig. 4.2.1.7[link] ). A wiggler consists of one or more dipole magnets with alternating magnetic field directions aligned transverse to the orbit. The critical wavelength can thus be shifted towards shorter values because the bending radius can be made small over a short section, especially when superconducting magnets are used. Such a device is called a wavelength shifter. If it has N dipoles, the radiation from the different poles is added to give an N-fold increase in intensity. Wigglers can be horizontal or vertical.

[Figure 4.2.1.7]

Figure 4.2.1.7| top | pdf |

Main components of a dedicated electron storage-ring synchrotron-radiation source. For clarity, only one bending magnet is shown. From Buras & Tazzari (1984[link]); courtesy of ESRP.

In a wiggler, the maximum divergence 2α of the electron beam is much larger than ψ, the vertical aperture of the radiation cone in the spectral region of interest (Fig. 4.2.1.5[link]). If [2\alpha\ll\psi] and if, in addition, the magnet poles of a multipole device have a short period [\lambda_0], the device becomes an undulator: interference will take place between the radiation emitted at two points [\lambda_0] apart on the electron trajectory (Fig. 4.2.1.8[link] ). The spectrum at an angle [\varphi] to the axis observed through a pin-hole will consist of a single spectral line and its harmonics of wavelengths [\lambda_i=i^{-1}\lambda_0[(E/mc^2)^{-2}+\alpha^{2}/2+\theta^{2}]/2\eqno (4.2.1.21)](Hofmann, 1978[link]). Typically, the bandwidth of the lines, δλ/λ, will be ∼0.01 to 0.1 and the photon flux per unit band width from the undulator will be many orders of magnitude greater than that from a bending magnet. Existing undulators have been designed for photon energies below 2 keV; higher energies, because of the relatively weak magnetic fields necessitated by the need to keep λ0 small [equation (4.2.1.21)[link]], require a high electron energy: undulators with a fundamental wavelength in the neighbourhood of 0.86 Å are planned for the European storage ring (Buras & Tazzari, 1984[link]).

[Figure 4.2.1.8]

Figure 4.2.1.8| top | pdf |

Electron trajectory within a multipole wiggler or undulator. λ0 is the spatial period, α the maximum deflection angle, and θ the observation angle. From Buras & Tazzari (1984[link]); courtesy of ESRP.

The wavelength spectra for a bending magnet, a wiggler and an undulator for the ESRF, are shown in Fig. 4.2.1.9[link] . A comparison of the spectra from an existing storage ring with the spectrum of a rotating-anode tube is shown in Fig. 4.2.1.10[link] .

[Figure 4.2.1.9]

Figure 4.2.1.9| top | pdf |

Spectral distribution and critical wavelengths for (a) a dipole magnet, (b) a wavelength shifter, and (c) a multipole wiggler for the proposed ESRF. From Buras & Tazzari (1984[link]); courtesy of ESRP.

[Figure 4.2.1.10]

Figure 4.2.1.10| top | pdf |

Comparison of the spectra from the storage ring SPEAR in photons s−1 mA−1 mrad−1 per 1% passband (1978 performance) and a rotating-anode X-ray generator. From Nagel (1980[link]); courtesy of K. O. Hodgson.

The important properties of synchrotron-radiation sources are:

  • (1) high intensity;

  • (2) very broad continuous spectral range;

  • (3) narrow angular collimation;

  • (4) small source size;

  • (5) high degree of polarization;

  • (6) regularly pulsed time structure;

  • (7) computability of properties.

Table 4.2.1.6[link] (after Buras & Tazzari, 1984[link]) compares the most important parameters of the European Synchrotron Radiation Facility (ESRF) with a number of other storage rings. In this table, BM and W signify bending-magnet and wiggler beam lines, respectively, [\sigma_x] and [\sigma'_z] is the source divergence; the flux is integrated in the vertical plane. The ESRF is seen to have a higher flux than other sources; even more impressive by virtue of the small dimensions of the source size and divergence are its improvements in spectral brightness (defined as the number of photons s−1 per unit solid angle per 0.1% bandwidth) and in spectral brilliance (defined as the number of photons s−1 per unit solid angle per unit area of the source per 0.1% bandwidth). In comparing different synchrotron-radiation sources with one another and with conventional sources (Fig. 4.2.1.10[link]), the relative quantity for comparison may be flux, brightness or brilliance, depending on the type of diffraction experiment and the type of collimation adopted. Table 4.2.1.7[link] (due to Farge & Duke, 1979[link]) attempts to compare intensity factors for a number of typical experiments. In general, a high brightness is important in experiments that do not embody focusing elements, such as mirrors or curved crystals, and a high brilliance in those experiments that do.

Table 4.2.1.6| top | pdf |
Comparison of storage-ring synchrotron-radiation sources; the parameters were correct in 1985 and, for some sources, may be substantially different from those at earlier or later periods; after Buras & Tazzari (1984[link]), courtesy of ESRP

Storage ringSource typeNo. of polesI (mA)E (GeV)R (m)[\sigma_x] (mm)[\sigma_z] (mm)[\sigma'_z] (mrad)[\lambda_c] (Å)Ec (keV)[{\rm Flux} \Big[{{\rm photons \, \, s}^{-1}\over {\rm mrad}\times 0.1\%\, \, {\rm BW}}\Big]]
at [\lambda^c]at 1.54 Å
(1) ESRFBM1005.020.00.0920.1000.0080.9148 × 10121 × 1013
(2) ESRFW301005.011.560.0620.0400.0160.5242.4 × 10143 × 1014
(3) ADONE (Frascati) BM1001.55.00.80.40.048.01.52.4 × 10125 × 1010
(4) ADONE (Frascati)W61001.52.61.40.240.084.331.4 × 10133.4 × 1012
(5) SRS (Daresbury)BM3002.05.562.70.230.054.031 × 10133 × 1012
(6) SRS (Daresbury)W13002.01.335.30.170.050.9131 × 10131.2 × 1013
(7) DCI (Orsay)BM2501.83.822.721.060.063.63.47 × 10122.4 × 1012
(8) DORIS (Hamburg)BM1003.712.221.00.30.051.39.26 × 10126.4 × 1012
(9) DORIS (Hamburg)BM405.012.221.30.650.0650.55233 × 10124.4 × 1013
(10) DORIS (Hamburg)W32 1003.720.571.50.40.0332.35.51.9 × 10141.3 × 1014
(11) CESR (Cornell)BM405.532.01.441.00.0651.011.53.5 × 10124 × 1012
(12) CESR (Cornell)W6405.513.21.91.20.050.4282 × 10133 × 1013
(13) NSLS X-ray (Brookhaven)BM3002.56.830.250.10.012.451 × 10138 × 1012
(14) SPEARBM1003.012.72.00.280.052.755 × 10123 × 1012
(15) SPEARW81003.05.573.20.150.031.0103.8 × 10134.5 × 1013
(16) SPEARW541003.08.363.20.150.031.772.6 × 10142.4 × 1014
(17) Photon Factory (Tsukuba)BM1502.58.662.20.60.143.046 × 10123 × 1012
(18) Photon Factory (Tsukuba) W31502.51.851.90.70.180.7191.8 × 10132.5 × 1013
(19) VEPP-3BM1002.26.156.150.080.023.043.5 × 10121.5 × 1012
One standard deviation of Gaussian distribution.

Table 4.2.1.7| top | pdf |
Intensity gain with storage rings over conventional sources; from Farge & Duke (1979[link]), courtesy of ESF

 GX6 rotating-anode tube 2.4 kW (Cu Kα emission)DCI 1.72 GeV and 240 mAESRF 5 GeV and 565 mA
[Scheme scheme1]Small-angle scattering with a double monochromator×500 to 1000×15000 to 3000
Protein crystallography with a single-focus monochromator  
  1 mm3 samples×50 to 160×900 to 1800
  Small samples×30 to 60×650 to 1300
Diffuse scattering (wide angles, low resolution and large samples) with a curved graphite monochromator×20 to 40×160 to 320
Non characteristic wavelength (continuous background) EXAFS experimental set-up with a 100 kW rotating anode×104×105

Many surveys of existing and planned synchrotron-radiation sources have been published since the compilation of Table 4.2.1.6[link]. Fig. 4.2.1.11[link] , taken from a recent review (Suller, 1992[link]), is a graphical illustration of the growth and the distribution of these sources. An earlier census is due to Huke & Kobayakawa (1989[link]). Many detailed descriptions of beam lines for particular purposes, such as protein crystallography (e.g. Fourme, 1992[link]) or at individual storage rings (e.g. Kusev, Raiko & Skuratowski, 1992[link]) have appeared: these are too numerous to list here and can be located by reference to Synchrotron Radiation News.

[Figure 4.2.1.11]

Figure 4.2.1.11| top | pdf |

The evolution of storage-ring synchrotron-radiation sources over the decades, as illustrated by their increasing number and range of machine energies (based on Suller, 1992[link]).

4.2.1.6. Plasma X-ray sources

| top | pdf |

Plasma sources of hard X-rays are being investigated in many laboratories. Most of the material in this section is derived from publications from the Laboratory for Laser Energetics, University of Rochester, USA. Plasma sources of very soft X-rays have been reviewed by Byer, Kuhn, Reed & Trail (1983[link]).

The peak wavelength of emission from a black-body radiator falls into the ultraviolet at about 105 K and into the X-ray region between 106 and 107 K. At these temperatures, matter is in the form of a plasma that consists of highly ionized atoms and of electrons with energies of several keV. The only successful methods of heating plasmas to temperatures in excess of 106 K is by means of high-energy laser beams with intensities of 1012 W mm−2 or more. The duration of the laser pulse must be less than 1 ns so that the plasma cannot flow away from the pulse. When the plasmas are created from elements with 15 [\lt] Z [\lt] 25, they consist mainly of ions stripped to the K shell, that is of hydrogen- and helium-like ions. The X-ray spectrum (Fig. 4.2.1.12[link] ) then contains a main group of lines with a bandwidth for the group of about 1%; the band is situated slightly below the K-absorption edge of the target material. The intensity of the band drops with increasing atomic number. For diffraction studies, Forsyth & Frankel (1980[link], 1984[link]) and Frankel & Forsyth (1979[link], 1985[link]) used a multi-stage Nd3+:glass laser (Seka, Soures, Lewis, Bunkenburg, Brown, Jacobs, Mourou & Zimmermann, 1980[link]), which was able to deliver up to 220 J per pulse of width 700 ps. They obtained [6\times10^{14}] photons pulse−1 for a Cl15+ plasma with a mean wavelength of about 4.45 Å and about [3\times10^{13}] photons pulse−1 for a Fe24+ plasma at about 1.87 Å (Yaakobi, Bourke, Conturie, Delettrez, Forsyth, Frankel, Goldman, McCrory, Seka, Soures, Burek & Deslattes, 1981[link]). More recently, the laser was fitted with a frequency conversion system that shifts the peak power of the laser light from the infrared (1.054 µm) to the ultraviolet (0.351 µm) (Seka, Soures, Lund & Craxton 1981[link]). This led to a more efficient X-ray production, which permitted a more than twofold increase in X-ray flux, even though the maximum pulse energies had to be reduced to ∼50 J to prevent damage to the optical components (Yaakobi, Boehli, Bourke, Conturie, Craxton, Delettrez, Forsyth, Frankel, Goldman, McCrory, Richardson, Seka, Shvarts & Soures, 1981[link]). Forsyth & Frankel (1984[link]) used the plasma X-ray source for diffraction studies with 4.45 Å X-rays with a focusing collimation system that delivered up to 1010 photons pulse−1 to the specimen over an area approximately 150 µm in diameter. More recently, by special target design (Forsyth, 1986, unpublished), fluxes have been increased by factors of 2 to 3 without altering the laser output. Other plasma sources have been described by Collins, Davanloo & Bowen (1986[link]) and by Rudakov, Baigarin, Kalinin, Korolev & Kumachov (1991[link]).

[Figure 4.2.1.12]

Figure 4.2.1.12| top | pdf |

X-ray emission from various laser-produced plasmas. From Forsyth & Frankel (1980[link]); courtesy of J. M. Forsyth.

The cost of plasma sources is about an order of magnitude greater than that of rotating-anode generators (Nagel, 1980[link]). Their use is at present confined to flash-diffraction experiments, since the duty cycle is a maximum of one flash every 30 min. Attempts are being made to increase the laser repetition rate; a substantial improvement could lead to a source that would rival storage-ring sources.

4.2.1.7. Other sources of X-rays

| top | pdf |

Parametric X-ray generation can be described as the diffraction of virtual photons associated with the field of a relativistic charged particle passing through a crystal. These diffracted photons appear as real photons with an energy that satisfies Bragg's law for the reflecting crystal planes, so that the energy can be tuned between 5 and 45 keV by rotating the mosaic graphite crystal. Linear accelerators with an energy between 100 and 500 MeV produce the incident relativistic electron beam (Maruyama, Di Nova, Snyder, Piestrup, Li, Fiorito & Rule, 1993[link]; Fiorito, Rule, Piestrup, Li, Ho & Maruyama, 1993[link]).

Transition-radiation X-rays with peak energies between 10 and 30 keV are produced when electrons from 100 to 400 MeV strike a stack of thin foils (Piestrup, Moran, Boyers, Pincus, Kephart, Gearhart & Maruyama, 1991[link]). Quasi-monochromatic X-rays result from a selection of target foils with appropriate K-, L- or M-edge frequencies (Piestrup, Boyers, Pincus, Harris, Maruyama, Bergstrom, Caplan, Silzer & Skopik, 1991[link]).

Channelling radiation, resulting from the incidence of electrons with an energy of only about 5 MeV on appropriately aligned diamond or silicon crystals hold out the hope of producing a bright tunable X-ray source.

One or more of these methods may, in the future, be developed as X-ray sources that can compete with synchrotron-radiation sources.

4.2.2. X-ray wavelengths

| top | pdf |
R. D. Deslattes,c E. G. Kessler Jr,f P. Indelicatoe and E. Lindrothg

4.2.2.1. Historical introduction

| top | pdf |

Wavelength tables in previous editions of this volume (Rieck, 1962[link]; Arndt, 1992[link]) were mainly obtained from the compilations prepared in Paris under the general direction of Professor Y. Cauchois (Cauchois & Hulubei, 1947[link]; Cauchois & Senemaud, 1978[link]). A separate effort by the late Professor J. A. Bearden and his collaborators (Bearden, 1967[link]) has been widely used in other aggregations of tabular data and was made available for some time through the Standard Reference Data Program at the National Institute of Standards and Technology (NIST). For simplicity in the following discussion, we use the Bearden database as a frame of reference with respect to which our current, and rather different, approach can be compared. Although a detailed comparison of the historical databases may be of some interest, the result would have only very small influence on the outcome presented here. To specify this framework, we begin with a brief description of the procedures used in establishing this reference database.

Bearden and his collaborators remeasured a group of five X-ray lines (Bearden, Henins, Marzolf, Sauder & Thomsen, 1964[link]), with the remaining entries in the wavelength table coming from a critically reviewed, and re-scaled, subset of earlier measurements (Bearden, 1967[link]). Line locations were given in Å* units, a scale defined by setting the wavelength of W Kα1 = 0.2090100 Å*. It was Bearden's intention that, for all but the most demanding applications, one could simply assign Å*/Å = 1, with an uncertainty arising from the fundamental physical constants, particularly NA and hc/e, combined with uncertainties arising from the measurement technology (Bearden, 1965[link]). Not long after the publication of the final compilation (Bearden, 1967[link]), it became clear that the fundamental constants used in defining Å* needed significant revision (Cohen & Taylor, 1973[link]), and that there were some inconsistencies in the metrology (Kessler, Deslattes & Henins, 1979[link]).

4.2.2.2. Known problems

| top | pdf |

Aside from the particular issues noted above, all previous wavelength tables had certain limitations arising from the procedures used in their generation. In particular, except for a small group of five [K\alpha ] spectra (Bearden, Thomsen et al., 1964[link]), the Bearden tables relied entirely on data previously reported in the literature. Both of the other tabulations also proceeded using only reported experimental values (Cauchois & Hulubei, 1947[link]; Cauchois & Senemaud, 1978[link]). In the Bearden compilation process, available data for each emission line were weighted according to claimed uncertainties, modified in certain cases by Bearden's detailed knowledge of the measurement practices of the major sources of experimental wavelength values. The complete documentation of this remarkable undertaking is, unfortunately, not widely accessible. Our evident need to understand the origin of the `recommended' values has been greatly aided by the availability of a copy of the full documentation (Bearden, Thomsen et al., 1964[link]).

The actual experimental data array from which the previous tables emerged is not complete, even for the prominent (`diagram') lines. In the cases where experimental data were not available [as can be seen only in the source documentation (Bearden, Thomsen et al., 1964[link])], the gaps were filled by interpolated values based on measurements available from nearby elements, plotted on a modified Moseley diagram in which the [Z^{2}] term dependence is taken into account (Burr, 1996[link]). In the end, such a smooth scaling with respect to nuclear charge suppresses the effects of the atomic shell structure, a practice that must be avoided in order to obtain the significant improvement in the database that we hope to provide. Also obscured in smooth Z scaling are detectable contributions arising from the fact that nuclear sizes do not change smoothly as a function of the nuclear charge, Z.

4.2.2.3. Alternative strategies

| top | pdf |

There are several possible approaches to generating an improved, `all-Z' table of X-ray wavelengths. These range from the option of conducting a massive measurement campaign to populate more fully the currently available tabular array to a large computational endeavor that might purport to carry out multiconfiguration, relativistic wavefunction calculations for the entire Periodic Table. It seems evident to us that there is little interest in, and even less support for, mounting the large effort needed to realize an improved tabulation of X-ray wavelengths by purely experimental means, while the possibility of proceeding in an entirely theoretical mode is not consistent with the evident need that at least some wavelengths be reported with uncertainties that approach the limit of what can be obtained from the naturally occurring X-ray lines. The actual location of any useful feature of a line is influenced not only by the physical and chemical environment of the emitting atom but also by inevitable multi-electron excitation processes that perturb the entire spectral profile. Calculation of such complexities currently lies beyond the limits of practicality, eliminating the option of proceeding without strong coupling to experimental profile locations, at least for crystallographically important X-ray lines. Similar considerations apply a fortiori to those lines needed as reference wavelengths for exotic atom measurements, such as those leading to masses of elementary particles and tests of basic theory [see e.g. Beyer, Indelicato, Finlayson, Liesen & Deslattes (1991[link])].

In constructing the accompanying tables, we have chosen a new procedure that differs from those described above, and accordingly requires some detailed commentary. We begin with the presently available network of well documented experimental measurements, originally established to provide a test bed for the theoretical methods developing at that time (Deslattes & Kessler, 1985[link]). This modest network was the first compilation to make use of the, then newly available, connection between the X-ray region and the base unit of the International System of measurement (the SI) based on optical interferometric measurement of a lattice period as revealed by X-ray interferometry. Details of the generation of this network and its subsequent expansion will be given below. Using this network as a test set gave clearer suggestions as to specific limitations of the theoretical modelling than had been evident from using other, less selective, experimental reference compilations available at that time. Extensive theoretical developments before and, especially, after the appearance of this new experimental reference set have shown a steady convergence toward these critically evaluated data. Following this evolution further, our long-term plan is to use these new theoretical calculations to provide a more structured and accurate interpolation procedure for estimating the spectra of elements lying between those for which we have accurate measurements, or spectra well connected to a directly established reference wavelength. The present table provides experimental and theoretical values for some of the more prominent K and L series lines and is a subset of a larger effort for all K and L series lines connecting the n = 1 to n = 4 shells. The more complete table will be published elsewhere and be made available on the the NIST Physical Reference Data web site. In addition, experimental values for the K and L edges are provided. Although the reference data are inadequate in both low and high ranges of Z, the general consistency of theory and experiment through the region 20 [\lt] Z [\lt] 90 for the strong K-series and L-series lines suggests that, in the absence of good reference measurements, the uncorrected theoretical values should be considered for applications not requiring the highest accuracy.

4.2.2.4. The X-ray wavelength scales, old and new

| top | pdf |

Historically, from the first realizations of refined spectroscopy in the X-ray region (ca 1915–1925) up to the period 1975–1985, the best measured X-ray wavelengths had to be expressed in some local unit, most often designated as the xu (x unit) or kxu (kilo x unit). Uncertainty in the conversion factor between the X-ray and optical scales was the dominant contributor to the total uncertainties in the wavelength values of the sharper X-ray emission lines, such as those most frequently used in crystallography. (For a discussion of the present values in relation to previously assigned numerical values on the various scales, see Subsection 4.2.2.14.[link]) This local unit was, for most of the time, `officially' defined by assigning a specific numerical value to the lattice period of a particular reflection from `the purest instance' of a particular crystal. Originally this was rocksalt; later it was calcite. In practice, most work used as de facto standards certain values for Cu [K\alpha _{1}] and Mo [K\alpha _{1}] whose inconsistency, though noted by some crystallographers earlier, was seriously addressed by Bearden and co-workers only in the 1960's (Bearden, Henins et al., 1964[link]). This early history was summarized in 1968 (Thomsen & Burr, 1968[link]). Connection of the X-ray wavelength scale to the primary realizations of the length (wavelength) unit in the `metric system' was primarily (at least after about 1930) through ruled grating measurements of longer-wavelength X-ray lines such as Al [K\alpha _{1,2}].

The remainder of the X-ray wavelength database was derived from relative measurements using crystal diffraction spectroscopy. Unfortunately, even the most refined among the ruled grating measurements did not give accuracies comparable to the precision accessible by relative wavelength measurements (Henins, 1971[link]). As noted above, in connection with establishing the previous wavelength table, Bearden introduced a new local unit, the Å*, based on an explicit value for the wavelength of W [K\alpha _{1}], chosen to give a conversion factor near unity. This transitional period will not be treated further in the present documentation since, to a substantial extent, developments described in the following paragraphs have effectively eliminated the need for a local scale for X-ray wavelength metrology.

Following the demonstration of crystal lattice interferometry in the X-ray region (Bonse & Hart, 1965a[link]), efforts to combine such an X-ray interferometer with various optical interferometers were undertaken in several (mostly national standards) laboratories. Although the earliest of these, carried out at the National Bureau of Standards (NBS) (now the National Institute of Standards and Technology, NIST) (Deslattes & Henins, 1973[link]) was, in the end, found to be burdened by a serious systematic error (1.8 × 10−6) in later work at the Physikalisch Technishe Bundesanstalt (PTB) (Becker et al., 1981[link]; Becker, Seyfried & Siegert, 1982[link]), it was clear that accuracy limitations associated with ruled grating measurements no longer dominated the metrology of X-ray wavelengths. The origin of the systematic error in the early NBS measurement was subsequently understood (Deslattes, Tanaka, Greene, Henins & Kessler, 1987[link]), and, more recently, excellent results were obtained in Italy (Basile, Bergamin, Cavagnero, Mana, Vittone & Zosi, 1994[link], 1995[link]) and Japan (Fujimoto, Fujii, Tanaka & Nakayama, 1997[link]). In all cases, the goal was to obtain an optical measurement of a crystal lattice period (thus far, only Si 220) and to use the calibrated crystals in diffraction spectrometry to establish optically based X-ray wavelengths. Such exercises have been undertaken for several X-ray lines, but the most detailed and well documented results to date were obtained in Jena (Härtwig, Hölzer, Wolf & Förster, 1993[link]; Hölzer, Fritsch, Deutsch, Härtwig & Förster, 1997[link]), where the Kα and Kβ spectra of the elements Cr to Cu were evaluated using silicon crystals well connected with the crystal spacings measured at the PTB.

4.2.2.5. K-series reference wavelengths

| top | pdf |

In addition to the Jena measurements noted above, a number of characteristic X-ray lines were measured on the optically based scale at NBS/NIST. The set of directly measured reference wavelengths is given in Table 4.2.2.1[link] in bold type. Most of the originally published (NBS/NIST) values were burdened by the 1.8 × 10−6 error in the silicon lattice period, as noted above. These have been corrected in the numerical results summarized in the table. The directly measured elements and lines appearing in this table were often chosen to meet the need for specific reference values in locations near those of certain optical transitions in highly charged ion spectra or spectra from pionic atoms. In addition, early NBS measurements specifically addressed the lines most often used in crystallography and the W Kα transition. In response to the needs of electron spectroscopy, Al and Mg K spectra were also determined (Schweppe, Deslattes, Mooney & Powell, 1994[link]). In this case, the original lattice error had been previously recognized, so that no rescaling was required. The remaining directly measured entries were obtained as opportunities to do so emerged.

Table 4.2.2.1| top | pdf |
K-series reference wavelengths in Å; bold numbers indicate a directly measured line

Numbers in parentheses are standard uncertainties in the least-significant figures.

ZSymbolA[K\alpha _2][K\alpha _1][K\beta _3][K\beta _1]References
12Mg 9.89153 (10)9.889554 (88)  (a)
13Al 8.341831 (58)8.339514 (58)  (a)
14Si 7.12801 (14)7.125588 (78)  (b)
16S 5.374960 (89)5.372200 (78)  (b)
17Cl 4.730693 (71)4.727818 (71)  (b)
18Ar 4.194939 (23)4.191938 (23)  (c)
19K 3.7443932 (68)3.7412838 (56)  (d)
24Cr 2.2936510 (30)2.2897260 (30)2.0848810 (40)2.0848810 (40)(e)
25Mn 2.1058220 (30)2.1018540 (30)1.9102160 (40)1.9102160 (40)(e)
26Fe 1.9399730 (30)1.9360410 (30)1.7566040 (40)1.7566040 (40)(e)
27Co 1.7928350 (10)1.7889960 (10)1.6208260 (30)1.6208260 (30)(e)
28Ni 1.6617560 (10)1.6579300 (10)1.5001520 (30)1.5001520 (30)(e)
29Cu 1.54442740 (50)1.54059290 (50)1.3922340 (60)1.3922340 (60)(e)
31Ga 1.3440260 (40)1.3401270 (96)1.208390 (75)1.207930 (34)(b), (f)
33As 1.108830 (31)1.104780 (12)0.992689 (79)0.992189 (53)(b), (f)
34Se  1.043836 (30)1.039756 (30)0.933284 (74)0.932804 (30)(b), (f)
36Kr 0.9843590 (44)0.9802670 (40)0.8790110 (70)0.8785220 (50)(b)
40Zr 0.7901790 (25)0.7859579 (27)0.7023554 (30)0.7018008 (30)(b)
42Mo 0.713607 (12)0.70931715 (41)0.632887 (13)0.632303 (13)(d), (f)
44Ru 0.6474205 (61)0.6430994 (61)0.5730816 (42)0.5724966 (42)(d), (f)
45Rh 0.6176458 (61)0.6132937 (61)0.5462139 (42)0.5456189 (42)(d), (f)
46Pd 0.5898351 (60)0.5854639 (46)0.5211363 (41)0.5205333 (41)(d), (f)
47Ag 0.5638131 (26)0.55942178 (76)0.4976977 (60)0.4970817 (60)(d), (f)
48Cd 0.5394358 (46)0.5350147 (46)0.4757401 (71)0.4751181 (71)(d), (f)
49In 0.5165572 (60)0.5121251 (46)0.4551966 (41)0.4545616 (41)(d), (f)
50Sn 0.4950646 (46)0.4906115 (46)0.4358821 (51)0.4352421 (51)(d), (f)
51Sb 0.4748391 (45)0.4703700 (45)0.4177477 (41)0.4170966 (31)(d), (f)
54Xe 0.42088103 (71)0.4163508 (14)0.3694051 (13)0.3687346 (13)(d)
56Ba 0.38968378 (74)0.38512464 (84)0.3415228 (11)0.34082708 (75)(d)
60Nd 0.3248079 (59)0.3201648 (59)0.283634 (59)0.282904 (44)(d), (f)
62Sm 0.31369830 (79)0.30904506 (46)0.273764 (30)0.273014 (30)(d), (f)
67Ho 0.26549088 (84)0.2607608 (42)0.230834 (30)0.230124 (30)(f), (g)
68Er 0.2571133 (11)0.25237359 (62)0.2234766 (14)0.22269866 (72)(d)
69Tm 0.24910095 (61)0.24434486 (44)0.216366 (30)0.21559182 (57)(f), (h)
74W 0.21383304 (50)0.20901314 (18)0.18518317 (70)0.1843768 (30)(d), (f)
79Au 0.18507664 (61)0.18019780 (47)0.1598249 (13)0.15899527 (77)(d)
82Pb 0.17029527 (56)0.16537816 (38)0.1468129 (10)0.14596836 (58)(d)
83Bi 0.1657183 (20)0.1607903 (46)0.142780 (11)0.1419492 (54)(f), (g)
90Th2300.13782600 (31)0.13282021 (36)0.11828686 (78)0.11740759 (59)(d)
91Pa2310.1343516 (29)0.1293302 (27)0.1152427 (21)0.1143583 (21)(i)
92U2380.13099111 (78)0.12595977 (36)0.11228858 (66)0.11140132 (65)(d)
93Np2370.1277287 (39)0.1226882 (36)0.1094230 (39)0.1085265 (28)(i)
94Pu2390.1245782 (15)0.11952120 (69)  (h)
94Pu2440.1245705 (25)0.1195140 (23)0.1066611 (18)0.1057595 (18)(i)
95Am2430.1215158 (24)0.1164463 (33)0.1039794 (17)0.1030803 (17)(i)
96Cm2480.1185427 (23)0.1134635 (21)0.1013753 (17)0.1004708 (16)(i)
97Bk2490.1156630 (54)0.1105745 (49)0.0988598 (55)0.0979514 (54)(i)
98Cf2500.1128799 (82)0.1077793 (75)  (i)

References: (a) Schweppe et al. (1994[link]); (b) Mooney (1996[link]); (c) Schweppe (1995[link]); (d) Deslattes & Kessler (1985[link]); (e) Hölzer et al. (1997[link]); (f) Bearden (1967[link]); (g) Borchert, Hansen, Jonson, Ravn & Desclaux (1980[link]); (h) Borchert (1976[link]); (i) Barreau, Börner, Egidy & Hoff (1982[link]).

The optically based data set was expanded by noting that several groups of accurate (relative) measurements in the literature either contained one of the directly measured lines (bold type) in Table 4.2.2.1[link] or were explicitly connected to one of them. Most often this situation was realized in reports that indicated a specific reference value, i.e. where it was stated numerical values are based on a scale where, for example, the wavelength of Mo [K\alpha _{1}] was taken as 707.831 xu. In such cases, and where other indicators of good measurement quality are presented, it is easy to re-scale the data reported so that it is consistent with the optically based data. This procedure was followed for important groups of measurements from earlier work by Bearden and co-workers, and from the X-ray laboratory at Uppsala. The rescaled numerical results are included in Table 4.2.2.1[link] in normal type along with the specific literature citations. The indicated uncertainties are standard uncertainties as defined by ISO (Taylor & Kuyatt, 1994[link]; Schwarzenbach, Abrahams, Flack, Prince & Wilson, 1995[link]).

4.2.2.6. L-series reference wavelengths

| top | pdf |

To date, only a very limited number of L-series emission lines have been directly measured on an optically based scale. Wavelengths for directly measured L-series lines are reported in Table 4.2.2.2[link] along with the literature citations. In the future, we hope to expand this limited data set by including other lines and elements that are well connected to these reference lines in the literature.

Table 4.2.2.2| top | pdf |
Directly measured L-series reference wavelengths in Å

Numbers in parentheses are standard uncertainties in the least significant figures.

ZSymbol[L\alpha _2][L\alpha _1][L\beta _1]References
36Kr7.82032(13)7.82032(13)7.574441(98)(a)
40Zr6.07710(48)6.070250(79)5.836214(76)(a)
54Xe3.025940(22)3.016582(15)2.806553(19)(b)
60Nd2.38079(52)2.370526(16)2.167008(19)(a)
62Sm2.210430(24)2.199873(13)1.998432(30)(a)
67Ho1.856472(15)1.845092(17)1.647484(32)(a)
68Er1.795701(45)1.784481(20)1.587466(86)(a)
69Tm1.738003(19)1.7267720(70)1.5302410(70)(a)

References: (a) Mooney (1996[link]); (b) Mooney et al. (1992[link]).

4.2.2.7. Absorption-edge locations

| top | pdf |

Only a small number of absorption-edge locations have been directly measured to high accuracy using the currently acceptable protocols. Some of the available data were obtained in order to provide wavelength determinations for spectra from highly charged and/or exotic atoms (Bearden, 1960[link]; Lum et al., 1981[link]). This small group was, however, significantly expanded very recently by an important set of new measurements, extending to Z = 51, that are well coupled to the optical wavelength scale (Kraft, Stümpel, Becker & Kuetgens, 1996[link]) The resulting experimental database is summarized in Table 4.2.2.3[link]. The effort that would be needed to expand the experimental database in a systematic way is quite large. Thus, we make use of a procedure, not previously used for this purpose, that combines available electron binding energy data with emission-line locations from our expanded reference set of emission-line data and emission lines that have been rescaled to be consistent with the optically based scale. At the same time, calculation of the location of absorption thresholds within the theoretical framework (see below) has been undertaken and will be made available in the longer publication and on the web site.

Table 4.2.2.3| top | pdf |
Directly measured and emission + binding energies (see text) K-absorption edges in Å

Numbers in parentheses are standard uncertainties in the least significant figures.

ZSymbolDirectly measuredEmission + binding energiesReferences
23V2.269211(21)2.26893(11)(a)
24Cr2.070193(14)2.07014(17)(a)
25Mn1.8964592(58)1.896457(42)(a)
26Fe1.7436170(49)1.743589(98)(a)
27Co1.6083510(42)1.60836(17)(a)
28Ni1.4881401(36)1.48823(25)(a)
29Cu1.3805971(31)1.38060(16)(a)
30Zn1.2833798(40)1.28338(15)(a)
39Y0.7277514(21)0.727750(23)(a)
40Zr0.6889591(31)0.688946(30)(a)
41Nb0.6531341(14)0.653112(29)(a)
42Mo0.61991006(62)0.619906(64)(a)
45Rh0.5339086(69)0.533951(10)(a)
46Pd0.5091212(42)0.509156(11)(a)
47Ag0.4859155(57)0.4859168(91)(a)
48Cd0.4641293(35)0.464135(12)(a)
49In0.4437454(48)0.443740(11)(a)
50Sn0.4245978(29)0.424590(13)(a)
51Sb0.4066324(27)0.406612(12)(a)
68Er0.2156801(75)0.2156762(50)(b)
82Pb0.1408821(74)0.1408836(11)(c)

References: (a) Kraft et al. (1996[link]); (b) Lum et al. (1981[link]); (c) Bearden (1960[link]).

The feature of absorption spectra customarily designated as `the absorption edge' has been variously associated with: the first inflection point of the absorption spectrum; the energy needed to produce a single inner vacancy with the photo-electron `at rest at infinity'; or the energy needed to remove an electron from an inner shell and place it in the lowest unoccupied energy level. A general discussion of this question has been given by Parratt (1959[link]). If we choose the second alternative, then it is easy to see that, with some care for symmetry restrictions, one can estimate the absorption-edge energy by combining the binding energy for any accessible outer shell with the energy of an emission line for which the transition terminus lies in the same outer shell. Of course, this procedure does not focus on the details of absorption thresholds, the locations of which are important for a number of structural applications. On the other hand, our choice gives greater regularity with respect to nuclear charge and facilitates use of electron binding energies, since they are referenced to the Fermi energy or the vacuum.

Electron binding energies have been tabulated for the principal electron shells of all the elements considered in the present table (Fuggle, Burr, Watson, Fabian & Lang, 1974[link]; Cardona & Ley, 1978[link]; Nyholm, Berndtsson & Mårtensson, 1980[link]; Nyholm & Mårtensson, 1980[link]; Lebugle, Axelsson, Nyholm & Mårtensson, 1981[link]; Powell, 1995[link]). The number of values available offers the possibility of consistency checking, since the K and L shells are connected by emission lines to several final hole states, each of which has (possibly) been evaluated by photoelectron spectroscopy. For each of the elements for which well qualified reference spectra are available, we evaluated edge location estimates using several alternative transition cycles and used the distribution of results to provide a measure of the uncertainty. Comparison of edge estimates obtained by this procedure with experimental data provides a quantitative test of the utility of the chosen approach to edge location estimation. In Table 4.2.2.3[link], the numerical results in the column labelled `Emission + binding energies' were obtained by combining emission energies and electron binding energies using all possible redundancies. The estimated uncertainties indicated were obtained from the distribution of the redundant routes. As can be seen, the results are in general agreement with the available directly measured values. Accordingly, we have used this protocol to obtain the edge locations listed in the summary tables below.

4.2.2.8. Outline of the theoretical procedures

| top | pdf |

Only recently has it become possible to understand the relativistic many-body problem in atoms with sufficient detail to permit meaningful calculation of transition energies between hole states (Indelicato & Lindroth, 1992[link]; Mooney, Lindroth, Indelicato, Kessler & Deslattes, 1992[link]; Lindroth & Indelicato, 1993[link], 1994[link]; Indelicato & Lindroth, 1996[link]). To deal with those hole states for atomic numbers ranging from 10 to 100, one needs to consider five kinds of contributions, all of which must be calculated in a relativistic framework, and the relative influence of which can change strongly as a function of the atomic number:

  • (i) nuclear size;

  • (ii) relativistic effects (corrections to Coulomb energy, magnetic and retardation energy);

  • (iii) Coulomb and Breit correlation;

  • (iv) radiative (QED) corrections (one- and two-electron Lamb shift etc.);

  • (v) Auger shift.

Such an undertaking, although much more advanced than any other done in the past, still suffers from severe limitations that need to be understood fully to make the best use of the table. The main limitation is probably that most lines are emitted by atoms in an elemental solid or a compound, while the calculation at present deals only with atoms isolated in vacuum. (A purely experimental database would have a similar limitation.) The second limitation is that it is not possible at present to include the coupling between the hole and open outer shells. Coupling between a [j={1\over2}], [j={3\over2}] or [j={5\over2}] hole and an external 3d or 4f shell can generate hundreds of levels, with splitting that can reach an eV. One then should calculate all radiative and Auger transition probabilities between hundreds of initial and final states. (The Auger final state would have one extra vacancy, leading eventually to thousands of final states.) Such an approach would give not only the mean line energy but also its shape and would thus be very desirable, but is impossible to do with present day theoretical tools and computers. We have thus limited ourselves to an approach in which one computes the weighted average energy for each hole state, and ignores possible distortion of the line profile due to the coupling between inner vacancies and outer shells.

Since we want to have good predictions for both light and heavy atoms, we have to include relativity non-perturbatively. To get a result approaching 1 ×  10−6 for uranium Kα by applying perturbation theory to the Schrödinger equation, for example, one would need to go to order 22 in powers of Zα = v/c. The natural framework in this case is thus to do a calculation exact to all orders in Zα by using the Dirac equation. We thus have used many-body methods, based on the Dirac equation, in which the main contributions to the transition energy are evaluated using the Dirac–Fock method. We use the Breit operator for the electron–electron interaction, to include magnetic (spin–spin, spin–other orbit and orbit–orbit interactions in the lower orders in Zα and (v/c)2 retardation effects. Higher-order retardation effects are also included. Many-body effects are calculated by using relativistic many-body perturbation theory (RMBPT). Since inner vacancy levels are auto-ionizing, one must include shifts in their energy due to the coupling between the discrete levels and Auger decay continuua.

In the following subsections, we describe in more detail the calculation of the different contributions.

4.2.2.9. Evaluation of the uncorrelated energy with the Dirac–Fock method

| top | pdf |

The first step in the calculation, following Indelicato and collaborators (Indelicato & Desclaux, 1990[link]; Indelicato & Lindroth, 1992[link]; Mooney et al., 1992[link]; Lindroth & Indelicato, 1993[link]; Indelicato & Lindroth, 1996[link]) consists in evaluating the best possible energy with relativistic corrections, within the independent electron approximation, for each hole state (here [1s_{{1\over2}}], [2p_{{1\over2}}], [2p_{{3\over2}}], [3p_{{1\over2}}], [3p_{{3\over2}}] for K, LII, LIII, MII, MIII, respectively). Such a calculation must provide a suitable starting point for adding all many-body and QED contributions. We have thus chosen the Dirac–Fock method in the implementation of Desclaux (1975[link], 1993[link]). This method, based on the Dirac equation, allows treatment of arbitrary atoms with arbitrary structure and has been widely used for this kind of calculation. We have used it with full exchange and relaxation (to account for inactive orbital rearrangement due to the hole presence). The electron–electron interaction used in this program contains all magnetic and retardation effects, which is very important to have good results at large Z. The magnetic interaction is treated on an equal footing with the Coulomb interaction, to account for higher-order effects in the wavefunction (which are also useful for evaluating radiative corrections to the electron–electron interaction). All these calculations must be done with proper nuclear charge models to account for finite-nuclear-size corrections to all contributions. For heavy nuclei, nuclear deformations must be accounted for (Blundell, Johnson & Sapirstein, 1990[link]; Indelicato, 1990[link]). For all elements for which experiments have been performed, we used experimental nuclear charge radii. For the others we used a formula from Johnson & Soff (1985[link]), corrected for nuclear deformations for Z [\gt] 90. Contribution of deformation to the r.m.s. radius (the only parameter of importance to the atomic calculation) is roughly constant (0.11 fm) for Z [\gt] 90. There is an unknown region, between Bi and Th (83 [\lt] Z [\lt] 90), where deformation effects start to be important, but for which they are not known. When experiments are done for a particular isotope, we calculated separately the energies for each isotope.

As mentioned in the introduction, there are special difficulties involved when dealing with atoms with open outer shells (obviously this is the most common case). Computing all energies [E_{{J}}] for total angular momentum J would be both impossible and useless. The Dirac–Fock method circumvents this difficulty. One can evaluate directly an average energy that corresponds to the barycentre of all [E_{{J}}] with weight ([2{J}+1]). There are still a few cases for which the average calculation cannot converge (when the open shells have identical symmetry). In that case, the outer electrons have been rearranged in an identical fashion for all hole states of the atom, to minimize possible shifts due to this procedure.

4.2.2.10. Correlation and Auger shifts

| top | pdf |

Once the Dirac–Fock energy is obtained, many-body effects beyond Dirac–Fock relaxation must be taken into account. These include relaxation beyond the spherical average, correlation (due to both Coulomb and magnetic interaction), and corrections due to the autoionizing nature of hole states (Auger shift). Since the many-body generalization of the Dirac–Fock method, the so-called MCDF (multiconfiguration Dirac–Fock), is very inefficient for hole states, we turned to RMBPT to evaluate those quantities. These many-body effects contribute very significantly to the final value. Coulomb correlation is mostly constant along the Periodic Table (at the level of a few eV). Magnetic correlations are very strong at high Z. Auger shift is very important for p states. The interested reader will find more details of these complicated calculations in the original references (Indelicato & Lindroth, 1992[link]; Mooney et al., 1992[link]; Lindroth & Indelicato, 1993[link]; Indelicato & Lindroth, 1996[link]). As these calculations are very time consuming, they are performed only for selected Z and interpolated. Since the Auger shifts do not always have a smooth Z dependence, care has been taken to evaluate them at as many different Z's as practical to ensure a good reproduction of irregularities.

4.2.2.11. QED corrections

| top | pdf |

The QED corrections originate in the quantum nature of both the electromagnetic and electron fields. They can be divided in two categories, radiative and non-radiative. The first one includes self-energy and vacuum polarization, which are the main contributions to the Lamb shift in one-electron atoms. These corrections scale as [{Z}^{4}/{n}^{3}] (n being the principal quantum number) and are thus very important for inner shells and high Z. The second category is composed of corrections to the electron–electron interaction that cannot be accounted for by RMBPT or MCDF. These corrections start at the two-photon interaction and include three-body effects. The two-photon, non-radiative QED contribution has been calculated recently only for the ground state of two-electron ions (Blundell, Mohr, Johnson & Sapirstein, 1993[link]; Lindgren, Persson, Salomonson & Labzowsky, 1995[link]) and cannot be evaluated in practice for atoms with more than two or three electrons.

The radiative corrections split up into two contributions. The first contribution is composed of one-electron radiative corrections (self-energy and vacuum polarization). For the self-energy and [{Z}\gt10], one must use all-order calculations (Mohr, 1974a,[link]b[link], 1975[link], 1982[link], 1992[link]; Mohr & Soff, 1993[link]). Vacuum polarization can be evaluated at the Uehling (1935[link]) and Wichmann & Kroll (1956[link]) level. Higher-order effects are much smaller than for the self-energy (Soff & Mohr, 1988[link]) and have been neglected. The second contribution is composed of radiative corrections to the electron–electron interaction, and scales as [{Z}^{3}/{n}^{3}]. Ab initio calculations have been performed only for few-electron ions (Indelicato & Mohr, 1990[link], 1991[link]). Here we use the Welton approximation which has been shown to reproduce very closely ab initio results in all examples that have been calculated (Indelicato, Gorceix & Desclaux 1987[link]; Indelicato & Desclaux 1990[link]; Kim, Baik, Indelicato & Desclaux, 1991[link]; Blundell, 1993a[link],b[link]).

4.2.2.12. Structure and format of the summary tables

| top | pdf |

Table 4.2.2.4[link] summarizes the theoretical and experimental results for prominent K-series lines and the K-absorption edge. For the emission lines, the upper number (in italics) is the theoretical estimate for this line and the lower number is the experimentally measured value (1) from Table 4.2.2.1[link] or (2) from the Bearden database or a reference that appeared after the Bearden database corrected to the optically based scale. For the K absorption edge, the upper number (also in italics) was obtained by combining emission lines and photoelectron spectroscopy (see Subsection 4.2.2.7[link]), and the lower number is the experimentally measured value (1) from Table 4.2.2.3[link] or (2) from the Bearden database or a reference that appeared after the Bearden database corrected to an optically based scale. For the experimental emission and absorption entries, bold type is used for wavelengths directly measured on an optically based scale. The numerical values for wavelengths in angstrom units (1 Å = 0.1 nm) are given to a number of significant figures commensurate with their estimated uncertainties, which appear in parentheses after each theoretical and experimental value.

Table 4.2.2.4| top | pdf |
Wavelengths of K-emission lines and K-absorption edges in Å; see text for explanation of typefaces

Numbers in parentheses are standard uncertainties in the least significant figures.

ZSymbolA[K\alpha _2 ][K\alpha _1 ][K\beta _3 ][K\beta _1 ][K\beta _2 ^{\rm II} ][K\beta _2 ^{\rm I}]K abs. edge
10Ne 14.6020(93)14.6006(93)    14.2391(26)
 14.6102(44)14.6102(44)14.4522(74)14.4522(74)  14.30201(15)
11Na 11.9013(59)11.8994(59)    11.4784(16)
 11.9103(13)11.9103(13)11.5752(30)11.5752(30)  11.5692(15)
12 Mg 9.8860(39)9.8840(39)    9.4479(10)
 9.89153(10)9.889554(88)9.5211(30)9.5211(30)  9.51234(15)
13Al 8.3372(27)8.3349(26)7.9412(49)   7.89928(67)
 8.341831(58)8.339514(58)7.9601(30)7.9601(30)  7.948249(74)
14Si 7.1269(19)7.1208(19)6.7317(26)   6.70091(46)
 7.12801(14)7.125588(78)6.7531(15)6.7531(15)  6.7381(15)
15P 6.1587(14)6.1539(14)5.7834(16)5.7914(27)  5.75537(33)
 6.1601(15)6.1571(15)5.7961(30)5.7961(30)  5.7841(15)
16S 5.3742(10)5.3701(10)5.0202(12)5.0246(15)  4.99591(24)
 5.374960(89)5.372200(78) 5.03168(30)  5.01858(15)
17Cl 4.72993(80)4.72560(77)4.39810(99)4.40038(77)  4.37679(18)
 4.730693(71)4.727818(71)4.40347(44)4.40347(44)  4.39717(15)
18 Ar 4.19448(62)4.19162(60)3.88506(71)3.88486(70)  3.86552(14)
 4.194939(23)4.191938(23)3.88606(30)3.88606(30)  3.870958(74)
19K 3.74352(50)3.74055(48)3.45189(69)3.45216(58)  3.42856(11)
 3.7443932(68)3.7412838(56)3.45395(30)3.45395(30)  3.43655(15)
20Ca 3.36223(39)3.35911(38)3.08855(45)3.08827(45)  3.061828(87)
 3.361710(44)3.358440(44)3.08975(30)3.08975(30)  3.07035(15)
21Sc 3.03479(33)3.03129(31)2.77919(50)2.77809(49)  2.754176(71)
 3.0344010(63)3.030854(14)2.77964(30)2.77964(30)  2.7620(15)
22Ti 2.75272(27)2.74886(26)2.51445(43)2.51262(45)  2.490681(59)
 2.7521950(57)2.7485471(57)2.513960(30)2.513960(30)  2.497377(74)
23V 2.50798(23)2.50383(21)2.28567(37)2.28332(40)  2.263194(49)
 2.507430(30)2.503610(30)2.284446(30)2.284446(30)  2.269211(21)
24Cr 2.29428(19)2.29012(18)2.08702(32)2.08478(35)  2.067898(41)
 2.2936510(30)2.2897260(30)2.0848810(40)2.0848810(40)  2.070193(14)
25Mn 2.10635(16)2.10210(15)1.91175(28)1.90960(31)  1.892275(36)
 2.1058220(30)2.1018540(30)1.9102160(40)1.9102160(40)  1.8964592(58)
26Fe 1.94043(14)1.93631(13)1.75784(25)1.75617(27)  1.739918(31)
 1.9399730(30)1.9360410(30)1.7566040(40)1.7566040(40)  1.7436170(49)
27Co 1.79321(12)1.78919(11)1.62166(22)1.62039(24)  1.605127(27)
 1.7928350(10)1.7889960(10)1.6208260(30)1.6208260(30)  1.6083510(42)
28Ni 1.66199(10)1.658049(96)1.50059(19)1.49964(21)  1.485300(24)
 1.6617560(10)1.6579300(10)1.5001520(30)1.5001520(30)  1.4881401(36)
29Cu 1.544324(93)1.540538(85)1.39246(17)1.39201(18)  1.379448(23)
 1.54442740(50)1.54059290(50)1.3922340(60)1.3922340(60)  1.3805971(31)
30Zn 1.438963(84)1.435151(74)1.29544(17)1.29506(16)  1.282346(20)
 1.439029(12)1.435184(12)1.295276(30)1.295276(30)1.283739(30)1.283739(30)1.2833798(40)
31Ga 1.343987(72)1.340095(65)1.20821(13)1.20774(14)1.195547(25) 1.194711(18)
 1.3440260(40)1.3401270(96)1.208390(75)1.207930(34)1.196018(30)1.196018(30)1.19582(15)
32Ge 1.257998(65)1.254054(58)1.12924(13)1.12877(13)1.116387(37) 1.115585(16)
 1.258030(13)1.254073(13)1.12938(13)1.128957(30)1.116877(30)1.116877(30)1.116597(74)
33As 1.179921(57)1.175932(52)1.05774(11)1.05724(11)1.044699(56)1.044836(20)1.043925(16)
 1.179959(17)1.17595600(90)1.057898(76)1.057368(33)1.045016(44)1.045016(44)1.04502(15)
34Se 1.108801(52)1.104778(47)0.992646(96)0.992152(95)0.979618(57)0.979716(26)0.978818(15)
 1.108830(31)1.104780(12)0.992689(79)0.992189(53)0.979935(74)0.979935(74)0.979755(15)
35Br 1.043841(47)1.039785(42)0.933275(87)0.932768(84)0.920344(49)0.920390(28)0.919501(13)
 1.043836(30)1.039756(30)0.933284(74)0.932804(30)0.920474(30)0.920474(30)0.92041(15)
36Kr 0.984347(42)0.980267(38)0.878967(81)0.878495(75)0.866209(36)0.866169(35)0.865324(12)
 0.9843590(44)0.9802670(40)0.8790110(70)0.8785220(50)0.86611(15)0.86611(15)0.865533(15)
37Rb 0.929713(39)0.925597(35)0.829174(71)0.828681(67)0.816459(33)0.816408(33)0.815270(12)
 0.929704(15)0.925567(13)0.829222(44)0.828692(30)0.816462(44)0.816462(44)0.815552(74)
38Sr 0.879444(36)0.875298(32)0.783413(63)0.782911(58)0.770774(33)0.770718(20)0.769359(11)
 0.879443(15)0.875273(15)0.783462(44)0.782932(30)0.770822(44)0.770822(44)0.769742(74)
39Y 0.833059(32)0.828875(29)0.741232(58)0.740716(53)0.728801(27)0.728663(21)0.727270(10)
 0.833063(15)0.828852(15)0.741271(44)0.740731(30)0.728651(59)0.728651(59)0.7277514(21)
40Zr 0.790181(30)0.785960(27)0.702296(53)0.701766(48)0.690079(28)0.689895(21)0.6884893(99)
 0.7901790(25)0.7859579(27)0.7023554(30)0.7018008(30)0.689940(59)0.689940(59)0.6889591(31)
41Nb 0.750448(28)0.746189(25)0.666266(49)0.665721(44)0.654328(31)0.654078(22)0.652890(93)
 0.750451(15)0.746211(15)0.666350(44)0.665770(30)0.654170(59)0.654170(59)0.6531341(14)
42Mo 0.713612(25)0.709328(22)0.632900(44)0.632345(38)0.621162(35)0.620941(21)0.6196481(87)
 0.713607(12)0.70931715(41)0.632887(13)0.632303(13)0.620999(30)0.620999(30)0.61991006(62)
43Tc 0.679318(24)0.675017(21)0.601881(40)0.601318(35)0.590423(40)0.590231(22)0.5889852(84)
 0.679330(44)0.675030(44)0.601889(59)0.601309(59)0.590249(74)0.590249(74)0.589069(15)
44Ru 0.647415(22)0.643088(20)0.573053(37)0.572478(32)0.561748(44)0.561587(22)0.5603122(81)
 0.6474205(61)0.6430994(61)0.5730816(42)0.5724966(42)0.561668(44)0.561668(44)0.560518(15)
45Rh 0.617652(21)0.613305(18)0.546191(34)0.545606(29)0.535110(48)0.534977(22)0.5337192(74)
 0.6176458(61)0.6132937(61)0.5462139(42)0.5456189(42)0.535038(30)0.535038(30)0.5339086(69)
46Pd 0.589822(20)0.585459(18)0.521117(29)0.520514(27)0.510283(46)0.510177(51)0.5090158(75)
 0.5898351(60)0.5854639(46)0.5211363(41)0.5205333(41)0.5102357(59)0.5102357(59)0.5091212(42)
47Ag 0.563804(18)0.559420(17)0.497673(29)0.497069(25)0.487060(55)0.487019(38)0.4857609(74)
 0.5638131(26)0.55942178(76)0.4976977(60)0.4970817(60)0.4870393(59)0.4870393(59)0.4859155(57)
48Cd 0.539426(18)0.535020(15)0.475739(27)0.475124(23)0.465335(62)0.465346(28)0.4640026(71)
 0.5394358(46)0.5350147(46)0.4757401(71)0.4751181(71)0.465335(10)0.465335(10)0.4641293(35)
49In 0.516551(17)0.512124(15)0.455178(25)0.454552(22)0.445014(58)0.445011(27)0.4435977(70)
 0.5165572(60)0.5121251(46)0.4551966(41)0.4545616(41)0.445007(15)0.445007(15)0.4437454(48)
50Sn 0.495060(16)0.490612(14)0.435878(24)0.435241(20)0.426120(12)0.425928(26)0.4244611(68)
 0.4950646(46)0.4906115(46)0.4358821(51)0.4352421(51)0.425921(12)0.425921(12)0.4245978(29)
51Sb 0.474840(15)0.470373(13)0.417736(22)0.417089(19)0.408017(57)0.408004(25)0.4064886(65)
 0.4748391(45)0.4703700(45)0.4177477(41)0.4170966(31)0.4079791(74)0.4079791(74)0.4066324(27)
52Te 0.455795(14)0.451310(13)0.400664(21)0.400008(18)0.391161(56)0.391135(27)0.3895899(64)
 0.4557908(44)0.4513018(44)0.4006650(59)0.4000010(44)0.3911079(89)0.3911079(89)0.389746(15)
53I 0.437834(13)0.433330(12)0.384576(20)0.383910(17)0.375286(54)0.375234(29)0.3736775(61)
 0.437836(10)0.4333245(74)0.3845698(59)0.3839108(59)0.375236(30)0.375236(30)0.373816(15)
54Xe 0.420879(42)0.416358(40)0.369407(40)0.368730(38)0.360326(72)0.36034(12)0.358683(27)
 0.42088103(71)0.4163508(14)0.3694051(13)0.3687346(13)0.360265(44)0.360265(44)0.35841(74)
55Cs 0.404848(13)0.400310(11)0.355067(17)0.354385(16)0.346197(49)0.346102(37)0.3444778(59)
 0.4048411(59)0.4002960(59)0.3550553(59)0.354369(10)0.346115(30)0.346115(30)0.344515(15)
56Ba 0.389684(12)0.385129(11)0.341517(16)0.340826(15)0.33285(14)0.332728(12)0.3310639(56)
 0.38968378(74)0.38512464(84)0.3415228(11)0.34082708(75)0.332775(15)0.332775(15)0.331045(15)
57La 0.375320(11)0.370748(10)0.328692(16)0.327993(14)0.32023(13)0.320101(11)0.3184025(55)
 0.3753186(30)0.3707426(30)0.3286909(59)0.3279879(44)0.320122(10)0.320122(10)0.318445(74)
58Ce 0.361685(11)0.3570964(97)0.316507(15)0.315795(14)0.30827(12)0.308131(10)0.3065382(54)
 0.3616884(30)0.3570974(30)0.3165248(59)0.3158207(30)0.308165(15)0.308165(15)0.306485(74)
59Pr 0.348755(10)0.3441494(94)0.304970(14)0.304249(13)0.29694(11)0.2967952(99)0.2952418(53)
 0.3487542(30)0.3441452(30)0.3049796(74)0.3042656(59)0.296794(30)0.296794(30)0.295184(74)
60Nd 0.336473(10)0.3318514(91)0.294021(13)0.293290(13)0.28619(10)0.2860408(94)0.2845288(52)
 0.33647921(73)0.33185689(62)0.2940366(40)0.2933086(40)0.28610(15)0.28610(15)0.284534(74)
61Pm 0.3247982(98)0.3201607(88)0.283620(13)0.282880(12)0.275992(98)0.2758335(91)0.2743634(53)
 0.3248079(59)0.3201648(59)0.283634(59)0.282904(44)0.27590(15)0.27590(15)0.274314(74)
62Sm 0.3136913(94)0.3090384(84)0.273732(12)0.272984(12)0.266459(91)0.2661277(87)0.2647027(51)
 0.31369830(79)0.30904506(46)0.273764(30)0.273014(30)0.26620(15)0.26620(15)0.264644(74)
63Eu 0.3031139(91)0.2984457(81)0.264322(12)0.263567(11)0.257069(85)0.2569028(81)0.2555123(51)
 0.3031225(30)0.2984505(30)0.2643360(74)0.2635810(74)0.256927(12)0.256927(12)0.255534(15)
64Gd 0.2930400(89)0.2883568(79)0.255371(11)0.254610(11)0.248289(23)0.2481186(76)0.2467265(48)
 0.2930424(30)0.2883573(30)0.255344(30)0.254604(30)0.248164(44)0.248164(44)0.246814(15)
65Tb 0.2834212(86)0.2787234(76)0.246818(11)0.246054(11)0.239916(21)0.2397496(75)0.2384335(49)
 0.2834273(30)0.2787242(30)0.246834(30)0.246084(30)0.23970(30)0.23970(30)0.238414(15)
66Dy 0.2742462(84)0.2695341(74)0.238671(10)0.237902(10)0.231955(12)0.2318190(53)0.2304867(46)
 0.2742511(30)0.2695370(30)0.238624(30)0.237884(30)0.23170(30)0.23170(30)0.230483(15)
67Ho 0.2654851(81)0.2607589(72)0.230896(10)0.230122(10)0.224320(18)0.2241536(66)0.2229099(45)
 0.26549088(84)0.2607608(42)0.230834(30)0.230124(30)0.22410(30)0.22410(30)0.222913(15)
68Er 0.2571059(79)0.2523659(71)0.2234662(97)0.2226875(98)0.217046(16)0.2168806(64)0.2156719(45)
 0.2571133(11)0.25237359(62)0.2234766(14)0.22269866(72)0.21670(30)0.21670(30)0.2156801(75)
69Tm 0.2490952(77)0.2443415(68)0.2163665(94)0.2155833(95)0.210099(15)0.2099331(62)0.2087587(44)
 0.24910095(61)0.24434486(44)0.216366(30)0.21559182(57)0.20980(30)0.20980(30)0.208803(74)
70Yb 0.2414274(75)0.2366603(67)0.2095741(93)0.2087863(95)0.203456(14)0.2032912(59)0.2021481(43)
 0.2414276(30)0.2366586(30)0.20960(15)0.208843(30)0.20330(30)0.20330(30)0.202243(74)
71Lu 0.2340857(73)0.2293053(65)0.2030802(88)0.2022872(90)0.191017(13)0.1969329(58)0.1957973(42)
 0.2340845(30)0.2293014(30)0.203093(59)0.202313(44)0.19690(30)0.19690(30)0.195853(74)
72Hf 0.2270507(72)0.2222572(64)0.1968603(86)0.1960622(88)0.191017(13)0.1908468(56)0.1897176(42)
 0.2270274(44)0.2222303(44)0.196863(59)0.196073(44)0.19080(30)0.19080(30)0.189823(74)
73Ta 0.2203039(70)0.2154977(63)0.1908986(83)0.1900954(86)0.185143(12)0.1849702(54)0.1838657(41)
 0.2203083(59)0.2155002(59)0.1908929(30)0.1900919(59)0.185191(13)0.185014(12)0.183943(15)
74W 0.2138327(69)0.2090134(61)0.1851834(81)0.1843751(83)0.179595(12)0.1794215(52)0.1783098(41)
 0.21383304(50)0.20901314(18)0.18518317(70)0.1843768(30)0.179603(15)0.179424(10)0.178373(15)
75Re 0.2076150(67)0.2027835(60)0.1796955(79)0.1788824(81)0.174234(11)0.1740571(51)0.1729509(40)
 0.2076141(15)0.2027840(30)0.1796997(44)0.1788827(44)0.174253(15)0.1740566(89)0.173023(15)
76Os 0.2016443(66)0.1968007(59)0.1744279(77)0.1736101(79)0.169085(11)0.1689066(50)0.1678092(40)
 0.2016420(30)0.1967970(30)0.1744336(44)0.1736136(44)0.169103(15)0.1689085(89)0.167873(15)
77Ir 0.1959045(65)0.1910492(57)0.1693667(75)0.1685444(77)0.164150(11)0.1639697(51)0.1628853(39)
 0.1959069(30)0.1910499(30)0.1693695(30)0.1685445(30)0.164152(15)0.163958(10)0.162922(15)
78Pt 0.1903859(61)0.1855187(55)0.1645026(72)0.1636756(74)0.1593872(99)0.1592048(46)0.1581346(38)
 0.1903839(59)0.1855138(59)0.1645035(44)0.1636775(44)0.159392(15)0.159202(15)0.158182(15)
79Au 0.1850702(64)0.1801914(57)0.1598202(73)0.1589887(75)0.1548206(99)0.1546363(48)0.1535699(40)
 0.18507664(61)0.18019780(47)0.1598249(13)0.15899527(77)0.154832(30)0.154620(13)0.1535953(74)
80Hg 0.1799628(61)0.1750720(54)0.1553217(69)0.1544857(72)0.1504204(94)0.1502334(46)0.1491786(38)
 0.1799607(44)0.1750706(44)0.1553233(44)0.1544893(44)0.150402(30)0.150202(30)0.149182(15)
81Tl 0.1750380(60)0.1701355(53)0.1509866(68)0.1501462(70)0.1461874(92)0.1459989(77)0.1449460(37)
 0.1750386(30)0.1701386(30)0.1509823(89)0.1501443(74)0.146142(15)0.145952(15)0.144952(15)
82Pb 0.1702924(59)0.1653781(53)0.1468107(67)0.1459663(68)0.1421118(88)0.1419201(75)0.1408707(37)
 0.17029527(56)0.16537816(38)0.1468129(10)0.14596836(58)0.142122(30)0.141912(15)0.1408821(74)
83Bi2090.1657170(58)0.1607911(52)0.1427865(65)0.1419372(66)0.1381841(87)0.1379910(72)0.1369439(37)
 0.1657183(20)0.1607903(46)0.142780(11)0.1419492(54)0.138172(15)0.137972(15)0.136942(15)
84Po2090.1613031(58)0.1563656(51)0.1389056(63)0.1380520(65)0.1343966(85)0.1342012(69)0.1331589(36)
 0.161302(15)0.156362(15)0.138922(30)0.138072(30)0.134382(30)0.134182(30) 
85At2100.1570444(56)0.1520953(50)0.1351623(62)0.1343044(63)0.1307448(83)0.1305470(67)0.1295098(36)
 0.157052(30)0.152102(30)0.135172(59)0.134322(59)0.130722(59)0.130522(59) 
86Rn2220.1529334(56)0.1479727(49)0.1315499(61)0.1306882(61)0.1272218(79)0.1270211(66)0.1259898(35)
 0.152942(44)0.147982(44)0.131552(74)0.130692(74)0.127192(74)0.126982(74) 
87Fr2230.1489599(56)0.1439878(50)0.1280599(60)0.1271937(61)0.1238183(79)0.1236157(63)0.1225852(36)
 0.148962(44)0.143992(44)0.128072(74)0.127192(74)0.123792(74)0.123582(74) 
88Ra2260.1451209(54)0.1401373(48)0.1246890(58)0.1238185(59)0.1205312(77)0.1203271(60)0.1192985(35)
 0.145119(20)0.140132(19)0.124689(15)0.123815(15)0.120535(14)0.120320(14) 
89Ac2270.1414083(54)0.1364131(47)0.1214301(57)0.1205554(58)0.1173552(73)0.1171477(59)0.1161246(34)
 0.141412(30)0.136419(12)0.121432(30)0.120552(30)0.117322(30)0.117112(30) 
90Th2320.1378266(53)0.1328194(47)0.1182861(56)0.1174071(56)0.1142910(71)0.1140810(57)0.1130642(34)
 0.13782600(31)0.13282021(36)0.11828686(78)0.11740759(59)0.114262(15)0.114042(13)0.113072(15)
91Pa2310.1343514(52)0.1293324(46)0.1152364(55)0.1143530(55)0.1113088(69)0.1110964(56)0.1101087(34)
 0.1343516(29)0.1293302(27)0.1152427(21)0.1143583(21)0.111292(30)0.111072(30) 
92U2380.1309879(52)0.1259572(46)0.1122860(53)0.1113979(54)0.1084449(67)0.1082301(54)0.1072452(33)
 0.13099111(78)0.12595977(36)0.11228858(66)0.11140132(65)0.108372(15)0.108182(15)0.107232(15)
93Np2370.1277298(51)0.1226871(45)0.1094299(52)0.1085378(53)0.1056621(66)0.1054450(53)0.1044744(33)
 0.1277287(39)0.1226882(36)0.1094230(39)0.1085265(28)0.105670(31)0.105457(31)0.1044605(62)
94Pu2440.1245763(50)0.1195212(45)0.1066627(51)0.1057661(52)0.1029688(64)0.1027494(52)0.1017982(33)
 0.1245705(25)0.1195140(23)0.1066611(18)0.1057595(18)0.1029724(26)0.1027429(26) 
95Am2430.1215172(50)0.1164501(45)0.1039811(51)0.1030805(51)0.1003579(63)0.1001364(51)0.0991999(33)
 0.1215158(24)0.1164463(33)0.1039794(17)0.1030803(17)0.1003537(24)0.1001357(24) 
96Cm2480.1185536(49)0.1134742(44)0.1013837(50)0.1004790(50)0.0978295(63)0.0976059(50)0.0966801(33)
 0.1185427(23)0.1134635(21)0.1013753(17)0.1004708(16)0.0978355(23)0.0975952(15) 
97Bk2490.1156777(49)0.1105860(43)0.0988636(48)0.0979546(49)0.0953724(61)0.0951469(49)0.0942405(32)
 0.1156630(54)0.1105745(49)0.0988598(55)0.0979514(54) 0.0942501(50) 
98Cf2500.1128873(48)0.1077832(43)0.0964130(47)0.0955000(48)0.0929867(61)0.0927593(48)0.0918695(32)
 0.1128799(82)0.1077793(75)0.0963915(83)0.0954860(90)0.0929715(82)0.0927508(84)0.091862(10)
99Es2510.1101788(47)0.1050620(43)0.0940403(46)0.0931231(47)0.0906838(60)0.0904543(47)0.0895840(32)
 0.1102072(98)0.1050554(89)0.094036(14)0.093090(14)  0.0895878(97)
100Fm2540.1075497(47)0.1024201(42)0.0917379(45)0.0908165(46)0.0884443(59)0.0882127(45)0.0873575(32)
 0.107514(14)0.102386(13)0.091715(10)0.0907943(98)0.0884212(100)0.0881872(99)0.0873356(80)

Figure 4.2.2.1[link] shows plots of the relative deviation between theoretical and experimental values for the K-series lines and the K-absorption edge as a function of Z. The error bars shown in the figure are the experimental uncertainties. In general, these plots show good agreement between theory and experiment except in the low-Z and high-Z regions. At the low-Z end of the table, the particular calculational approach used is not optimum, and the experimental data are surprisingly weak. At the high end, experimental data have rather large uncertainties, and thus do not provide an accurate test of the theory.

[Figure 4.2.2.1]

Figure 4.2.2.1| top | pdf |

Relative deviations between theoretical and experimental results for K-series spectra. The topmost data set concerns the K-edge location, while the other data sets, beginning at the bottom, refer to the Kα2, Kα1, Kβ3 and Kβ1, respectively. The ordinate scales have been displaced for clarity by the indicated multiples of 0.001.

Table 4.2.2.5[link] summarizes the theoretical and experimental results for prominent L-series lines and the L-absorption edges. The experimental database of high-accuracy emission data is much more limited than was the case for the K series, and there have been very few high-accuracy edge-location measurements. The format of this table is similar to that of Table 4.2.2.4[link]. For the emission lines, the upper number (in italics) is the theoretical estimate for this line, and the lower number is the experimentally measured value. Numbers in bold type were directly measured on the optical scale (see Table 4.2.2.2[link]), and numbers in normal type are from the Bearden database or a reference that appeared after the Bearden database corrected to an optically based scale. For the L-absorption edges, the upper number (also in italics) is obtained by combining emission lines and photoelectron spectroscopy (see Subsection 4.2.2.7[link]) and the lower number is the experimentally measured value. The numbers in bold type are recent measurements by Kraft, Stümpel, Becker & Kuetgens (1996[link]), and the numbers in normal type are from the Bearden database or a reference that appeared after the Bearden database corrected to an optically based scale. Fig. 4.2.2.2[link] shows relative deviations between the theoretical and experimental values for most of the tabulated data. The error bars shown in the figure are the experimental uncertainties.

Table 4.2.2.5| top | pdf |
Wavelengths of L-emission lines and L-absorption edges in Å; see text for explanation of typefaces

Numbers in parentheses are standard uncertainties in the least significant figures.

ZSymbolALα2Lα1Lβ1Lβ2LI abs. edgeLII abs. edgeLIII abs. edge
20Ca     28.275(32)35.384(40)35.7704(68)
 36.331(30)36.331(30)35.941(30)  35.131(15)35.491(15)
21Sc 30.947(46) 30.587(47) 24.896(15)30.718(17)31.109(36)
 31.350(44)31.350(44)31.020(30)    
22Ti 27.215(37) 26.843(37) 22.099(24)26.953(14)27.3105(36)
 27.420(30)27.420(30)27.050(30)  27.290(15)27.290(15)
23V 24.143(30) 23.764(30) 19.779(19)23.8561(89)24.206(10)
 24.250(44)24.250(44)23.880(59)    
24Cr 21.640(24)21.490(11)21.276(24) 17.804(15)21.246(18)21.5867(49)
 21.640(44)21.640(44)21.270(15) 16.70(15)17.90(15)20.70(15)
25Mn 19.390(20)19.359(21)19.036(20) 16.113(19)19.0781(57)19.4063(43)
 19.450(15)19.450(15)19.110(30)    
26Fe 17.525(17)17.503(17)17.194(17) 14.611(34)17.2248(92)17.5402(35)
 17.590(30)17.590(30)17.260(15)  17.2023(74)17.5253(74)
27Co 15.922(14)15.905(15)15.610(14) 13.4000(86)15.627(14)15.9290(44)
 15.9722(89)15.9722(89)15.666(12)  15.6182(74)15.9152(74)
28Ni 14.532(12)14.520(12)14.236(12) 12.295(13)14.251(23)14.5396(57)
 14.5612(44)14.5612(44)14.2712(89)  14.2422(74)14.5252(74)
29Cu 13.341(10)13.336(11)13.063(10) 11.292(16)13.016(14)13.2934(64)
 13.3362(44)13.3362(44)13.0532(44)  13.0142(15)13.2882(15)
30Zn 12.2529(90)12.2489(90)11.9819(93) 10.361(12)11.8652(66)12.134(14)
 12.2542(44)12.2542(44)11.9832(44) 13.060(15)11.8622(15)12.1312(15)
31Ga 11.2916(77)11.2858(78)11.0226(78) 9.518(11)10.8414(29)11.1040(29)
 11.2922(15)11.2922(15)11.0232(30) 9.5171(74)10.8282(74)11.1002(15)
32Ge 10.4371(68)10.4306(68)10.1717(69) 8.775(12)9.9340(27)10.1849(46)
 10.4363(12)10.4363(12)10.1752(15) 8.7731(15)9.9241(15)10.1872(15)
33As 9.6744(60)9.6680(60)9.4126(59) 8.092(13)9.1182(17)9.3649(29)
 9.6710(12)9.6710(12)9.4142(12) 8.1071(15)9.1251(15)9.3671(15)
34Se 8.9914(52)8.9852(52)8.7335(52) 7.498(13)8.4105(58)8.64459(77)
 8.99013(74)8.99013(74)8.73593(74) 7.5031(15)8.4071(15)8.6461(15)
35Br 8.3776(46)8.3715(46)8.1233(46) 6.958(14)7.7669(35)7.9991(30)
 8.37473(74)8.37473(74)8.12522(74) 6.9591(74)7.7531(74)7.9841(74)
36Kr 7.8242(41)7.8180(41)7.5736(40) 6.4561(41)7.1630(21)7.3841(17)
 7.82032(13)7.82032(13)7.574441(98) 6.470(15)7.1681(15)7.3921(15)
37Rb 7.3226(37)7.3164(36)7.0749(36) 6.0010(11)6.6449(59)6.8643(67)
 7.32521(44)7.31841(30)7.07601(44) 6.0081(74)6.6441(15)6.8621(15)
38Sr 6.8674(33)6.8610(32)6.6224(33) 5.5945(16)6.17624(70)6.38937(84)
 6.86980(44)6.86290(30)6.62400(44) 5.5921(74)6.1731(15)6.3871(15)
39Y 6.4539(30)6.4466(29)6.2110(29) 5.22968(53)5.75742(82)5.9658(15)
 6.45590(44)6.44890(30)6.21209(44) 5.2171(74)5.7561(15)5.9621(15)
40Zr 6.0766(27)6.0684(26)5.8357(26) 4.89881(41)5.3773(15)5.5816(15)
 6.0766(27)6.070250(79)5.836214(76)5.58638(44)4.8791(74)5.3781(15)5.5791(15)
41Nb 5.7326(24)5.7226(23)5.4931(23) 4.59975(43)5.03480(63)5.23529(98)
 5.73199(44)5.72439(30)5.49238(44)5.23798(44)4.5751(74)5.0311(15)5.2301(15)
42Mo 5.4151(22)5.4054(21)5.1778(21)4.91857(74)4.32423(40)4.72145(60)4.9179(31)
 5.41445(12)5.40663(12)5.17716(12)4.92327(30)4.3041(74)4.7191(15)4.9131(15)
43Tc 5.1228(20)5.1139(19)4.8880(19)4.6341(13) 4.4368(13)4.62991(94)
  5.11488(44)4.8874(12) 4.0581(74)4.4361(15)4.6301(15)
44Ru 4.8541(18)4.8449(17)4.6210(17)4.3681(13)3.8443(16)4.17814(78)4.36776(32)
 4.85388(10)4.845823(74)4.620649(44)4.37187(30)3.8351(74)4.1801(15)4.3691(15)
45Rh 4.6055(16)4.5966(16)4.3744(16)4.1277(12)3.6334(17)3.94053(55)4.12730(50)
 4.60552(13)4.59750(13)4.374206(59)4.13106(30)3.6291(74)3.94256(74)4.12996(74)
46Pd 4.3753(15)4.3672(15)4.1461(14)3.9088(10)3.43948(61)3.72251(52)3.90655(62)
 4.37595(10)4.367736(74)4.146282(74)3.908929(59)3.4371(15)3.72286(15)3.90746(15)
47Ag 4.1623(14)4.1541(13)3.9347(13)3.7034(10)3.25639(29)3.51704(26)3.69817(53)
 4.163002(74)4.154492(44)3.934789(44)3.703406(44)3.25645(15)3.51645(15)3.69996(15)
48Cd 3.9644(13)3.9560(12)3.7382(12)3.51355(97)3.08443(17)3.32528(29)3.50348(45)
 3.965020(89)3.956409(59)3.738286(59)3.514133(59)3.08495(15)3.32575(15)3.50475(15)
49In 3.7802(12)3.7716(11)3.5553(11)3.33796(83)2.92533(19)3.14784(47)3.32322(42)
 3.780787(89)3.771977(59)3.555363(59)3.338430(44)2.92604(15)3.14735(15)3.32375(15)
50Sn 3.6084(11)3.5997(10)3.38472(100)3.17475(77)2.776792(71)2.98309(56)3.15521(70)
 3.606964(59)3.599994(44)3.384921(44)3.175098(44)2.77694(15)2.98234(15)3.15575(15)
51Sb 3.44794(99)3.43913(93)3.22551(92)3.02325(67)2.638437(69)2.82990(51)2.99986(66)
 3.448452(89)3.439462(59)3.225718(59)3.023395(44)2.63884(15)2.82944(74)3.00035(15)
52Te 3.29788(92)3.28894(86)3.07663(85)2.88209(61)2.50998(50)2.687685(87)2.85523(35)
 3.29851(13)3.289249(89)3.076816(89)2.88221(12)2.50994(15)2.68794(15)2.85554(15)
53I 3.15734(85)3.14828(81)2.93720(78)2.75031(54)2.38965(37)2.55532(31)2.72067(32)
 3.157957(89)3.148647(89)2.937484(89)2.75057(12)2.38804(74)2.55424(74)2.71964(74)
54Xe 3.02568(78)3.01640(76)2.80659(69)2.62740(47)2.273869(70)2.427862(95)2.590303(89)
 3.025940(22)3.016582(15)2.806553(19) 2.27373(15)2.42924(15)2.59264(15)
55Cs 2.90167(73)2.89237(69)2.68362(66)2.51216(47)2.1676(29)2.3135(17)2.47326(16)
 2.90204(30)2.89244(30)2.68374(30)2.51184(30)2.16733(74)2.31393(15)2.47404(15)
56Ba 2.78522(68)2.77580(64)2.56812(61)2.40421(26)2.0697(15)2.20482(12)2.363082(97)
 2.785572(74)2.775992(74)2.568249(74)2.404386(89)2.06783(74)2.20483(15)2.36294(15)
57La 2.67563(64)2.66607(60)2.45941(57)2.30307(24)1.97705(28)2.10317(10)2.25958(20)
 2.675383(60)2.665740(74)2.458947(74)2.303312(98)1.97803(74)2.10533(74)2.2610(15)
58Ce 2.57122(59)2.56108(56)2.35598(53)2.20843(21)1.89320(71)2.01084(14)2.16586(39)
 2.57059(18)2.56163(17)2.35580(18)2.20900(17)1.89343(74)2.01243(74)2.1660(15)
59Pr 2.47329(55)2.46280(52)2.25890(49)2.11936(20)1.81477(33)1.92607(36)2.07945(22)
 2.47294(44)2.46304(30)2.25883(44)2.11943(59)1.81413(74)1.92553(74)2.07913(74)
60Nd 2.38081(51)2.36999(48)2.16724(45)2.03554(18)1.73904(18)1.84373(16)1.99616(19)
 2.38079(52)2.370526(16)2.167008(19)2.035448(88)1.73903(15)1.84403(15)1.99673(15)
61Pm 2.29340(48)2.28227(45)2.08060(42)1.95675(18)   
 2.29263(59)2.28223(44)2.07973(59)1.95593(89)1.66743(74)1.76763(74)1.91913(15)
62Sm 2.21054(48)2.19926(42)1.99850(42)1.88225(17)1.60201(12)1.69495(13)1.84534(42)
 2.210430(24)2.199873(13)1.998432(30)1.882206(41)1.60022(15)1.69533(15)1.84573(15)
63Eu 2.13214(42)2.12081(40)1.92080(37)1.81237(16)1.54065(17)1.62830(21)1.77767(16)
 2.13156(17)2.120673(95)1.92053(17)1.81215(17)1.53812(15)1.62712(15)1.77613(15)
64Gd 2.05817(40)2.04670(37)1.84744(34)1.74582(14)1.47922(25)1.56264(23)1.71092(21)
 2.05783(30)2.04683(30)1.84683(30)1.74553(30)1.47842(15)1.56322(15)1.71173(15)
65Tb 1.98699(37)1.97586(35)1.77701(32)1.68377(14)1.42285(98)1.50195(80)1.65023(44)
 1.98753(30)1.97653(30)1.77683(44)1.68303(30)1.42232(15)1.50232(15)1.64972(15)
66Dy 1.91986(35)1.90883(33)1.71052(30)1.62497(12)1.37058(41)1.44500(20)1.59241(33)
 1.919939(44)1.908839(44)1.71065(10)1.62371(10)1.36922(15)1.44452(15)1.59162(15)
67Ho 1.85606(33)1.84511(31)1.64732(28)1.56818(11)1.31957(28)1.39091(27)1.53614(34)
 1.856472(15)1.845092(17)1.647484(32)1.567168(50)1.31902(15)1.39052(15)1.53682(15)
68Er 1.79537(31)1.78449(29)1.58720(26)1.51486(10)1.27145(14)1.33792(26)1.48318(27)
 1.795701(45)1.784481(20)1.587466(86)1.51401(13)1.27062(15)1.33862(15)1.48352(15)
69Tm 1.73758(29)1.72677(27)1.52995(24)1.464210(95)1.22612(28)1.28942(27)1.43366(27)
 1.738003(19)1.7267720(70)1.5302410(70)1.46402(30)1.22502(15)1.28922(15)1.43342(15)
70Yb 1.68248(29)1.67177(26)1.47538(24)1.416041(89)1.18266(60)1.243391(70)1.3858(10)
 1.682875(74)1.671915(59)1.475672(74)1.415521(74)1.18182(15)1.24282(15)1.38622(15)
71Lu 1.63031(26)1.61949(24)1.42361(21)1.370061(85)1.14043(22)1.197954(60)1.341053(93)
 1.630314(74)1.619534(44)1.423611(44)1.370141(44)1.14022(15)1.19852(15)1.34052(15)
72Hf 1.58049(25)1.56959(23)1.37419(20)1.326241(78)1.10009(24)1.1550(10)1.2972(14)
 1.580484(74)1.569604(74)1.374121(74)1.326410(74)1.1002640(49)1.1548587(22)1.2971383(68)
73Ta 1.53290(23)1.52194(22)1.32697(19)1.282314(74)1.06152(30)1.11368(14)1.25506(34)
 1.532953(30)1.521993(30)1.327000(44)1.284559(30)1.06132(15)1.11372(15)1.25532(15)
74W 1.48748(22)1.47642(21)1.28188(18)1.244447(70)0.91604(28)1.07431(38)1.21543(99)
 1.487452(30)1.4763112(95)1.281812(13)1.2443048(98)1.024685(74)1.07452(15)1.21552(15)
75Re 1.44399(21)1.43288(19)1.23872(17)1.206487(67)0.98968(21)1.03670(20)1.17673(27)
 1.443982(74)1.432922(59)1.238599(30)1.206618(59)0.98941(15)1.03712(15)1.17732(15)
76Os 1.40238(20)1.39121(18)1.19742(16)1.170095(62)0.95583(36)1.000786(57)1.14002(23)
 1.402361(74)1.391231(74)1.197288(74)1.16981(12)0.95581(15)1.00142(15)1.14082(15)
77Ir 1.36252(19)1.35130(19)1.15786(15)1.135812(72)0.9240(12)0.96675(18)1.10535(22)
 1.362520(74)1.351300(44)1.157827(44)1.135337(44)0.92361(15)0.96711(15)1.10582(15)
78Pt 1.32434(18)1.31308(17)1.11995(14)1.102006(63)0.8933(14)0.93395(27)1.07200(36)
 1.324340(30)1.313060(44)1.119917(30)1.102017(44)0.893213(19)0.9341861(21)1.0722721(19)
79Au 1.28773(17)1.27643(16)1.08359(13)1.070479(53)0.86383(45)0.90263(12)1.04009(27)
 1.287739(44)1.276419(44)1.083546(44)1.070236(44)0.863683(30)0.9027409(46)1.0401625(52)
80Hg 1.25261(16)1.24126(15)1.04869(13)1.039584(51)0.83546(43)0.87238(26)1.00919(30)
 1.25266(10)1.241219(74)1.048696(74)1.03977(10)0.83531(15)0.87221(15)1.00912(15)
81Tl 1.21890(15)1.20750(14)1.01519(12)1.01029(20)0.80795(15)0.843512(77)0.97953(25)
 1.218768(44)1.207408(59)1.015145(59)1.010325(44)0.80811(15)0.84341(15)0.97931(15)
82Pb 1.18651(15)1.17507(14)0.98298(11)0.98221(19)0.78172(24)0.81575(18)0.95113(22)
 1.186498(74)1.175028(30)0.982925(44)0.98222(10)0.7818404(49)0.8157395(16)0.9511590(22)
83Bi2091.15540(14)1.14390(13)0.95205(11)0.95526(18)0.75649(58)0.789102(88)0.92387(11)
 1.155377(15)1.143877(30)0.951992(13)0.955194(59)0.75711(15)0.78871(15)0.92341(15)
84Po2091.12549(13)1.11393(12)0.92228(10)0.92932(18)0.7332(13)0.76325(13)0.897554(85)
 1.125497(74)1.113877(59)0.92201(30)0.929384(74)   
85At2101.09670(13)1.08510(12)0.893639(96)0.90444(17) 0.73868(13) 
 1.096726(74)1.085016(74)0.89350(13)    
86Rn2221.06900(12)1.05735(11)0.866054(91)0.88055(15) 0.71511(13) 
 1.069006(74)1.057246(74)0.86606(13)    
87Fr2231.04232(11)1.03063(11)0.839482(86)0.85751(15) 0.69240(13)0.8251(27)
 1.042316(74)1.030505(74)0.83941(13)0.8580(30)   
88Ra2261.01662(11)1.00489(10)0.813866(82)0.83533(16)0.64449(15)0.67077(12)0.802768(44)
 1.016575(74)1.004745(74)0.813762(74)0.835383(74)0.64451(15)0.67071(15)0.80281(15)
89Ac2270.99185(11)0.980070(98)0.789163(78)0.81406(14) 0.64970(13) 
 0.991795(74)0.979945(74)0.78904(13)    
90Th2320.96798(10)0.956154(94)0.765343(75)0.79354(13)0.60569(11)0.62966(11)0.760637(99)
 0.9679082(23)0.9560826(15)0.7652610(14)0.7935516(15)0.60591(15)0.62991(15)0.76071(15)
91Pa2310.944896(96)0.933002(90)0.742301(71)0.77321(12)0.58759(12)0.610354(92)0.740958(97)
 0.944834(74)0.932854(74)0.742331(74)0.77371(15)   
92U2380.922622(93)0.910674(86)0.720056(68)0.75462(12)0.569885(39)0.591930(66)0.722319(52)
 0.922572(13)0.910653(13)0.719995(12)0.754692(13)0.56951(15)0.59191(15)0.72231(15)
93Np2370.901230(88)0.889223(83)0.698624(65)0.73623(11)   
 0.901059(13)0.889141(13)0.698488(13)0.736241(13)0.55239(34)0.57368(37)0.704136(20)
94Pu2440.880355(85)0.868290(79)0.677776(60)0.71848(11)   
     0.53651(15)0.55721(15)0.68671(15)
95Am2430.860288(84)0.848190(81)0.657686(59)0.70134(10)   
        
96Cm2480.840918(80)0.828776(78)0.638265(56)0.684815(98)   
        
97Bk2490.822159(76)0.809987(69)0.619449(53)0.668638(94)   
     0.49060(49)0.50851(52)0.63748(98)
98Cf2500.803608(73)0.791421(66)0.601005(50)0.652873(89)   
     0.476569(92)0.493804(98)0.62300(19)
99Es2510.786043(70)0.773837(63)0.583354(49)0.638227(82)   
        
100Fm2540.769077(67)0.756843(60)0.566272(47)0.623826(82)   
 0.76904(62)0.75674(60)0.56619(34)0.62369(41)0.44966(13)0.46534(12)0.59414(20)
These values are for the unresolved [L\beta _2] and [L\beta _{15}] emission lines.
[Figure 4.2.2.2]

Figure 4.2.2.2| top | pdf |

Comparison of L-series data with experiment for the indicated range of Z. Indicated data, beginning at the bottom, refer to the Lα2, Lα1, and Lβ1 emission lines and the LIII, LII, and LI absorption edges. For clarity, the plots have been displaced vertically by multiples of 0.002 for the emission lines and 0.004 for the absorption edges.

4.2.2.13. Availability of a more complete X-ray wavelength table

| top | pdf |

This article and the accompanying X-ray wavelength tables are an updated version of the contribution to the International Tables for Crystallography, Volume C, 2nd edition that was published in 1999. This article has been subject to more critical review and analysis and the data are consistent with the most recent adjustment of the fundamental physical constants (Mohr & Taylor, 2000[link]). We believe that these data represent a significant improvement in consistency, coverage and accuracy over previously available resources. The results presented here are a subset of a larger effort that includes all K- and L-series lines connecting the n = 1 to n = 4 shells. The more complete table has been submitted for archival publication and will be made available on the NIST Physical Reference Data web site. Electronic publishing of this resource will provide a convenient data resource to the scientific community that can be more easily up-dated and expanded.

4.2.2.14. Connection with scales used in previous literature

| top | pdf |

In order to compare historical data for X-ray spectra with the results in the present tabulation, certain conversion factors are needed. As discussed in the introduction, the principal units found in the literature are the xu and the Å* unit. There is the additional complication that there were several different definitions in use at various times and at the same time in different laboratories. For the convenience of the reader, we summarize in Table 4.2.2.6[link] the main conversion factors needed. The numerical values for the wavelengths in Å can be converted to energies in electron volts by using the conversion factor 12398.41857 (49) eV Å (Mohr & Taylor, 2000[link]).

Table 4.2.2.6| top | pdf |
Wavelength conversion factors

Numbers in parentheses are standard uncertainties in the least-significant figures.

 Cu Kα1Mo Kα1W Kα1
λ (Å)1.54059292(45)0.70931713(41)0.20901313(18)
λ (Å*)1.540562(3)0.709300(1)0.2090100
l kxu1.5374000.707831 
Å*/Å1.0000201(20)1.0000242(22)1.00001498(86)
kxu/Å1.00207683(29)1.00209955(58) 

4.2.3. X-ray absorption spectra

| top | pdf |
D. C. Creaghb

4.2.3.1. Introduction

| top | pdf |

4.2.3.1.1. Definitions

| top | pdf |

This section deals with the manner in which the photon scattering and absorption cross sections of an atom varies with the energy of the incident photon. Further information concerning these cross sections and tables of the X-ray attenuation coefficients are given in Section 4.2.4[link]. Information concerning the anomalous-dispersion corrections is given in Section 4.2.6[link].

When a highly collimated beam of monoenergetic photons passes through a medium of thickness t, it suffers a decrease in intensity according to the relation [I=I_0\exp(-\mu_lt), \eqno (4.2.3.1)]where μl is the linear attenuation coefficient. Most tabulations express μl in c.g.s. units, μl having the units cm−1.

An alternative, often more convenient, way of expressing the decrease in intensity involves the measurement of the mass per unit area mA of the specimen rather than the specimen thickness, in which case equation (4.2.3.1)[link] takes the form [I=I_0\exp[-(\mu/\rho)m_A], \eqno (4.2.3.2)]where ρ is the density of the material and (μ/ρ) is the mass absorption coefficient. The linear attenuation coefficient of a medium comprising atoms of different types is related to the mass absorption coefficients by [\mu_l=\rho\textstyle\sum\limits_ig_i(\mu/\rho)_i, \eqno (4.2.3.3)]where [g_i] is the mass fraction of the atoms of the ith species for which the mass absorption coefficient is [(\mu/\rho)_i]. The summation extends over all the atoms comprising the medium. For a crystal having a unit-cell volume of [V_c], [\mu_l={1\over V_c}\sum\sigma_i, \eqno (4.2.3.4)]where [\sigma_i] is the photon scattering and absorption cross section. If [\sigma_i] is expressed in terms of barns/atom then [V_c] must be expressed in terms of Å3 and μl is in cm−1. (1 barn = 10−28 m2.)

The mass attenuation coefficient μ/ρ is related to the total photon–atom scattering cross section σ according to [\eqalignno{{\mu\over\rho}({\rm cm}^2/{\rm g}) &=(N_{\rm A}/M)\sigma ({\rm cm}^2/{\rm atom})\cr &=(N_{\rm A}/M)\times10^{-24}\sigma\,({\rm barns/atom}), &(4.2.3.5)}]where [N_{\rm A}=] Avogadro's number = 6.0221367 (36) × 1023 atoms/gram atom (Cohen & Taylor, 1987[link]) and M = atomic weight relative to M(12C) = 12.0000.

4.2.3.1.2. Variation of X-ray attenuation coefficients with photon energy

| top | pdf |

When a photon interacts with an atom, a number of different absorption and scattering processes may occur. For an isolated atom at photon energies of less than 100 keV (the limit of most conventional X-ray generators), contributions to the total cross section come from the photo-effect, coherent (Rayleigh) scattering, and incoherent (Compton) scattering. [\sigma=\sigma_{\rm pe}+\sigma_R+\sigma_C. \eqno (4.2.3.6)]The relation between the photo-effect absorption cross section [\sigma_{\rm pe}] and the X-ray anomalous-dispersion corrections will be discussed in Section 4.2.6[link].

The magnitudes of these scattering cross sections depend on the type of atom involved in the interactions and on the energy of the photon with which it interacts. In Fig. 4.2.3.1[link] , the theoretical cross sections for the interaction of photons with a carbon atom are given. Values of [\sigma_{\rm pe}] are from calculations by Scofield (1973[link]), and those for Rayleigh and Compton scattering are from tabulations by Hubbell & Øverbø (1979[link]) and Hubbell (1969[link]), respectively. Note the sharp discontinuities that occur in the otherwise smooth curves. These correspond to photon energies that correspond to the energies of the K and [L_{\rm I}] [L_{\rm II}] [L_{\rm III}] shells of the carbon energies. Notice also that [\sigma_{\rm pe}] is the dominant interaction cross section, and that the Rayleigh scattering cross section remains relatively constant for a broad range of photon energies, whilst the Compton scattering peaks at a particular photon energy (∼100 keV). Other interaction mechanisms exist [e.g. Delbrück (Papatzacos & Mort, 1975[link]; Alvarez, Crawford & Stevenson, 1958[link]), pair production, nuclear Thompson], but these do not become significant interaction processes for photon energies less than 1 MeV. This section will not address the interaction of photons with atoms for which the photon energy exceeds 100 keV.

[Figure 4.2.3.1]

Figure 4.2.3.1| top | pdf |

Theoretical cross sections for photon interactions with carbon showing the contributions of photoelectric, elastic (Rayleigh), inelastic (Compton), and pair-production cross sections to the total cross sections. Also shown are the experimental data (open circles). From Gerstenberg & Hubbell (1982[link]).

4.2.3.1.3. Normal attenuation, XAFS, and XANES

| top | pdf |

The curves shown in Fig. 4.2.3.1[link] are the result of theoretical calculations of the interactions of an isolated atom with a single photon. Experiments are not usually performed on isolated atoms, however. When experiments are performed on ensembles of atoms, a number of points of difference emerge between the experimental data and the theoretical calculations. These effects arise because the presence of atoms in proximity with one another can influence the scattering process. In short: the total attenuation coefficient of the system is not the sum of all the individual attenuation coefficients of the atoms that comprise the system.

Perhaps the most obvious manifestation of this occurs when the photon energy is close to an absorption edge of an atom. In Fig. 4.2.3.2[link] , the mass attenuation of several germanium compounds is plotted as a function of photon energy. The energy scale measures the distance from the K-shell edge energy of germanium (11.104 keV). These curves are taken from Hubbell, McMaster, Del Grande & Mallett (1974[link]). Not only does the experimental curve depart significantly from the theoretically predicted curve, but there is a marked difference in the complexity of the curves between the various germanium compounds.

[Figure 4.2.3.2]

Figure 4.2.3.2| top | pdf |

The dependence of the X-ray attenuation coefficient on energy for a range of germanium compounds, taken in the neighbourhood of the germanium absorption edge (from IT IV, 1974[link]).

Far from the absorption edge, the theoretical calculations and the experimental data are in reasonable agreement with what one might expect using the sum rule for the various scattering cross sections and one could say that this region is one in which normal attenuation coefficients may be found.

Closer to the edge, the almost periodic variation of the mass attenuation coefficient is called the extended X-ray absorption fine structure (XAFS). Very close to the edge, more complicated fluctuations occur. These are referred to as X-ray absorption near edge fine structure (XANES). The boundary of the XAFS and XANES regions is somewhat arbitrary, and the physical basis for making the distinction between the two will be outlined in Subsection 4.2.3.4[link].

Even in the region where normal attenuation may be thought to occur, cooperative effects can exist, which can affect both the Rayleigh and the Compton scattering contributions to the total attenuation cross section. The effect of cooperative Rayleigh scattering has been discussed by Gerward, Thuesen, Stibius-Jensen & Alstrup (1979[link]), Gerward (1981[link], 1982[link], 1983[link]), Creagh & Hubbell (1987[link]), and Creagh (1987a[link]). That the Compton scattering contribution depends on the physical state of the scattering medium has been discussed by Cooper (1985[link]).

Care must therefore be taken to consider the physical state of the system under investigation when estimates of the theoretical interaction cross sections are made.

4.2.3.2. Techniques for the measurement of X-ray attenuation coefficients

| top | pdf |

4.2.3.2.1. Experimental configurations

| top | pdf |

Experimental configurations that set out to determine the X-ray linear attenuation coefficient [\mu_l] or the corresponding mass absorption coefficients (μ/ρ) must have characteristics that reflect the underlying assumptions from which equation (4.2.3.1)[link] was derived, namely:

  • (i) the incident and transmitted beams are parallel and there is no divergence in the transmitted beam;

  • (ii) the photons in the incident and transmitted beams have the same energy;

  • (iii) the specimen is of sufficient thickness.

Because of the considerable discrepancies that often exist in X-ray attenuation measurements (see, for example, IT IV, 1974[link]), the IUCr Commission on Crystallographic Apparatus set up a project to determine which, if any, of the many techniques for the measurement of X-ray attenuation coefficients is most likely to yield correct results. In the project, a number of different experimental configurations were used. These are shown in Fig. 4.2.3.3[link] . The configurations used ranged in complexity from that of Fig. 4.2.3.3(a)[link], which uses a slit-collimated beam from a sealed tube and a β-filter to select its characteristic radiation, and a proportional counter and associated electronics to detect the transmitted-beam intensity, to that of Fig. 4.2.3.3[link](f), which uses a modification to a commercial X-ray-fluorescence analyser. Sources of X-rays included conventional sealed X-ray tubes, X-ray-fluorescence sources, radioisotope sources, and synchrotron-radiation sources. Detectors ranged from simple ionization chambers, which have no capacity for photon energy detection, to solid-state detectors, which provide a relatively high degree of energy discrimination. In a number of cases (Figs. 4.2.3.3c, d, e, and f[link]), monochromatization of the beam was effected using single Bragg reflection from silicon single crystals. In Fig. 4.2.3.3(i)[link], the incident-beam monochromator is using reflections from two Bragg reflectors tuned so as to eliminate harmonic radiation from the source.

[Figure 4.2.3.3]

Figure 4.2.3.3| top | pdf |

Schematic representations of experimental apparatus used in the IUCr X-ray Attenuation Project (Creagh & Hubbell, 1987[link]; Creagh, 1985[link]). X: characteristic line from sealed X-ray tube; b: Bremsstrahlung from a sealed X-ray tube; r: radioactive source; s: synchrotron-radiation source; β: β-filter for characteristic X-rays; S: collimating slits; M: monochromator.

The performance of these systems was evaluated for a range of materials that included:

  • (i) highly perfect silicon single crystals (Creagh & Hubbell, 1987[link]);

  • (ii) polycrystalline copper foils that exhibited a high degree of preferred orientation; and

  • (iii) pyrolytic graphite that contained a high density of regular voids.

The results of this study are outlined in Section 4.2.3.2.3[link].

4.2.3.2.2. Specimen selection

| top | pdf |

Although the most important component in the experiment is the specimen itself, examination of the data files held at the US National Institute of Standards and Technology (Gerstenberg & Hubbell, 1982[link]; Saloman & Hubbell, 1986[link]; Hubbell, Gerstenberg & Saloman, 1986[link]) has shown that, in general, insufficient care has been taken in the past to select an experimental device with characteristics that are appropriate to the specimen chosen. Nor has sufficient care been taken in the determination of the dimensions, homogeneity, and defect structure of the specimens. To achieve the best results, the following procedures should be followed.

  • (i) The dimensions of the specimen should be determined using at least two different techniques, and sample thicknesses should be chosen such that the Nordfors (1960[link]) criterion, later confirmed by Sears (1983[link]), that the condition [2\le\ln(I_0/I)\le4\eqno (4.2.3.7)]be satisfied. This enables the best compromise between achieving good counting statistics and avoiding multiple photon scattering within the sample.

    Wherever possible, different sample thicknesses should be chosen to enable a test of equation (4.2.3.1)[link] to be made. If deviations from equation (4.2.3.1)[link] exist, either the sample material or the experimental configuration, or both, are not appropriate for the measurement of [\mu_l]. If the attenuation of the material under test falls outside the limits set by the Nordfors criterion and the material is in the form of a powder, the mixing of this powder with one with low attenuation and no absorption edge in the region of interest can be used to bring the total attenuation of the sample within the Nordfors range.

  • (ii) The sample should be examined by as many means as possible to ascertain its regularity, homogeneity, defect structure, and, especially for very thin specimens, freedom from pinholes and cracks. Where a diluent has been used to reduce the attenuation so that the Nordfors criterion is satisfied, care must be taken to ensure intimate mixing of the two materials and the absence of voids.

    Since the theory upon which equation (4.2.3.1)[link] is based envisages that each atom scatters as an individual, it is necessary to be aware of whether such cooperative effects as Laue–Bragg scattering (which may become significant in single-crystal specimens) and small-angle X-ray scattering (SAXS) (which may occur if a distribution of small voids or inclusions exists) occur in polycrystalline and amorphous specimens. Knowledge that cooperative scattering may occur influences the choice of collimation of the beam.

  • (iii) The sample should be mounted normal to the beam.

4.2.3.2.3. Requirements for the absolute measurement of μl or (μ/ρ)

| top | pdf |

The following prescription should be followed if accurate, absolute measurements of [\mu_l] and [(\mu/\rho)] are to be obtained.

  • (i) X-ray source and X-ray monochromatization. The energy of the incident photons should be measured directly using reflections from a single-crystal silicon monochromator, and the energy spread of the beam should be measured. Measurements should be made of the state of polarization, since X-ray-polarization effects are known to be significant in some measurements (Templeton & Templeton, 1982[link], 1985a[link], 1986[link]). The results of a survey on X-ray polarization were given by Jennings (1984[link]). If a single-crystal monochromator is employed, it should be placed between the sample and the detector.

  • (ii) Collimation. It is of some advantage if both the incident-beam- and the transmitted-beam-defining slits can be varied in width.

    Should it be necessary to combat the effects of Laue–Bragg scattering in a single-crystal specimen, an incident beam with a high degree of collimation is required (Gerward, 1981[link]).

    To counter the effects of small-angle X-ray scattering, it may be necessary to widen the detector aperture (Chipman, 1969[link]). That these effects can be marked has been shown by Parratt, Porteus, Schnopper & Watanabe (1959[link]), who investigated the influence collimator and monochromator configurations have on X-ray-attenuation measurements.

  • (iii) Detection. Detectors that give some degree of energy discrimination should be used. Compromise may be necessary between sensitivity and energy resolution, however, and these factors should be taken into account when a choice is being made between proportional and solid-state detectors.

Whichever detection system is chosen, it is essential that the system dead-time be determined experimentally. For descriptions of techniques for the determination of system dead-time, see, for example, Bertin (1975[link]).

4.2.3.3. Normal attenuation coefficients

| top | pdf |

Fig. 4.2.3.1[link] shows that the X-ray attenuation coefficients are a smooth function of photon energy over a relatively large range of photon energies, and that discontinuities occur whenever the photon energy corresponds to a resonance in the electron cloud surrounding the nucleus. In Fig. 4.2.3.2[link], the effect of the interaction of the ejected photoelectron with the electron's neighbouring atoms is shown. Such edge effects (XAFS) can extend 1000 eV from the edge.

It is conventional, however, to extrapolate the smooth curve to the edge value, and a curve of normal attenuation coefficients results. These are taken to be the attenuation coefficients of the individual atoms. Tables of these normal attenuation coefficients are given in Section 4.2.4[link].

4.2.3.4. Attenuation coefficients in the neighbourhood of an absorption edge

| top | pdf |

4.2.3.4.1. XAFS

| top | pdf |

Although the existence of XAFS has been known for more than 60 years following experiments by Fricke (1920[link]) and Hertz (1920[link]), it is only in the last decade that a proper theoretical description has been developed. Kronig (1932a[link]) suggested a long-range-order theory based on quantum-mechanical precepts, although later (Kronig, 1932b[link]) he applied a short-range-order (SRO) theory to explain the existence of XAFS in molecular spectra. As time progressed, important suggestions were made by others, notably Kostarev (1941[link], 1949[link]), who applied this SRO theory to condensed matter, Sawada, Tsutsumi, Shiraiwa, Ishimura & Obashi (1959[link]), who accounted for the lifetime of the excited photoelectron and the core-hole state in terms of a mean free path, and Schmidt (1961a[link],b[link], 1963[link]), who showed the influence atomic vibrations have on the phase of the back-scattered waves.

Nevertheless, neither the experimental data nor the theories were sufficiently good to enable Azaroff & Pease (1974[link]) to decide which theory was the correct one to use. However, Sayers, Lytle & Stern (1970[link]) produced a theoretical approach based on SRO theory, later extended by Lytle, Stern & Sayers (1975[link]), and this is the foundation upon which all modern work has been built. Since 1970, a great deal of theoretical effort has been expended to improve the theory because of the need to interpret the wealth of data that became available through the increasing use of synchrotron-radiation sources in XAFS experiments.

A number of major reviews of XAFS theory and its use for the resolution of experimental data have been published. Contributions have been made by Stern, Sayers & Lytle (1975[link]), Lee, Citrin, Eisenberger & Kincaid (1981[link]), Lee (1981[link]), and Teo (1981[link]). The rapid growth of the use of synchrotron-radiation sources has led to the development of the use of XAFS in a wide variety of research fields. The XAFS community has met regularly at conferences, producing conference proceedings that demonstrate the maturation of the technique. The reader is directed to the proceedings edited by Mustre de Leon, Stern, Sayers, Ma & Rehr (1988[link]), Hasnain (1990[link]), and Kuroda, Ohta, Murata, Udagawa & Nomura (1992[link]), and to the papers contained therein. In the following section, a brief, simplified, description will be given of the theory of XAFS and of the application of that theory to the interpretation of XAFS data.

4.2.3.4.1.1. Theory

| top | pdf |

The theory that will be outlined here has evolved through the efforts of many workers over the past decade. The oscillatory part of the X-ray attenuation relative to the `background' absorption may be written as [\chi(E)={\mu_l(E)-\mu_{l0}(E)\over\mu_{l0}(E)}, \eqno (4.2.3.8)]where [\mu_l(E)] is the measured value of the linear attenuation coefficient at a photon energy E and [\mu_{l0}(E)] is the `background' linear attenuation coefficient. This is sometimes the extrapolation of the normal attenuation curve to the edge energy, although it is usually found necessary to modify this extrapolation somewhat to improve the matching of the higher-energy data with the XAFS data (Dreier, Rabe, Malzfeldt & Niemann, 1984[link]). In most computer programs, the normal attenuation curve is fitted to the data using cubic spline fitting routines.

The origin of XAFS lies in the interaction of the ejected photoelectron with electrons in its immediate vicinity. The wavelength of a photoelectron ejected when a photon is absorbed is given by λ = 2π/k, where [k=[(2m/\hbar^2)(E-E_0)]{}^{1/2}. \eqno (4.2.3.9)]

This outgoing spherical wave can be back-scattered by the electron clouds of neighbouring atoms. This back-scattered wave interferes with the outgoing wave, resulting in the oscillation of the absorption rate that is observed experimentally and called XAFS. Equation (4.2.3.8)[link] was written with the assumption that the absorption rate was directly proportional to the linear absorption coefficient.

It is conventional to express [\chi(E)] in terms of the momentum of the ejected electron, and the usual form of the theoretical expression for χ(k) is [\chi({\bf k})=\textstyle\sum\limits_i(N_i/kr^2_j)|\,f_i(k)|\exp(\sigma^2_i k^2-r_i/\rho)\sin[2kr_i + \varphi_i(k)]. \eqno (4.2.3.10)]Here the summation extends over the shells of atoms that surround the absorbing atom, [N_i] representing the number of atoms in the ith shell, which is situated a distance [r_i] from the absorbing atom. The back-scattering amplitude from this shell is [f_i(k)] for which the associated phase is [\varphi_i(k)]. Deviations due to thermal motions of the electrons are incorporated through a Debye–Waller factor, [\exp(-\sigma^2_ik^2)], and ρ is the mean free path of the electron.

The amplitude function [f_i(k)] depends only on the type of back-scattering atom. The phase, however, contains contributions from both the absorber and the back-scatterer: [\varphi^l_i(k)=\varphi^l_j(k)+\varphi_i(k)-l\pi,\eqno (4.2.3.11)]where l = 1 for K and [L_{\rm I}] edges, and l = 2 or 0 for [L_{\rm II}] and [L_{\rm III}] edges. The phase is sensitive to variations in the energy threshold, the magnitude of the effect being larger for small electron energies than for electrons with considerable kinetic energy, i.e. the effect is more marked in the neighbourhood of the absorption edge. Since the position of the edge varies somewhat for different compounds (Azaroff & Pease, 1974[link]), some impediment to the analysis of experimental data might occur, since the determination of the interatomic distance [r_i] depends upon the precise knowledge of the value of [\varphi_i(k)].

In fitting the experimental data based on an empirical value of threshold energy using theoretically determined phase shifts, the difference between the theoretical and the experimental threshold energies [\Delta E_0] cannot produce a good fit at an arbitrarily chosen distance [r_i], since the effect will be seen primarily at low k values [(\sim0.3r\Delta E_0/k)], whereas changing [r_i] affects [\varphi_i(k)] at high k values [(\sim2k\Delta r)]. This was first demonstrated by Lee & Beni (1977[link]).

The significance of the Debye–Waller factor [\exp(-\sigma^2_ik^2)] should not be underestimated in this type of investigation. In XAFS studies, one is seeking to determine information regarding such properties of the system as nearest- and next-nearest-neighbour distances and the number of nearest and next-nearest neighbours. The theory is a short-range-order theory, hence deviations of atoms from their expected positions will influence the analysis significantly. Thus, it is often of value, experimentally, to work at liquid-nitrogen temperatures to reduce the effect of atomic vibrations.

Two distinct types of disorder are observed: vibrational, where the atom vibrates about a mean position in the structure, and static, where the atom occupies a position not expected theoretically. These terms can be separated from one another if the variation of XAFS spectra with temperature is studied, because the two have different temperature dependences. A discussion of the effect of a thermally activated disorder that is large compared with the static order has been given by Sevillano, Meuth & Rehr (1978[link]). For systems with large static disorders, e.g. liquids and amorphous solids, equation (4.2.3.10)[link] has to be modified somewhat. The XAFS equation has to be averaged over the pair distribution function g(r) for the system: [\chi(k)={ F(k)\over k}\int\limits^\propto_0g(r)\exp(-2r/\rho){ \sin(2kr+\varphi_k)\over r^2}{\,\rm d}r. \eqno (4.2.3.12)]

Other factors that must be taken into account in XAFS analyses include: inelastic scattering (due to multiple scattering in the absorbing atom and excitations of the atoms surrounding the atom from which the photoelectron was ejected) and multiple scattering of the photoelectron. Should multiple scattering be significant, the simple model given in equation (4.2.3.10)[link] is inappropriate, and more complex models such as those proposed by Pendry (1983[link]), Durham (1983[link]), Gurman (1988[link], 1995[link]), Natoli (1990[link]), and Rehr & Albers (1990[link]) should be used. Several computer programs are now available commercially for use in personal computers (EXCURVE, FEFF5, MSCALC). Readers are referred to scientific journals to find how best to contact the suppliers of these programs.

4.2.3.4.1.2. Techniques of data analysis

| top | pdf |

Three assumptions must be made if XAFS data are to be used to provide accurate structural and chemical information:

  • (i) XAFS occurs through the interaction of waves singly scattered by neighbouring atoms;

  • (ii) the amplitude function of the atoms is insensitive to the type of chemical bond (the postulate of transferability), which implies that one can use the same amplitude function for a given atom in problems involving compounds of that atom, whatever the nature of its neighbours or the nature of the bond; and

  • (iii) the phase function can be transferred for each pair of absorber–back-scatterer atoms.

Of these three assumptions, (ii) is of the most questionable validity. See, for example, Stern, Bunker & Heald (1981[link]).

It is usual, when analysing XAFS data, to search the literature for, or make sufficient measurements of, [\mu_{l0}] remote from the absorption edge to produce a curve of [\mu_{l0}(E)] versus E that can be extrapolated to the position of the edge. From equation (4.2.3.8)[link], it is possible to produce a curve of [\chi(E)] versus E from which the variation of [\chi(k)] with k can be deduced using equation (4.2.3.9)[link].

It is also customary to multiply [\chi(k)] by some power of k to compensate for the damping of the XAFS amplitudes with increasing k. The power chosen is somewhat arbitrary but [k^3] is a commonly used weighting function.

Two different techniques may be used to analyse the new data set, the Fourier-transform technique or the curve-fitting technique.

In the Fourier-transform technique (FF), the Fourier transform of the [k^n\chi(k)] is determined for that region of momentum space from the smallest, [k_1], to the largest, [k_2], wavevectors of the photoelectron, yielding the radical distribution function [\rho_n(r')] in coordinate [(r')] space. [\rho_n(r')={ 1\over(2\pi)^{1/2}}\int\limits^{k_2}_{k_1}k^n\chi(k)\exp(i2kr'){\, \rm d}k. \eqno (4.2.3.13)]

The Fourier spectrum contains peaks indicating that the nearest-neighbour, next-nearest-neighbour, etc. distances will differ from the true spacings by between 0.2 to 0.5 Å depending on the elements involved. These position shifts are determined for model systems and then transferred to the unknown systems to predict interatomic spacings. Fig. 4.2.3.4[link] illustrates the various steps in the Fourier-transform analysis of XAFS data.

[Figure 4.2.3.4]

Figure 4.2.3.4| top | pdf |

Steps in the reduction of data from an XAFS experiment using the Fourier transform technique: (a) after the removal of background χ(k) versus k; (b) after multiplication by a weighting function (in this case k3); (c) after Fourier transformation to determine r′.

The technique works best for systems having well separated peaks. Its primary weakness as a technique lies in the fact that the phase functions are not linear functions of k, and the spacing shift will depend on [E_0], the other factors including the weighting of data before the Fourier transforms are made, the range of k space transformed, and the Debye–Waller factors of the system.

In the curve-fitting technique (CF), least-squares refinement is used to fit the spectra in k space using some structural model for the system. Such techniques, however, can only indicate which of several possible choices is more likely to be correct, and do not prove that that structure is the correct structure.

It is possible to combine the FF and CF techniques to simplify the data analysis. Also, for data containing single-scatter peaks, the phase and amplitude components can be separated and analysed separately using either theory or model compounds (Stern, Sayers & Lytle, 1975[link]).

Each XAFS data set depends on two sets of strongly correlated variables: [\{F(k),\sigma,\rho,N\}] and [\{\varphi(k),E_0,r\}]. The elements of each set are not independent of one another. To determine N and σ, one must know F(k) well. To determine r, [\varphi(k)] must be known accurately.

Attempts have been made by Teo & Lee (1979[link]) to calculate F(k) and [\varphi(k)] from first principles using an electron–atom scattering model. Parametrized versions have been given by Teo, Lee, Simons, Eisenberger & Kincaid (1977[link]) and Lee et al. (1981[link]). Claimed accuracies for r, σ, and N in XAFS determinations are 0.5, 10, and 20%, respectively.

Acceptable methods for data analysis must conform to a number of basic criteria to have any validity. Amongst these are the following:

  • (i) the data analysis must not give rise to systematic error in the sense that it must provide unbiased estimates of parameters;

  • (ii) the assumed (hypothetical) model must be able to describe the data adequately;

  • (iii) the number of parameters used to describe the best fit of data must not exceed the number of independent data points;

  • (iv) where multiple solutions exist, supplementary information or assumptions used to resolve the ambiguity must conform to the philosophy of choice of the model structure.

The techniques for estimation of the parameters must always be given, including all known sources of uncertainty.

A complete list of criteria for the correct analysis and presentation of XAFS data is given in the reports of the International Workshops on Standards and Criteria in XAFS (Lytle, Sayers & Stern, 1989[link]; Bunker, Hasnain & Sayers, 1990[link]).

4.2.3.4.1.3. XAFS experiments

| top | pdf |

The variety and number of experiments in which XAFS experiments have been used is so large that it is not possible here to give a comprehensive list. By consulting the papers given in such texts as those edited by Winick & Doniach (1980[link]), Teo & Joy (1981[link]), Bianconi, Incoccia & Stipcich (1983[link]), Mustre de Leon et al. (1988[link]), Hasnain (1990[link]), and Kuroda et al. (1992[link]), the reader may find references to a wide variety of experiments in fields of research ranging from archaeology to zoology.

In crystallography, XAFS experiments have been used to assist in the solution of crystal structures; the large variations in the atomic scattering factors can be used to help solve the phase problem. Helliwell (1984[link]) reviewed the use of these techniques in protein crystallography. A further discussion of the use of these anomalous-dispersion techniques in crystallography has been given by Creagh (1987b[link]). The relation that exists between the attenuation (related to the imaginary part of the dispersion correction, f′′) and intensity (related to the atomic form factor and the real part of the dispersion correction, f′) is discussed by Creagh in Section 4.2.6[link]. Specifically, modulations occur in the observed diffracted intensities from a specimen as the incident photon energy is scanned through the absorption edge of an atomic species present in the specimen. This technique, referred to as diffraction anomalous fine structure (DAFS) is complementary to XAFS. Because of the dependence of intensity on the geometrical structure factor, and the fact that the structure factor itself depends on the positional coordinates of the absorbing atom, it is possible to discriminate, in some favourable cases, between the anomalous scattering between atoms occupying different sites in the unit cell (Sorenson et al., 1994[link]).

In many systems of biological interest, the arrangement of radicals surrounding an active site must be found in order that the role of that site in biochemical processes may be assessed. A study of the XAFS spectrum of the active atom yields structural information that is specific to that site. Normal crystallographic techniques yield more general information concerning the crystal structure. An example of the use of XAFS in biological systems is the study of iron–sulfur proteins undertaken by Shulman, Weisenberger, Teo, Kincaid & Brown (1978[link]). Other, more recent, studies of biological systems include the characterization of the Mn site in the photosynthetic oxygen evolving complexes including hydroxylamine and hydroquinone (Riggs, Mei, Yocum & Penner-Hahn, 1993[link]) and an XAFS study with an in situ electrochemical cell on manganese Schiff-base complexes as a model of a photosystem (Yamaguchi et al., 1993[link]).

It must be stressed that the theoretical expression (equation 4.2.3.10[link]) does not take into account the state of polarization of the incident photon. Templeton & Templeton (1986[link]) have shown that polarization effects may be observed in some materials, e.g. sodium bromate. Given that most XAFS experiments are undertaken using the highly polarized radiation from synchrotron-radiation sources, it is of some importance to be aware of the possibility that dichroic effects may occur in some specimens.

Because XAFS is a short-range-order phenomenon, it is particularly useful for the structural study of such disordered systems as liquid metals and amorphous solids. The analysis of such disordered systems can be complicated, particularly in those cases where excluded-volume effects occur. Techniques for analysis for these cases have been suggested by Crozier & Seary (1980[link]). Fuoss, Eisenberger, Warburton & Bienenstock (1981[link]) suggested a technique for the investigation of amorphous solids, which they call the differential anomalous X-ray scattering (DAS) technique. This method has some advantages when compared with conventional XAFS methods because it makes more effective use of low-k information, and it does not depend on a knowledge of either the electron phase shifts or the mean free paths.

Both the conventional XAFS and DAS techniques may be used for studies of surface effects and catalytic processes such as those investigated by Sinfelt, Via & Lytle (1980[link]), Hida et al. (1985[link]), and Caballero, Villain, Dexpert, Le Peltier & Lynch (1993[link]).

It must be stressed that in all the foregoing discussion it has been assumed that the detection of XAFS has been by measurement of the linear attenuation coefficient of the specimen. However, the process of photon absorption followed by the ejection of a photoelectron has as its consequence both X-ray fluorescence and surface XAFS (SEXAFS) and Auger electron emission. All of these techniques are extremely useful in the analysis of dilute systems.

SEXAFS techniques are extremely sensitive to surface conditions since the mean free path of electrons is only about 20 Å. Discussions of the use of SEXAFS techniques have been given by Citrin, Eisenberger & Hewitt (1978[link]) and Stohr, Denley & Perfettii (1978[link]). A major review of the topic is given in Lee et al. (1981[link]). SEXAFS has the capacity of sensing thin films deposited on the surface of substrates, and has applications in experiments involving epitaxic growth and absorption by catalysts.

Fluorescence techniques are important in those systems for which the absorption of the specimen under investigation contributes only very slightly to the total attenuation coefficient since it detects the fluorescence of the absorbing atom directly. Experiments by Hastings, Eisenberger, Lengeler & Perlman (1975[link]) and Marcus, Powers, Storm, Kincaid & Chance (1980[link]) proved the importance of this technique in analysing dilute alloy and biological specimens. Materlik, Bedzyk & Frahm (1984[link]) have demonstrated its use in determining the location of bromine atoms absorbed on single-crystal silicon substrates. Oyanagi, Matsushita, Tanoue, Ishiguro & Kohra (1985[link]) and Oyanagi, Takeda, Matsushita, Ishiguro & Sasaki (1986[link]) have also used fluorescence XAFS techniques for the characterization of very thin films. More recently, Oyanagi et al. (1987[link]) have applied the technique to the study of short-range order in high-temperature superconductors. Oyanagi, Martini, Saito & Haga (1995[link]) have studied in detail the performance of a 19-element high-purity Ge solid-state detector array for fluorescence X-ray absorption fine structure studies.

A less-sensitive technique, but one that can be usefully employed for thin-film studies, is that in which XAFS modulations are detected in the beam reflected from a sample surface. This technique, ReflEXAFS, has been used by Martens & Rabe (1980[link]) to investigate superficial regions of copper oxide films by means of reflection of the X-rays close to the critical angle for total reflection.

If a thin film is examined in a transmission electron microscope, the electron beam loses some of its kinetic energy in interactions between the electron beam and the electrons within the film. If the resultant energy loss is analysed using a magnetic analyser, XAFS-like modulations are observed in the electron energy spectrum. These modulations, electron-energy-loss fine structure (EELS), which were first observed in a conventional transmission electron microscope by Leapman & Cosslett (1976[link]), are now used extensively for microanalyses of light elements incorporated into heavy-element matrices. Most major manufacturers of transmission electron microscopes supply electron-energy-loss spectrometers for their machines. There are more problems in analysing electron-energy-loss spectra than there are for XAFS spectra. Some of the difficulties encountered in producing reliable techniques for the routine analysis of EELS have been outlined by Joy & Maher (1985[link]). This matter is discussed more fully in §4.3.4.4.2[link] .

A more recent development has been the observation of topographic XAFS (Bowen, Stock, Davies, Pantos, Birnbaum & Chen, 1984[link]). This fine structure is observed in white-beam topographs taken using synchrotron-radiation sources. The technique provides the means of simultaneously determining spatially resolved microstructural and spectroscopic information for the specimen under investigation.

In all the preceding discussion, however, the electron was assumed to undergo only single-scattering processes. If multiple scattering occurs, then the theory has to be changed somewhat. §4.2.3.4.2[link] discusses the effect of multiple scattering.

4.2.3.4.2. X-ray absorption near edge structure (XANES)

| top | pdf |

In Fig. 4.2.3.2(c)[link], there appears to be one cycle of strong oscillation in the neighbourhood of the absorption edge before the quasi-periodic variation of the XAFS commences. The electrons that cause this strong modulation of the photoelectric scattering cross section have low k values, and the electron is strongly scattered by neighbouring atoms. It was mentioned in §4.2.3.4.1[link] that conventional XAFS theory assumes a weak, single-scattering interaction between the ejected photoelectron and its environment. A schematic diagram illustrating the difference between single- and multiple-scattering processes is given in Fig. 4.2.3.5[link] . Evidently, the multiple-scattering process is very complicated and a discussion of the theory of XANES is too complex to be given here. The reader is directed to papers by Pendry (1983[link]), Lee (1981[link]), and Durham (1983[link]). A more recent review of the study of fine structure in ionization cross sections and their use in surface science has been given by Woodruff (1986[link]).

[Figure 4.2.3.5]

Figure 4.2.3.5| top | pdf |

Schematic representations of the scattering processes undergone by the ejected photoelectron in the single-scattering (XAFS) case and the full multiple-scattering regime (XANES).

The data from XANES experiments can be analysed to determine structural information such as coordination geometry, the symmetry of unoccupied valence electronic states, and the effective charge on the absorbing atom (Natoli, Misemer, Doniach & Kutzler, 1980[link]; Kutzler, Natoli, Misemer, Doniach & Hodgson, 1981[link]). XANES experiments have been performed to resolve many problems, inter alia: the origin of white lines (Lengeler, Materlik & Müller, 1983[link]); absorption of gases on metal surfaces (Norman, Durham & Pendry, 1983[link]); the effect of local symmetry in 3d elements (Petiau & Calas, 1983[link]); and the determination of valence states in materials (Lereboures, Dürr, d'Huysser, Bonelle & Lenglet, 1980[link]).

4.2.3.5. Comments

| top | pdf |

For reliable experiments using XAFS and XANES to be undertaken, intense-radiation sources must be used. Synchrotron-radiation sources are such a source of highly intense X-rays. Their ready availability to experimenters and the comparative simplicity of the equipment required to perform the experiments have made experiments involving XAFS and XANES very much easier to perform than has hitherto been the case.

At some synchrotron-radiation sources, database and program libraries for the storage and analysis of XAFS and XANES data exist. These are usually part of the general computing facilities (Pantos, 1982[link]).

Crystallographers seeking information concerning the nature and extent of these computer facilities can find such information by contacting the computer centre at one of the synchrotron-radiation establishments listed in Table 4.2.3.1[link].

Table 4.2.3.1| top | pdf |
Some synchrotron-radiation facilities providing XAFS databases and analysis utilities

CountrySynchrotron sourceAddress
FranceLUREUniversité Paris-Sud, LURE, 91405 Orsay, France
ItalyFrascatiLaboratori Nationali di Frascati, CP 13, 00044 Frascati, Italy
JapanPhoton FactoryPhoton Factory, National Laboratory for High Energy Physics, 1-1 Oho, Tsukuba-gun, Ibaraki 305, Japan
GermanyDESYDESY, Notkestrasse 85, 2000 Hamburg 52, Germany
United KingdomSRC/DaresburyDaresbury Laboratory, Daresbury, Warrington WA4 4AD, England
USACHESSCHESS, Cornell University, Ithaca, New York 14853, USA
NSLSNSLS, Brookhaven National Laboratory, Upton, New York 11973, USA
SPEARSSRL, Stanford University, Bin 69, PO Box 4349, Stanford, California 94305, USA

4.2.4. X-ray absorption (or attenuation) coefficients

| top | pdf |
D. C. Creaghb and J. H. Hubbelld

4.2.4.1. Introduction

| top | pdf |

This data set is intended to supersede those data sets given in International Tables for X-ray Crystallography, Vols. III (Koch, MacGillavry & Milledge, 1962[link]) and IV (Hubbell, McMaster, Del Grande & Mallett, 1974[link]).

It is not intended here to give a detailed bibliography of experimental data that have been obtained in the past 90 years. This has been the subject of a number of publications, e.g. Saloman & Hubbell (1987[link]), Hubbell, Gerstenberg & Saloman (1986[link]), Saloman & Hubbell (1986[link]), and Saloman, Hubbell & Scofield (1988[link]). Further commentary on the validity and the quality of the experimental data in existing tabulations has been given by Creagh & Hubbell (1987[link]) and Creagh (1987a[link]).

Existing tabulations of X-ray attenuation (or absorption) cross sections fall into three distinct categories: purely theoretical, purely experimental, and an evaluated mixture of theoretical and experimental data.

Compilations of the purely theoretically derived data exist for:

  • photo-effect absorption cross sections (Storm & Israel, 1970[link]; Cromer & Liberman, 1970[link]; Scofield, 1973[link]; Hubbell, Veigele, Briggs, Brown, Cromer & Howerton, 1975[link]; Band, Kharitonov & Trzhaskovskaya, 1979[link]; Yeh & Lindau, 1985[link]);

  • Compton scattering cross sections (Hubbell et al., 1975[link]);

  • Rayleigh scattering cross sections (Hubbell et al., 1975[link]; Hubbell & Øverbø, 1979[link]; Schaupp, Schumacher, Smend, Rullhusen & Hubbell, 1983[link]).

Many purely experimental compilations exist, and the cross-section data given in computer programs used in the analysis of results in X-ray-fluorescence spectroscopy, electron-probe microanalysis, and X-ray diffraction are usually (evaluated) compilations of several of the following compilations: Allen (1935[link], 1969[link]), Victoreen (1949[link]), Liebhafsky, Pfeiffer, Winslow & Zemany (1960[link]), Koch et al. (1962[link]), Heinrich (1966[link]), Theisen & Vollath (1967[link]), Veigele (1973[link]), Leroux & Thinh (1977[link]), Montenegro, Baptista & Duarte (1978[link]), and Plechaty, Cullen & Howerton (1981[link]). If a comparison is made between these data sets, significant discrepancies are found, and questions must be asked concerning the reliability of the data sets that are compared. Jackson & Hawkes (1981[link]) and Gerward (1986[link]) have produced sets of parametric tables to simplify the application of X-ray attenuation data for the solution of problems in computer-aided tomography and X-ray-fluorescence analysis.

Compilations by Henke, Lee, Tanaka, Shimambukuro & Fujikawa (1982[link]) and the earlier tables of McMaster, Del Grande, Mallett & Hubbell (1969/1970[link]) are examples of the judicious application of both theoretical and experimental data to produce a comprehensive data set of X-ray interaction cross sections.

Because of the discrepancies that appear to exist between experimental data sets, the IUCr Commission on Crystallographic Apparatus set up a project to establish which, if any, of the existing methods for measuring X-ray interaction cross sections (X-ray attenuation coefficients) and which theoretical calculations could be considered to be the most reliable. A discussion of some of the major results of this project is given in Section 4.2.3[link]. A more detailed description of this project has been given by Creagh & Hubbell (1987[link], 1990[link]).

In this section, tabulations of the total X-ray interaction cross sections σ and the mass absorption coefficient [\mu_m] are given for a range of characteristic X-ray wavelengths [Ti Kα 2.7440 Å (or 4.509 keV) to Ag Kβ 0.4470 Å (or 24.942 keV)]. The interaction cross sections are expressed in units of barns/atom (1 barn = 10−28 m2) whilst the mass absorption coefficient is given in cm2 g−1. Table 4.2.4.1[link] sets out the wavelengths of the characteristic wavelengths used in Tables 4.2.4.2[link] and 4.2.4.3[link], which list values of σ and [\mu_m], respectively.

Table 4.2.4.1| top | pdf |
Table of wavelengths and energies for the characteristic radiations used in Tables 4.2.4.2[link] and 4.2.4.3[link]

Radiationλ (Å)E (keV)
Ag [K \bar \alpha]0.560822.103
Ag [K \beta _1]0.497024.942
Pd [K \bar \alpha]0.586921.125
Pd [K \beta _1]0.520523.819
Rh [K \bar \alpha]0.614720.169
Rh [K \beta _1]0.545622.724
Mo [K \bar \alpha]0.710717.444
Mo [K \beta _1]0.632319.608
Zn [K \bar \alpha]1.43648.631
Zn [K \beta _1]1.29529.572
Cu [K \bar \alpha]1.54188.041
Cu [K \beta _1]1.39228.905
Ni [K \bar \alpha]1.65917.472
Ni [K \beta _1]1.50018.265
Co [K \bar \alpha]1.79056.925
Co [K \beta _1]1.62087.629
Fe [K \bar \alpha]1.93736.400
Fe [K \beta _1]1.75657.038
Mn [K \bar \alpha]2.10315.895
Mn [K \beta _1]1.91026.490
Cr [K \bar \alpha]2.29095.412
Cr [K \beta _1]2.08485.947
Ti [K \bar \alpha]2.74964.509
Ti [K \beta _1]2.51384.932

Table 4.2.4.2| top | pdf |
Total phonon interaction cross section (barns/atom)

RadiationEnergy (MeV)12345678
HydrogenHeliumLithiumBerylliumBoronCarbonNitrogenOxygen
Ag K[\beta _1]2.494E−026.10E−011.26E+002.01E+002.97E+004.40E+006.59E+001.00E+011.52E+01
Pd K[\beta _1]2.382E−026.10E−011.26E+002.01E+002.99E+004.44E+006.68E+001.02E+011.55E+01
Rh K[\beta _1]2.272E−026.12E−011.27E+002.04E+003.06E+004.47E+006.78E+001.05E+011.62E+01
Ag K[\bar {\alpha }]2.210E−026.14E−011.28E+002.06E+003.13E+004.79E+007.45E+001.17E+011.82E+01
Pd K[\bar {\alpha }]2.112E−026.16E−011.29E+002.09E+003.23E+005.05E+008.02E+001.28E+012.02E+01
Rh K[\bar {\alpha }]2.017E−026.18E−011.30E+002.13E+003.35E+005.35E+008.68E+001.41E+012.25E+01
Mo K[\beta _1]1.961E−026.19E−011.31E+002.16E+003.42E+005.56E+009.14E+001.50E+012.41E+01
Mo K[\bar {\alpha }]1.744E−026.24E−011.34E+002.28E+003.83E+006.61E+001.15E+011.96E+013.25E+01
Zn K[\beta _1]9.572E−036.47E−011.69E+004.19E+001.07E+012.54E+015.37E+011.03E+021.80E+02
Cu K[\beta _1]8.905E−036.50E−011.78E+004.74E+001.28E+013.10E+016.64E+011.27E+022.24E+02
Zn K[\bar {\alpha }]8.631E−036.51E−011.82E+005.02E+001.38E+013.39E+017.28E+011.40E+022.46E+02
Ni K[\beta _1]8.265E−036.53E−011.89E+005.46E+001.54E+013.83E+018.28E+011.59E+022.80E+02
Cu K[\bar {\alpha}]8.041E−036.55E−011.94E+005.76E+001.66E+014.15E+018.99E+011.73E+023.04E+02
Co K[\beta _1]7.649E−036.58E−012.04E+006.40E+001.90E+014.80E+011.04E+022.01E+023.54E+02
Ni K[\bar {\alpha}]7.472E−036.59E−012.09E+006.73E+002.02E+015.14E+011.12E+022.16E+023.80E+02
Fe K[\beta _1]7.058E−036.63E−012.23E+007.65E+002.37E+016.09E+011.33E+022.57E+024.51E+02
Co K[\bar {\alpha}]6.925E−036.64E−012.28E+007.99E+002.50E+016.45E+011.41E+022.72E+024.78E+02
Mn K[\beta _1]6.490E−036.69E−012.48E+009.34E+003.01E+017.84E+011.72E+023.31E+025.81E+02
Fe K[\bar {\alpha}]6.400E−036.70E−012.53E+009.67E+003.13E+018.18E+011.79E+023.46E+026.06E+02
Cr K[\beta _1]5.947E−036.77E−012.83E+001.16E+013.88E+011.02E+022.24E+024.32E+027.56E+02
Mn K[\bar {\alpha}]5.895E−036.78E−012.87E+001.19E+013.99E+011.05E+022.30E+024.44E+027.76E+02
Cr K[\bar {\alpha}]5.412E−036.89E−013.31E+001.50E+015.14E+011.36E+022.99E+025.75E+021.00E+03
Ti K[\beta _1]4.932E−037.04E−013.94E+001.94E+016.82E+011.81E+023.98E+027.62E+021.33E+03
Ti K[\bar {\alpha}]4.509E−037.24E−014.73E+002.51E+018.97E+012.39E+025.23E+021.00E+031.73E+03

RadiationEnergy (MeV)910111213141516
FluorineNeonSodiumMagnesiumAluminiumSiliconPhosphorusSulfur
Ag K[\beta _1]2.494E−022.27E+013.33E+014.77E+016.68E+019.16E+011.23E+021.62E+022.10E+02
Pd K[\beta _1]2.382E−022.32E+013.40E+014.88E+016.85E+019.40E+011.26E+021.67E+022.16E+02
Rh K[\beta _1]2.272E−022.50E+013.62E+015.07E+018.26E+011.05E+021.40E+021.85E+022.02E+02
Ag K[\bar {\alpha }]2.210E−022.77E+014.12E+015.96E+018.42E+011.16E+021.56E+022.06E+022.67E+02
Pd K[\bar {\alpha }]2.112E−023.11E+014.65E+016.75E+019.55E+011.32E+021.78E+022.35E+023.05E+02
Rh K[\bar {\alpha }]2.017E−023.50E+015.26E+017.67E+011.09E+021.51E+022.03E+022.69E+023.49E+02
Mo K[\beta _1]1.961E−023.76E+015.68E+015.30E+011.18E+021.63E+022.20E+022.92E+023.78E+02
Mo K[\bar {\alpha }]1.744E−025.15E+017.86E+011.16E+021.65E+022.29E+023.10E+024.10E+025.32E+02
Zn K[\beta _1]9.572E−032.95E+024.57E+026.77E+029.67E+021.34E+031.79E+032.36E+033.03E+03
Cu K[\beta _1]8.905E−033.66E+025.67E+028.39E+021.20E+031.65E+032.21E+032.90E+033.72E+03
Zn K[\bar {\alpha }]8.631E−034.02E+026.22E+029.20E+021.31E+031.81E+032.42E+033.17E+034.06E+03
Ni K[\beta _1]8.265E−034.58E+027.08E+021.05E+031.49E+032.05E+032.75E+033.59E+034.60E+03
Cu K[\bar {\alpha }]8.041E−034.98E+027.68E+021.14E+031.61E+032.22E+032.97E+033.88E+034.97E+03
Co K[\beta _1]7.649E−035.78E+028.92E+021.32E+031.87E+032.57E+033.43E+034.48E+035.72E+03
Ni K[\bar {\alpha }]7.472E−036.20E+029.56E+021.41E+032.00E+032.75E+033.67E+034.78E+036.11E+03
Fe K[\beta _1]7.058E−037.36E+021.13E+031.67E+032.36E+033.24E+034.32E+035.62E+037.17E+03
Co K[\bar {\alpha }]6.925E−037.79E+021.20E+031.76E+032.50E+033.42E+034.56E+035.93E+037.56E+03
Mn K[\beta _1]6.490E−039.46E+021.45E+032.13E+033.02E+034.13E+035.49E+037.12E+039.06E+03
Fe K[\bar {\alpha }]6.400E−039.86E+021.51E+032.22E+033.14E+034.29E+035.71E+037.41E+039.42E+03
Cr K[\beta _1]5.947E−031.23E+031.88E+032.75E+033.88E+035.30E+037.03E+039.10E+031.15E+04
Mn K[\bar {\alpha }]5.895E−031.26E+031.93E+032.83E+033.98E+035.43E+037.20E+039.33E+031.18E+04
Cr K[\bar {\alpha }]5.412E−031.62E+032.48E+033.62E+035.09E+036.93E+039.16E+031.18E+041.50E+04
Ti K[\beta _1]4.932E−032.14E+033.26E+034.74E+036.64E+039.01E+031.19E+041.53E+041.93E+04
Ti K[\bar{\alpha}]4.509E−032.79E+034.23E+036.13E+038.57E+031.16E+041.52E+041.95E+042.45E+04

RadiationEnergy (MeV)1718192021222324
ChlorineArgonPotassiumCalciumScandiumTitaniumVanadiumChromium
Ag K[\beta _1]2.494E−022.68E+023.36E+024.17E+025.12E+026.20E+027.44E+028.85E+021.04E+03
Pd K[\beta _1]2.382E−022.75E+023.45E+024.29E+025.26E+026.37E+027.64E+029.09E+021.07E+03
Rh K[\beta _1]2.272E−023.15E+023.84E+024.84E+026.30E+027.25E+028.63E+029.98E+021.19E+03
Ag K[\bar {\alpha }]2.210E−023.41E+024.29E+025.32E+026.52E+027.89E+029.47E+021.12E+031.33E+03
Pd K[\bar {\alpha }]2.112E−023.89E+024.89E+026.06E+027.42E+028.99E+021.08E+031.28E+031.51E+03
Rh K[\bar {\alpha }]2.017E−024.45E+025.59E+026.93E+028.48E+021.03E+031.23E+031.46E+031.72E+03
Mo K[\beta _1]1.961E−024.83E+026.06E+027.52E+029.20E+021.11E+031.33E+031.58E+031.86E+03
Mo K[\bar {\alpha }]1.744E−026.78E+028.51E+021.05E+031.29E+031.56E+031.86E+032.20E+032.58E+03
Zn K[\beta _1]9.572E−033.82E+034.74E+035.80E+037.02E+038.38E+039.93E+031.16E+041.33E+04
Cu K[\beta _1]8.905E−034.68E+035.80E+037.09E+038.57E+031.02E+041.21E+041.41E+041.60E+04
Zn K[\bar {\alpha }]8.631E−035.11E+036.34E+037.73E+039.34E+031.11E+041.32E+041.53E+041.73E+04
Ni K[\beta _1]8.265E−035.78E+037.15E+038.72E+031.05E+041.25E+041.48E+041.72E+041.96E+04
Cu K[\bar {\alpha }]8.041E−036.24E+037.72E+039.40E+031.13E+041.35E+041.39E+041.85E+042.13E+04
Co K[\beta _1]7.649E−037.17E+038.86E+031.08E+041.30E+041.54E+041.80E+042.11E+042.53E+04
Ni K[\bar {\alpha }]7.472E−037.66E+039.46E+031.15E+041.38E+041.64E+041.91E+042.25E+042.74E+04
Fe K[\beta _1]7.058E−038.97E+031.11E+041.34E+041.61E+041.91E+042.20E+042.62E+043.32E+04
Co K[\bar {\alpha }]6.925E−039.46E+031.17E+041.41E+041.69E+042.01E+042.31E+042.75E+043.53E+04
Mn K[\beta _1]6.490E−031.13E+041.39E+041.69E+042.01E+042.39E+042.74E+043.26E+044.15E+04
Fe K[\bar {\alpha }]6.400E−031.18E+041.44E+041.75E+042.09E+042.47E+042.85E+043.37E+044.25E+04
Cr K[\beta _1]5.947E−031.44E+041.76E+042.13E+042.54E+043.01E+043.53E+044.05E+045.79E+03
Mn K[\bar {\alpha }]5.895E−031.47E+041.80E+042.18E+042.60E+043.08E+043.63E+044.14E+045.93E+03
Cr K[\bar {\alpha }]5.412E−031.86E+042.27E+042.74E+043.26E+043.85E+044.69E+046.32E+037.50E+03
Ti K[\beta _1]4.932E−032.38E+042.91E+043.49E+044.15E+044.87E+046.79E+038.16E+039.68E+03
Ti K[\bar {\alpha }]4.509E−033.01E+043.69E+044.41E+045.20E+046.03E+048.68E+031.04E+041.24E+04

RadiationEnergy (MeV)2526272829303132
ManganeseIronCobaltNickelCopperZincGalliumGermanium
Ag K[\beta _1]2.494E−021.22E+031.42E+031.64E+031.88E+032.14E+032.43E+032.74E+033.08E+03
Pd K[\beta _1]2.382E−021.25E+031.46E+031.68E+031.93E+032.20E+032.49E+032.81E+033.16E+03
Rh K[\beta _1]2.272E−021.37E+031.65E+031.89E+032.19E+032.49E+032.88E+033.21E+033.55E+03
Ag K[\bar {\alpha }]2.210E−021.55E+031.80E+032.07E+032.38E+032.71E+033.07E+033.46E+033.87E+03
Pd K[\bar {\alpha }]2.112E−021.76E+032.04E+032.35E+032.70E+033.07E+033.47E+033.91E+034.38E+03
Rh K[\bar {\alpha }]2.017E−022.01E+032.33E+032.68E+033.07E+033.49E+033.95E+034.44E+034.97E+03
Mo K[\beta _1]1.961E−022.17E+032.52E+032.90E+033.31E+033.77E+034.26E+034.80E+035.37E+03
Mo K[\bar {\alpha }]1.744E−023.02E+033.49E+034.01E+034.57E+035.18E+035.86E+036.60E+037.38E+03
Zn K[\beta _1]9.572E−031.55E+041.78E+042.02E+042.27E+042.53E+043.90E+034.46E+035.08E+03
Cu K[\beta _1]8.905E−031.88E+042.15E+042.43E+042.72E+044.13E+034.75E+035.44E+036.19E+03
Zn K[\bar {\alpha }]8.631E−032.05E+042.34E+042.63E+042.94E+044.50E+035.18E+035.92E+036.75E+03
Ni K[\beta _1]8.265E−032.29E+042.61E+042.93E+044.41E+035.07E+035.83E+036.67E+037.60E+03
Cu K[\bar {\alpha }]8.041E−032.46E+042.80E+043.14E+044.76E+035.47E+036.29E+037.19E+038.19E+03
Co K[\beta _1]7.649E−032.80E+043.17E+044.71E+035.46E+036.27E+037.21E+038.24E+039.38E+03
Ni K[\bar {\alpha }]7.472E−032.97E+043.35E+045.02E+035.82E+036.68E+037.68E+038.79E+031.00E+04
Fe K[\beta _1]7.058E−033.42E+045.04E+035.87E+036.80E+037.81E+038.98E+031.03E+041.17E+04
Co K[\bar {\alpha }]6.925E−033.58E+045.31E+036.18E+037.16E+038.23E+039.46E+031.08E+041.23E+04
Mn K[\beta _1]6.490E−035.40E+036.34E+037.39E+038.56E+039.83E+031.13E+041.29E+041.47E+04
Fe K[\bar {\alpha }]6.400E−035.62E+036.59E+037.68E+038.89E+031.02E+041.17E+041.34E+041.53E+04
Cr K[\beta _1]5.947E−036.87E+038.06E+039.40E+031.09E+041.25E+041.43E+041.64E+041.86E+04
Mn K[\bar {\alpha }]5.895E−037.04E+038.26E+039.62E+031.11E+041.28E+041.47E+041.68E+041,91E+04
Cr K[\bar {\alpha }]5.412E−038.90E+031.04E+041.22E+041.41E+041.61E+041.85E+042.12E+042.41E+04
Ti K[\beta _1]4.932E−031.15E+041.35E+041.57E+041.81E+042.08E+042.39E+042.72E+043.09E+04
Ti K[\bar {\alpha }]4.509E−031.47E+041.72E+042.00E+042.31E+042.65E+043.04E+043.46E+043.93E+04

RadiationEnergy (MeV)3334353637383940
ArsenicSeleniumBromineKryptonRubidiumStrontiumYttriumZirconium
Ag K[\beta _1]2.494E−023.44E+033.84E+034.26E+034.72E+035.21E+035.72E+036.25E+036.79E+03
Pd K[\beta _1]2.382E−023.53E+033.94E+034.37E+034.84E+035.34E+035.86E+036.41E+036.96E+03
Rh K[\beta _1]2.272E−024.04E+034.53E+035.00E+035.50E+035.98E+036.48E+037.27E+037.80E+03
Ag K[\bar {\alpha }]2.210E−024.33E+034.82E+035.35E+035.92E+036.52E+037.15E+037.80E+038.47E+03
Pd K[\bar {\alpha }]2.112E−024.89E+035.45E+036.04E+036.68E+037.35E+038.06E+038.79E+039.52E+03
Rh K[\bar {\alpha }]2.017E−025.55E+036.18E+036.83E+037.55E+038.30E+039.09E+039.90E+031.07E+04
Mo K[\beta _1]1.961E−025.99E+036.66E+037.36E+038.13E+038.94E+039.78E+031.06E+041.15E+04
Mo K[\bar {\alpha }]1.744E−028.22E+039.11E+031.00E+041.10E+041.21E+041.32E+041.43E+042.47E+03
Zn K[\beta _1]9.572E−035.77E+036.52E+037.34E+038.24E+039.21E+031.03E+041.14E+041.26E+04
Cu K[\beta _1]8.905E−037.03E+037.94E+038.94E+031.00E+041.12E+041.25E+041.39E+041.54E+04
Zn K[\bar {\alpha }]8.631E−037.65E+038.64E+039.73E+031.09E+041.22E+041.36E+041.51E+041.67E+04
Ni K[\beta _1]8.265E−038.62E+039.73E+031.10E+041.23E+041.37E+041.53E+041.70E+041.88E+04
Cu K[\bar {\alpha }]8.041E−039.29E+031.05E+041.18E+041.32E+041.48E+041.65E+041.83E+042.03E+04
Co K[\beta _1]7.649E−031.06E+041.20E+041.35E+041.52E+041.69E+041.89E+042.09E+042.32E+04
Ni K[\bar {\alpha }]7.472E−031.13E+041.28E+041.44E+041.61E+041.80E+042.01E+042.23E+042.47E+04
Fe K[\beta _1]7.058E−031.32E+041.49E+041.68E+041.88E+042.10E+042.34E+042.60E+042.87E+04
Co K[\bar {\alpha }]6.925E−031.39E+041.57E+041.77E+041.98E+042.22E+042.47E+042.73E+043.02E+04
Mn K[\beta _1]6.490E−031.66E+041.88E+042.11E+042.36E+042.64E+042.94E+043.26E+043.60E+04
Fe K[\bar {\alpha }]6.400E−031.73E+041.95E+042.19E+042.45E+042.74E+043.05E+043.38E+043.74E+04
Cr K[\beta _1]5.947E−032.11E+042.38E+042.67E+042.99E+043.34E+043.72E+044.12E+044.55E+04
Mn K[\bar {\alpha }]5.895E−032.16E+042.44E+042.74E+043.07E+043.42E+043.81E+044.22E+044.66E+04
Cr K[\bar {\alpha }]5.412E−032.72E+043.07E+043.45E+043.85E+044.30E+044.78E+045.29E+045.84E+04
Ti K[\beta _1]4.932E−033.50E+043.94E+044.42E+044.94E+045.51E+046.12E+046.77E+047.47E+04
Ti K[\bar {\alpha }]4.509E−034.44E+045.00E+045.61E+046.27E+046.98E+047.75E+048.56E+049.43E+04

RadiationEnergy (MeV)4142434445464748
NiobiumMolybdenumTechnetiumRutheniumRhodiumPalladiumSilverCadmium
Ag K[\beta _1]2.494E−027.41E+039.36E+038.65E+039.33E+031.00E+041.00E+042.00E+032.18E+03
Pd K[\beta _1]2.382E−027.59E+039.61E+038.86E+039.56E+031.03E+041.88E+032.05E+032.23E+03
Rh K[\beta _1]2.272E−028.57E+039.30E+039.95E+031.07E+041.18E+032.10E+032.29E+032.49E+03
Ag K[\bar {\alpha }]2.210E−029.22E+031.15E+041.07E+041.92E+032.10E+032.30E+032.51E+032.73E+03
Pd K[\bar {\alpha }]2.112E−021.04E+041.23E+041.20E+042.17E+032.38E+032.60E+032.84E+033.09E+03
Rh K[\bar {\alpha }]2.017E−021.16E+041.27E+042.24E+032.46E+032.70E+032.94E+033.21E+033.50E+03
Mo K[\beta _1]1.961E−021.25E+042.19E+032.42E+032.65E+032.91E+033.18E+033.47E+033.78E+03
Mo K[\bar {\alpha }]1.744E−022.73E+033.00E+033.32E+033.64E+033.99E+034.36E+034.76E+035.18E+03
Zn K[\beta _1]9.572E−031.40E+041.54E+041.69E+041.85E+042.02E+042.21E+042.40E+042.61E+04
Cu K[\beta _1]8.905E−031.70E+041.87E+042.05E+042.25E+042.45E+042.67E+042.91E+043.16E+04
Zn K[\bar {\alpha }]8.631E−031.85E+042.03E+042.23E+042.44E+042.67E+042.91E+043.16E+043.44E+04
Ni K[\beta _1]8.265E−032.07E+042.28E+042.51E+042.74E+043.00E+043.27E+043.55E+043.86E+04
Cu K[\bar {\alpha }]8.041E−032.23E+042.46E+042.70E+042.95E+043.23E+043.52E+043.82E+044.15E+04
Co K[\beta _1]7.649E−032.55E+042.81E+043.08E+043.37E+043.68E+044.01E+044.36E+044.73E+04
Ni K[\bar {\alpha }]7.472E−032.72E+042.99E+043.28E+043.59E+043.92E+044.27E+044.64E+045.03E+04
Fe K[\beta _1]7.058E−033.17E+043.48E+043.82E+044.18E+044.56E+044.96E+045.39E+045.84E+04
Co K[\bar {\alpha }]6.925E−033.33E+043.66E+044.02E+044.39E+044.79E+045.21E+045.66E+046.14E+04
Mn K[\beta _1]6.490E−033.96E+044.36E+044.77E+045.22E+045.69E+046.19E+046.72E+047.28E+04
Fe K[\bar {\alpha }]6.400E−034.12E+044.52E+044.95E+045.42E+045.91E+046.42E+046.97E+047.55E+04
Cr K[\beta _1]5.947E−035.01E+045.50E+046.02E+046.58E+047.17E+047.79E+048.45E+049.15E+04
Mn K[\bar {\alpha }]5.895E−035.13E+045.63E+046.16E+046.73E+047.34E+047.97E+048.64E+049.36E+04
Cr K[\bar {\alpha }]5.412E−036.42E+047.05E+047.71E+048.41E+049.16E+049.94E+041.08E+051.17E+05
Ti K[\beta _1]4.932E−038.21E+048.99E+049.83E+041.07E+051.17E+051.27E+051.37E+051.48E+05
Ti K[\bar {\alpha }]4.509E−031.04E+051.13E+051.24E+051.35E+051.47E+051.59E+051.72E+051.86E+05

RadiationEnergy (MeV)4950515253545556
IndiumTinAntimonyTelluriumIodineXenonCaesiumBarium
Ag K[\beta _1]2.494E−022.37E+032.57E+032.79E+033.02E+033.26E+033.52E+033.79E+034.08E+03
Pd K[\beta _1]2.382E−022.43E+032.64E+032.86E+033.09E+033.34E+033.61E+033.89E+034.18E+03
Rh K[\beta _1]2.272E−022.64E+033.00E+033.20E+033.50E+033.71E+034.04E+034.04E+034.76E+03
Ag K[\bar {\alpha }]2.210E−022.97E+033.23E+033.50E+033.78E+034.09E+034.41E+034.75E+035.11E+03
Pd K[\bar {\alpha }]2.112E−023.36E+033.65E+033.95E+034.28E+034.62E+034.98E+035.37E+035.78E+03
Rh K[\bar {\alpha }]2.017E−023.81E+034.13E+034.48E+034.85E+035.24E+035.65E+036.09E+036.55E+03
Mo K[\beta _1]1.961E−024.11E+034.46E+034.84E+035.23E+035.65E+036.09E+036.57E+037.06E+03
Mo K[\bar {\alpha }]1.744E−025.63E+036.11E+036.62E+037.16E+037.73E+038.34E+038.98E+039.65E+03
Zn K[\beta _1]9.572E−032.83E+043.06E+043.31E+043.57E+043.84E+044.13E+044.44E+044.76E+04
Cu K[\beta _1]8.905E−033.43E+043.71E+044.00E+044.32E+044.64E+044.99E+045.36E+045.74E+04
Zn K[\bar {\alpha }]8.631E−033.72E+044.03E+044.35E+044.69E+045.04E+045.42E+045.81E+046.23E+04
Ni K[\beta _1]8.265E−034.18E+044.52E+044.88E+045.25E+045.65E+046.07E+046.51E+046.98E+04
Cu K[\bar {\alpha }]8.041E−034.50E+044.86E+045.25E+045.65E+046.07E+046.52E+047.00E+047.50E+04
Co K[\beta _1]7.649E−035.12E+045.54E+045.98E+046.43E+046.92E+047.42E+047.96E+048.52E+04
Ni K[\bar {\alpha }]7.472E−035.45E+045.89E+046.35E+046.84E+047.35E+047.88E+048.45E+049.04E+04
Fe K[\beta _1]7.058E−036.32E+046.83E+047.37E+047.94E+048.52E+049.14E+049.77E+041.04E+05
Co K[\bar {\alpha }]6.925E−036.64E+047.18E+047.74E+048.34E+048.96E+049.60E+041.03E+051.09E+05
Mn K[\beta _1]6.490E−037.87E+048.50E+049.17E+049.88E+041.06E+051.13E+051.21E+051.29E+05
Fe K[\bar {\alpha }]6.400E−038.17E+048.82E+049.51E+041.02E+051.10E+051.18E+051.26E+051.34E+05
Cr K[\beta _1]5.947E−039.90E+041.07E+051.15E+051.24E+051.33E+051.42E+051.60E+051.47E+05
Mn K[\bar {\alpha }]5.895E−031.01E+051.09E+051.18E+051.27E+051.36E+051.45E+051.63E+051.50E+05
Cr K[\bar {\alpha }]5.412E−031.26E+051.36E+051.46E+051.57E+051.68E+051.57E+051.77E+051.34E+05
Ti K[\beta _1]4.932E−031.60E+051.73E+051.85E+051.98E+052.11E+052.48E+055.70E+047.16E+04
Ti K[\bar {\alpha }]4.509E−032.00E+052.15E+052.00E+051.60E+056.17E+045.78E+047.28E+047.62E+04

RadiationEnergy (MeV)5758596061626364
LanthanumCeriumPraseodymiumNeodymiumPromethiumSamariumEuropiumGadolinium
Ag K[\beta _1]2.494E−023.97E+034.26E+034.56E+034.89E+035.22E+035.57E+035.93E+036.32E+03
Pd K[\beta _1]2.382E−024.05E+034.82E+035.15E+035.51E+035.90E+036.29E+036.71E+037.15E+03
Rh K[\beta _1]2.272E−025.10E+035.47E+035.85E+036.25E+036.69E+037.14E+037.59E+038.09E+03
Ag K[\bar {\alpha }]2.210E−025.49E+035.89E+036.29E+036.73E+037.20E+037.69E+038.17E+038.72E+03
Pd K[\bar {\alpha }]2.112E−026.20E+036.63E+037.11E+037.62E+038.14E+038.69E+039.23E+039.84E+03
Rh K[\bar {\alpha }]2.017E−027.03E+037.69E+038.07E+038.62E+039.22E+039.84E+031.05E+041.11E+04
Mo K[\beta _1]1.961E−027.59E+038.12E+038.70E+039.29E+039.94E+031.06E+041.13E+041.20E+04
Mo K[\bar {\alpha }]1.744E−021.04E+041.11E+041.19E+041.27E+041.36E+041.44E+041.54E+041.63E+04
Zn K[\beta _1]9.572E−035.10E+045.42E+045.78E+046.15E+046.57E+046.97E+049.93E+047.83E+04
Cu K[\beta _1]8.905E−036.14E+046.56E+047.00E+047.42E+047.90E+048.36E+048.88E+049.40E+04
Zn K[\bar {\alpha }]8.631E−036.67E+047.12E+047.58E+048.04E+048.44E+049.06E+049.59E+041.02E+05
Ni K[\beta _1]8.265E−037.47E+047.96E+048.49E+049.00E+049.56E+041.01E+051.07E+059.84E+04
Cu K[\bar {\alpha }]8.041E−038.03E+048.56E+049.12E+049.68E+041.03E+051.08E+051.02E+051.05E+05
Co K[\beta _1]7.649E−039.11E+049.70E+041.03E+051.09E+051.16E+051.07E+051.21E+058.75E+04
Ni K[\bar {\alpha }]7.472E−039.68E+041.03E+051.09E+051.16E+051.23E+051.14E+058.70E+049.29E+04
Fe K[\beta _1]7.058E−031.11E+051.19E+051.26E+051.18E+051.22E+059.56E+041.03E+053.99E+04
Co K[\bar {\alpha }]6.925E−031.17E+051.24E+051.32E+051.21E+059.63E+049.89E+044.04E+044.20E+04
Mn K[\beta _1]6.490E−031.38E+051.27E+051.39E+051.05E+051.13E+054.39E+044.67E+044.93E+04
Fe K[\bar {\alpha }]6.400E−031.43E+051.31E+051.02E+051.09E+054.19E+044.59E+044.87E+045.09E+04
Cr K[\beta _1]5.947E−031.48E+051.15E+051.22E+054.74E+045.01E+045.52E+045.65E+046.14E+04
Mn K[\bar {\alpha }]5.895E−031.50E+051.19E+054.52E+044.86E+045.18E+045.64E+046.03E+046.29E+04
Cr K[\bar {\alpha }]5.412E−035.19E+045.54E+045.57E+046.01E+046.43E+046.97E+047.54E+047.78E+04
Ti K[\beta _1]4.932E−036.55E+046.98E+047.02E+047.52E+048.11E+048.74E+049.34E+049.76E+04
Ti K[\bar {\alpha }]4.509E−038.19E+048.31E+048.77E+049.51E+041.02E+051.09E+051.16E+051.22E+05

RadiationEnergy (MeV)6566676869707172
TerbiumDysprosiumHolmiumErbiumThuliumYtterbiumLutetiumHafnium
Ag K[\beta _1]2.494E−026.66E+037.15E+037.59E+038.05E+038.52E+039.02E+039.53E+031.01E+04
Pd K[\beta _1]2.382E−027.52E+038.07E+038.57E+039.08E+039.62E+031.02E+041.08E+041.14E+04
Rh K[\beta _1]2.272E−028.15E+039.15E+039.69E+031.03E+041.08E+041.15E+041.22E+041.29E+04
Ag K[\bar {\alpha }]2.210E−029.16E+039.85E+031.01E+041.11E+041.17E+041.24E+041.31E+041.38E+04
Pd K[\bar {\alpha }]2.112E−021.03E+041.11E+041.18E+041.25E+041.32E+041.40E+041.48E+041.56E+04
Rh K[\bar {\alpha }]2.017E−021.17E+041.26E+041.33E+041.41E+041.50E+041.58E+041.67E+041.76E+04
Mo K[\beta _1]1.961E−021.26E+041.35E+041.43E+041.52E+041.61E+041.70E+041.80E+041.90E+04
Mo K[\bar {\alpha }]1.744E−021.72E+041.84E+041.95E+042.07E+042.19E+042.31E+042.44E+042.57E+04
Zn K[\beta _1]9.572E−038.20E+048.74E+049.20E+048.53E+046.59E+046.90E+047.26E+047.70E+04
Cu K[\beta _1]8.905E−039.87E+049.31E+047.17E+047.42E+048.27E+043.10E+043.52E+043.56E+04
Zn K[\bar {\alpha }]8.631E−039.29E+041.02E+057.78E+047.97E+043.28E+043.36E+043.81E+043.85E+04
Ni K[\beta _1]8.265E−031.08E+058.53E+048.60E+043.42E+043.67E+043.76E+044.24E+044.30E+04
Cu K[\bar {\alpha }]8.041E−038.38E+049.23E+043.53E+043.67E+043.93E+044.08E+044.53E+044.59E+04
Co K[\beta _1]7.649E−039.40E+043.72E+044.00E+044.14E+044.46E+044.57E+045.17E+045.21E+04
Ni K[\bar {\alpha }]7.472E−033.89E+043.94E+044.24E+044.39E+044.47E+044.86E+045.49E+045.54E+04
Fe K[\beta _1]7.058E−034.22E+044.53E+044.87E+045.05E+045.50E+045.63E+046.33E+046.40E+04
Co K[\bar {\alpha }]6.925E−034.43E+044.75E+045.12E+045.30E+045.78E+045.92E+046.65E+046.73E+04
Mn K[\beta _1]6.490E−035.22E+045.59E+046.02E+046.22E+046.82E+047.01E+047.84E+047.91E+04
Fe K[\bar {\alpha }]6.400E−035.44E+045.77E+046.24E+046.44E+047.10E+047.21E+048.13E+048.21E+04
Cr K[\beta _1]5.947E−036.46E+046.93E+047.45E+047.75E+048.55E+048.73E+049.85E+049.90E+04
Mn K[\bar {\alpha }]5.895E−036.68E+047.07E+047.67E+047.92E+048.75E+048.94E+041.01E+051.01E+05
Cr K[\bar {\alpha }]5.412E−038.29E+048.77E+049.51E+049.78E+041.08E+051.11E+051.25E+051.26E+05
Ti K[\beta _1]4.932E−031.05E+051.11E+051.20E+051.23E+051.39E+051.41E+051.59E+051.60E+05
Ti K[\bar {\alpha }]4.509E−031.31E+051.39E+051.50E+051.54E+051.73E+051.78E+052.00E+052.01E+05

RadiationEnergy (MeV)7374757677787980
TantalumTungstenRheniumOsmiumIridiumPlatinumGoldMercury
Ag K[\beta _1]2.494E−021.17E+041.24E+041.30E+041.37E+041.44E+041.52E+041.60E+041.68E+04
Pd K[\beta _1]2.382E−021.20E+041.27E+041.34E+041.41E+041.48E+041.56E+041.64E+041.72E+04
Rh K[\beta _1]2.272E−021.35E+041.45E+041.52E+041.59E+041.69E+041.78E+041.77E+041.94E+04
Ag K[\bar {\alpha }]2.210E−021.46E+041.54E+041.62E+041.71E+041.80E+041.89E+041.99E+042.09E+04
Pd K[\bar {\alpha }]2.112E−021.65E+041.74E+041.83E+041.93E+042.02E+042.13E+042.24E+042.35E+04
Rh K[\bar {\alpha }]2.017E−021.86E+041.96E+042.07E+042.17E+042.28E+042.40E+042.52E+042.65E+04
Mo K[\beta _1]1.961E−022.00E+042.11E+042.22E+042.34E+042.46E+042.58E+042.71E+042.84E+04
Mo K[\bar {\alpha }]1.744E−022.72E+042.86E+043.01E+043.16E+043.31E+043.48E+043.65E+043.82E+04
Zn K[\beta _1]9.572E−033.30E+043.29E+043.67E+043.73E+044.01E+044.08E+044.24E+044.29E+04
Cu K[\beta _1]8.905E−033.76E+043.97E+044.43E+044.48E+044.83E+044.89E+045.08E+045.18E+04
Zn K[\bar {\alpha }]8.631E−034.05E+044.26E+044.78E+044.84E+045.20E+045.28E+045.58E+045.58E+04
Ni K[\beta _1]8.265E−034.41E+044.80E+045.34E+045.41E+045.81E+045.66E+046.14E+046.20E+04
Cu K[\bar {\alpha }]8.041E−034.85E+045.13E+045.72E+045.80E+046.24E+046.34E+046.69E+046.68E+04
Co K[\beta _1]7.649E−035.39E+045.83E+046.50E+046.55E+047.09E+046.87E+047.49E+046.91E+04
Ni K[\bar {\alpha }]7.472E−035.84E+046.19E+046.88E+046.97E+047.51E+047.62E+048.03E+048.02E+04
Fe K[\beta _1]7.058E−036.60E+047.15E+047.95E+048.03E+048.68E+048.45E+049.17E+049.14E+04
Co K[\bar {\alpha }]6.925E−036.92E+047.51E+048.33E+048.44E+049.11E+049.20E+049.71E+049.71E+04
Mn K[\beta _1]6.490E−038.16E+048.82E+049.83E+049.94E+041.07E+051.04E+051.14E+051.09E+05
Fe K[\bar {\alpha }]6.400E−038.46E+049.15E+041.02E+051.02E+051.11E+051.12E+051.18E+051.19E+05
Cr K[\beta _1]5.947E−031.01E+051.10E+051.22E+051.23E+051.34E+051.31E+051.41E+051.36E+05
Mn K[\bar {\alpha }]5.895E−031.06E+051.12E+051.25E+051.27E+051.37E+051.34E+051.45E+051.42E+05
Cr K[\bar {\alpha }]5.412E−031.31E+051.40E+051.55E+051.57E+051.70E+051.66E+051.79E+051.80E+05
Ti K[\beta _1]4.932E−031.64E+051.77E+051.96E+051.99E+052.15E+052.11E+052.23E+052.33E+05
Ti K[\bar {\alpha }]4.509E−032.60E+052.18E+052.46E+052.49E+052.65E+052.66E+052.80E+052.98E+05

RadiationEnergy (MeV)8182838485868788
ThalliumLeadBismuthPoloniumAstatineRadonFranciumRadium
Ag K[\beta _1]2.494E−021.76E+041.84E+041.93E+042.02E+042.11E+042.20E+042.30E+042.41E+04
Pd K[\beta _1]2.382E−021.80E+041.89E+041.98E+042.07E+042.16E+042.26E+042.36E+042.46E+04
Rh K[\beta _1]2.272E−021.99E+042.05E+042.27E+042.37E+042.49E+042.55E+042.60E+042.68E+04
Ag K[\bar {\alpha }]2.210E−022.19E+042.29E+042.40E+042.51E+042.62E+042.73E+042.85E+042.98E+04
Pd K[\bar {\alpha }]2.112E−022.46E+042.58E+042.70E+042.82E+042.94E+043.07E+043.20E+043.34E+04
Rh K[\bar {\alpha }]2.017E−022.78E+042.91E+043.04E+043.18E+043.31E+043.45E+043.60E+043.76E+04
Mo K[\beta _1]1.961E−022.98E+043.12E+043.27E+043.41E+043.56E+043.71E+043.87E+044.03E+04
Mo K[\bar {\alpha }]1.744E−024.01E+044.19E+044.38E+044.58E+044.07E+043.98E+043.22E+043.30E+04
Zn K[\beta _1]9.572E−034.83E+045.11E+045.42E+045.66E+045.56E+046.22E+046.55E+046.56E+04
Cu K[\beta _1]8.905E−035.81E+046.15E+046.51E+046.80E+046.68E+047.48E+047.87E+047.80E+04
Zn K[\bar {\alpha }]8.631E−036.29E+046.66E+047.04E+047.35E+047.21E+048.11E+048.92E+048.55E+04
Ni K[\beta _1]8.265E−037.03E+047.44E+047.86E+048.22E+048.04E+049.06E+049.94E+049.53E+04
Cu K[\bar {\alpha }]8.041E−037.54E+047.98E+048.43E+048.81E+048.65E+049.72E+041.02E+051.02E+05
Co K[\beta _1]7.649E−038.58E+049.01E+049.57E+041.00E+059.82E+041.11E+051.16E+051.16E+05
Ni K[\bar {\alpha }]7.472E−039.11E+049.64E+041.02E+051.06E+051.04E+051.17E+051.23E+051.23E+05
Fe K[\beta _1]7.058E−031.05E+051.12E+051.17E+051.22E+051.20E+051.36E+051.42E+051.43E+05
Co K[\bar {\alpha }]6.925E−031.11E+051.17E+051.23E+051.29E+051.26E+051.43E+051.49E+051.49E+05
Mn K[\beta _1]6.490E−031.31E+051.39E+051.46E+051.51E+051.49E+051.69E+051.77E+051.76E+05
Fe K[\bar {\alpha }]6.400E−031.36E+051.48E+051.51E+051.57E+051.54E+051.75E+051.82E+051.83E+05
Cr K[\beta _1]5.947E−031.64E+051.74E+051.82E+051.89E+051.86E+052.11E+052.21E+052.19E+05
Mn K[\bar {\alpha }]5.895E−031.68E+051.78E+051.86E+051.93E+051.91E+052.16E+052.23E+052.25E+05
Cr K[\bar {\alpha }]5.412E−032.22E+052.22E+052.32E+051.99E+052.37E+052.70E+052.82E+052.79E+05
Ti K[\beta _1]4.932E−032.66E+052.82E+052.94E+052.88w+052.99E+053.43E+053.56E+053.53E+05
Ti K[\bar {\alpha }]4.509E−033.36E+053.56E+053.67E+053.17E+053.78E+054.33E+054.49E+054.99E+05

RadiationEnergy (MeV)8990919293949596
ActiniumThoriumProtactiniumUraniumNeptuniumPlutoniumAmericiumCurium
Ag K[\beta _1]2.494E−022.51E+042.62E+042.73E+042.84E+042.95E+043.07E+043.18E+042.89E+04
Pd K[\beta _1]2.382E−022.57E+042.68E+042.79E+042.90E+043.02E+043.14E+043.26E+042.96E+04
Rh K[\beta _1]2.272E−022.83E+043.09E+043.00E+043.54E+043.42E+043.03E+043.34E+042.36E+04
Ag K[\bar {\alpha }]2.210E−023.11E+043.23E+043.42E+043.50E+042.99E+042.27E+042.50E+042.46E+04
Pd K[\bar {\alpha }]2.112E−023.48E+043.62E+043.77E+043.40E+044.08E+044.24E+044.39E+044.05E+04
Rh K[\bar {\alpha }]2.017E−023.92E+044.07E+044.24E+042.74E+044.58E+044.76E+044.93E+044.57E+04
Mo K[\beta _1]1.961E−023.42E+043.80E+044.55E+042.96E+044.91E+045.10E+045.29E+044.92E+04
Mo K[\bar {\alpha }]1.744E−025.40E+043.70E+043.87E+044.03E+042.57E+041.62E+041.89E+042.01E+04
Zn K[\beta _1]9.572E−031.07E+056.57E+046.74E+047.11E+047.47E+047.29E+047.63E+047.95E+04
Cu K[\beta _1]8.905E−031.14E+058.70E+048.11E+048.54E+048.92E+048.75E+049.27E+049.51E+04
Zn K[\bar {\alpha }]8.631E−031.16E+059.84E+048.78E+049.23E+049.67E+049.48E+041.01E+051.03E+05
Ni K[\beta _1]8.265E−031.19E+051.10E+059.82E+041.03E+051.08E+051.06E+051.14E+051.15E+05
Cu K[\bar {\alpha }]8.041E−031.43E+051.18E+051.06E+051.12E+051.23E+051.13E+051.44E+051.38E+05
Co K[\beta _1]7.649E−031.50E+051.34E+051.19E+051.26E+051.32E+051.28E+051.46E+051.41E+05
Ni K[\bar {\alpha }]7.472E−031.67E+051.42E+051.27E+051.33E+051.40E+051.36E+051.48E+051.47E+05
Fe K[\beta _1]7.058E−031.47E+051.54E+051.46E+051.54E+051.61E+051.57E+051.73E+051.73E+05
Co K[\bar {\alpha }]6.925E−031.74E+051.72E+051.53E+051.61E+051.69E+051.65E+051.81E+051.79E+05
Mn K[\beta _1]6.490E−031.96E+051.87E+051.81E+051.90E+051.98E+051.94E+052.15E+052.11E+05
Fe K[\bar {\alpha }]6.400E−032.00E+051.96E+051.88E+051.97E+052.17E+052.02E+052.43E+052.42E+05
Cr K[\beta _1]5.947E−032.32E+052.35E+052.27E+052.37E+052.48E+052.43E+052.72E+052.62E+05
Mn K[\bar {\alpha }]5.895E−032.37E+052.40E+052.31E+052.43E+052.53E+052.48E+052.78E+052.68E+05
Cr K[\bar {\alpha }]5.412E−032.79E+052.96E+052.88E+053.03E+053.14E+053.08E+053.49E+053.33E+05
Ti K[\beta _1]4.932E−033.32E+053.77E+053.83E+053.82E+053.97E+053.90E+054.47E+054.22E+05
Ti K[\bar {\alpha }]4.509E−033.95E+054.76E+054.84E+054.88E+053.79E+053.65E+054.26E+054.03E+05

RadiationEnergy (MeV)9798
BerkeliumCalifornium
Ag K[\beta _1]2.494E−022.13E+043.06E+04
Pd K[\beta _1]2.382E−022.18E+043.44E+04
Rh K[\beta _1]2.272E−022.41E+043.86E+04
Ag K[\bar {\alpha }]2.210E−022.50E+042.89E+04
Pd K[\bar {\alpha }]2.112E−022.98E+044.62E+04
Rh K[\bar {\alpha }]2.017E−023.37E+045.21E+04
Mo K[\beta_1]1.961E−023.64E+045.59E+04
Mo K[\bar {\alpha }]1.744E−022.01E+042.09E+04
Zn K[\beta_1]9.572E−037.63E+048.67E+04
Cu K[\beta _1]8.905E−039.27E+041.04E+04
Zn K[\bar {\alpha }]8.631E−031.01E+051.13E+05
Ni K[\beta _1]8.265E−031.13E+051.26E+05
Cu K[\bar {\alpha }]8.041E−031.43E+051.50E+05
Co K[\beta _1]7.649E−031.46E+051.52E+05
Ni K[\bar {\alpha }]7.472E−031.48E+051.61E+05
Fe K[\beta _1]7.058E−031.73E+051.87E+05
Co K[\bar {\alpha }]6.925E−031.82E+051.96E+05
Mn K[\beta _1]6.490E−032.16E+052.30E+05
Fe K[\bar {\alpha }]6.400E−032.43E+052.53E+05
Cr K[\beta _1]5.947E−032.72E+052.86E+05
Mn K[\bar {\alpha }]5.895E−032.78E+052.93E+05
Cr K[\bar {\alpha }]5.412E−033.49E+053.63E+05
Ti K[\beta _1]4.932E−034.47E+054.59E+05
Ti K[\bar {\alpha }]4.509E−034.26E+054.38E+05

Table 4.2.4.3| top | pdf |
Mass attenuation coefficients (cm2 g−1)

RadiationEnergy (MeV)12345678
HydrogenHeliumLithiumBerylliumBoronCarbonNitrogenOxygen
Ag K[\beta _1]2.494E−023.63E−011.89E−011.72E−011.95E−012.37E−013.15E−014.04E−013.29E−01
Pd K[\beta _1]2.382E−023.65E−011.90E−011.74E−012.00E−012.47E−013.35E−014.37E−015.82E−01
Rh K[\beta _1]2.272E−023.66E−011.92E−011.77E−012.05E−012.59E−013.58E−014.77E−016.44E−01
Ag K[\bar {\alpha }]2.210E−023.67E−011.93E−011.79E−012.09E−012.67E−013.74E−015.03E−016.85E−01
Pd K[\bar {\alpha }]2.112E−023.68E−011.94E−011.82E−012.16E−012.81E−014.02E−015.51E−017.60E−01
Rh K[\bar {\alpha }]2.017E−023.69E−011.96E−011.85E−012.24E−012.98E−014.35E−016.07E−018.48E−01
Mo K[\beta _1]1.961E−023.70E−011.97E−011.87E−012.29E−013.09E−014.58E−016.45E−019.08E−01
Mo K[\bar {\alpha }]1.744E−023.73E−012.02E−011.98E−012.56E−013.68E−015.76E−018.45E−011.22E+00
Zn K[\beta _1]9.572E−033.86E−012.55E−013.64E−017.16E−011.41E+002.69E+004.42E+006.78E+00
Cu K[\beta _1]8.905E−033.88E−012.68E−014.12E−018.53E−011.73E+003.33E+005.48E+008.42E+00
Zn K[\bar {\alpha }]8.631E−033.89E−012.74E−014.36E−019.23E−011.89E+003.65E+006.01E+009.25E+00
Ni K[\beta _1]8.265E−033.90E−012.85E−014.73E−011.03E+002.14E+004.15E+006.85E+001.05E+01
Cu K[\bar {\alpha }]8.041E−033.91E−012.92E−015.00E−011.11E+002.31E+004.51E+007.44E+001.15E+01
Co K[\beta _1]7.649E−033.93E−013.07E−015.55E−011.27E+002.67E+005.24E+008.66E+001.33E+01
Ni K[\bar {\alpha }]7.472E−033.94E−013.14E−015.84E−011.35E+002.87E+005.62E+009.29E+001.43E+01
Fe K[\beta _1]7.058E−033.96E−013.35E−016.63E−011.58E+003.39E+006.68E+001.10E+011.70E+01
Co K[\bar {\alpha }]6.925E−033.97E−013.43E−016.93E−011.67E+003.59E+007.07E+001.17E+011.80E+01
Mn K[\beta _1]6.490E−034.00E−013.74E−018.10E−012.01E+004.37E+008.62E+001.42E+012.19E+01
Fe K[\bar {\alpha }]6.400E−034.00E−013.81E−018.39E−012.09E+004.55E+008.99E+001.49E+012.28E+01
Cr K[\beta _1]5.947E−034.05E−014.25E−011.01E+002.59E+005.69E+001.12E+011.86E+012.84E+01
Mn K[\bar {\alpha }]5.895E−034.05E−014.31E−011.03E+002.66E+005.84E+001.16E+011.91E+012.92E+01
Cr K[\bar {\alpha }]5.412E−034.12E−014.98E−011.30E+003.44E+007.59E+001.50E+012.47E+013.78E+01
Ti K[\beta _1]4.932E−034.21E−015.92E−011.68E+004.56E+001.01E+011.99E+013.28E+014.99E+01
Ti K[\bar {\alpha }]4.509E−034.33E−017.12E−012.18E+006.00E+001.33E+012.62E+014.30E+016.52E+01

RadiationEnergy (MeV)910111213141516
FluorineNeonSodiumMagnesiumAluminiumSiliconPhosphorusSulfur
Ag K[\beta _1]2.494E−026.60E−019.06E−011.13E+001.50E+001.85E+002.38E+002.84E+003.55E+00
Pd K[\beta _1]2.382E−027.35E−011.02E+001.28E+001.70E+002.10E+002.71E+003.24E+004.05E+00
Rh K[\beta _1]2.272E−028.22E−011.15E+001.45E+001.93E+002.39E+003.09E+003.70E+004.64E+00
Ag K[\bar {\alpha }]2.210E−028.79E−011.23E+001.56E+002.09E+002.59E+003.35E+004.01E+005.02E+00
Pd K[\bar {\alpha }]2.112E−029.84E−011.39E+001.77E+002.37E+002.94E+003.81E+004.57E+005.72E+00
Rh K[\bar {\alpha }]2.017E−021.11E+001.57E+002.01E+002.70E+003.36E+004.36E+005.23E+006.55E+00
Mo K[\beta _1]1.961E−021.19E+001.69E+002.17E+002.92E+003.64E+004.73E+005.67E+007.11E+00
Mo K[\bar {\alpha }]1.744E−021.63E+002.35E+003.03E+004.09E+005.11E+006.64E+007.97E+009.99E+00
Zn K[\beta _1]9.572E−039.35E+001.36E+011.77E+012.40E+012.98E+013.85E+014.58E+015.68E+01
Cu K[\beta _1]8.905E−031.16E+011.69E+012.20E+012.96E+013.68E+014.75E+015.64E+016.98E+01
Zn K[\bar {\alpha }]8.631E−031.28E+011.86E+012.41E+013.25E+014.03E+015.20E+016.17E+017.63E+01
Ni K[\beta_1]8.265E−031.45E+012.11E+012.74E+013.69E+014.58E+015.89E+016.98E+018.63E+01
Cu K[\bar {\alpha }]8.041E−031.58E+012.29E+012.97E+014.00E+014.96E+016.37E+017.55E+019.33E+01
Co K[\beta _1]7.649E−031.83E+012.66E+013.45E+014.63E+015.73E+017.36E+018.70E+011.07E+02
Ni K[\bar {\alpha }]7.472E−031.97E+012.85E+013.69E+014.96E+016.13E+017.87E+019.30E+011.15E+02
Fe K[\beta _1]7.058E−032.33E+013.38E+014.37E+015.85E+017.23E+019.27E+011.09E+021.35E+02
Co K[\bar {\alpha }]6.925E−032.47E+013.58E+014.62E+016.19E+017.64E+019.78E+011.15E+021.42E+02
Mn K[\beta _1]6.490E 033.00E+014.34E+015.59E+017.47E+019.21E+011.18E+021.39E+021.70E+02
Fe K[\bar {\alpha }]6.400E−033.13E+014.52E+015.82E+017.78E+019.59E+011.22E+021.44E+021.77E+02
Cr K[\beta _1]5.947E−033.89E+015.61E+017.21E+019.62E+011.18E+021.51E+021.77E+022.17E+02
Mn K[\bar {\alpha }]5.895E−033.99E+015.76E+017.40E+019.87E+011.21E+021.54E+021.81E+022.22E+02
Cr K[\bar {\alpha }]5.412E−035.15E+017.41E+019.49E+011.26E+021.55E+021.96E+022.30E+022.81E+02
Ti K[\beta _1]4.932E−036.78E+019.72E+011.24E+021.65E+022.01E+022.55E+022.97E+023.62E+02
Ti K[\bar {\alpha }]4.509E−038.84E+011.26E+021.61E+022.12E+022.59E+023.27E+023.79E+024.60E+02

RadiationEnergy (MeV)1718192021222324
ChlorineArgonPotassiumCalciumScandiumTitaniumVanadiumChromium
Ag K[\beta _1]2.494E−024.09E+004.56E+005.78E+006.92E+007.47E+008.43E+009.42E+001.09E+01
Pd K[\beta _1]2.382E−024.67E+005.21E+006.60E+007.90E+008.53E+009.61E+001.07E+011.24E+01
Rh K[\beta _1]2.272E−025.35E+005.96E+007.56E+009.04E+009.76E+001.10E+011.23E+011.42E+01
Ag K[\bar {\alpha }]2.210E−025.79E+006.46E+008.19E+009.79E+001.06E+011.19E+011.33E+011.54E+01
Pd K[\bar {\alpha }]2.112E−026.61E+007.37E+009.33E+001.12E+011.20E+011.36E+011.51E+011.75E+01
Rh K[\bar {\alpha }]2.017E−027.55E+008.42E+001.07E+011.27E+011.38E+011.55E+011.73E+011.99E+01
Mo K[\beta _1]1.961E−028.20E+009.14E+001.16E+011.38E+011.49E+011.68E+011.87E+012.15E+01
Mo K[\bar {\alpha }]1.744E−021.15E+011.28E+011.62E+011.93E+012.08E+012.34E+012.60E+012.99E+01
Zn K[\beta _1]9.572E−036.48E+017.14E+018.94E+011.05E+021.12E+021.25E+021.37E+021.55E+02
Cu K[\beta _1]8.905E−037.95E+018.75E+011.09E+021.29E+021.37E+021.52E+021.66E+021.85E+02
Zn K[\bar {\alpha }]8.631E−038.69E+019.55E+011.19E+021.40E+021.49E+021.66E+021.81E+022.01E+02
Ni K[\beta _1]8.265E−039.81E+011.08E+021.34E+021.58E+021.67E+021.86E+022.03E+022.27E+02
Cu K[\bar {\alpha }]8.041E−031.06E+021.16E+021.45E+021.70E+021.80E+022.00E+022.19E+022.47E+02
Co K[\beta _1]7.649E−031.22E+021.34E+021.66E+021.95E+022.06E+022.27E+022.50E+022.93E+02
Ni K[\bar {\alpha }]7.472E−031.30E+021.43E+021.77E+022.08E+022.20E+022.40E+022.66E+023.18E+02
Fe K[\beta _1]7.058E−031.52E+021.67E+022.07E+022.42E+022.56E+022.77E+023.09E+023.85E+02
Co K[\bar {\alpha }]6.925E−031.61E+021.76E+022.18E+022.55E+022.69E+022.91E+023.25E+024.08E+02
Mn K[\beta _1]6.490E−031.92E+022.10E+022.60E+023.03E+023.19E+023.45E+023.85E+024.80E+02
Fe K[\bar {\alpha }]6.400E−032.00E+022.18E+022.70E+023.14E+023.32E+023.58E+023.99E+024.92E+02
Cr K[\beta _1]5.947E−032.44E+022.66E+023.28E+023.82E+024.03E+024.44E+024.79E+026.70E+01
Mn K[\bar {\alpha }]5.895E−032.50E+022.72E+023.36E+023.91E+024.12E+024.57E+024.89E+026.86E+01
Cr K[\bar {\alpha }]5.412E−033.16E+023.42E+024.21E+024.90E+025.16E+025.90E+027.47E+018.68E+01
Ti K[\beta _1]4.932E−034.04E+024.38E+025.38E+026.24E+026.52E+028.54E+019.65E+011.12E+02
Ti K[\bar {\alpha }]4.509E−035.11E+025.56E+026.80E+027.81E+028.08E+021.09E+021.23E+021.43E+02

RadiationEnergy (MeV)2526272829303132
ManganeseIronCobaltNickelCopperZincGalliumGermanium
Ag K[\beta _1]2.494E−021.21E+011.38E+011.51E+011.74E+011.83E+012.02E+012.14E+012.31E+01
Pd K[\beta _1]2.382E−021.37E+011.57E+011.72E+011.98E+012.08E+012.30E+012.43E+012.62E+01
Rh K[\beta _1]2.272E−021.57E+011.79E+011.96E+012.26E+012.38E+012.62E+012.77E+012.98E+01
Ag K[\bar {\alpha }]2.210E−021.70E+011.94E+012.12E+012.44E+012.56E+012.82E+012.98E+013.21E+01
Pd K[\bar {\alpha }]2.112E−021.93E+012.20E+012.41E+012.77E+012.91E+013.20E+013.38E+013.64E+01
Rh K[\bar {\alpha }]2.017E−022.20E+012.51E+012.74E+013.15E+013.30E+013.63E+013.84E+014.13E+01
Mo K[\beta _1]1.961E−022.38E+012.71E+012.96E+013.40E+013.57E+013.93E+014.15E+014.46E+01
Mo K[\bar {\alpha }]1.744E−023.31E+013.76E+014.10E+014.69E+014.91E+015.40E+015.70E+016.12E+01
Zn K[\beta _1]9.572E−031.70E+021.92E+022.06E+022.33E+022.40E+023.59E+013.85E+014.22E+01
Cu K[\beta _1]8.905E−032.07E+022.32E+022.48E+022.79E+023.92E+014.38E+014.70E+015.14E+01
Zn K[\bar {\alpha }]8.631E 032.24E+022.52E+022.69E+023.02E+024.27E+014.77E+015.12E+015.59E+01
Ni K[\beta _1]8.265E−032.51E+022.81E+023.00E+024.53E+014.80E+015.37E+015.76E+016.30E+01
Cu K[\bar {\alpha }]8.041E 032.70E+023.02E+023.21E+024.88E+015.18E+015.79E+016.21E+016.79E+01
Co K[\beta _1]7.649E−033.06E+023.42E+024.81E+015.60E+015.94E+016.64E+017.12E+017.78E+01
Ni K[\bar {\alpha }]7.472E−033.25E+023.62E+025.13E+015.97E+016.33E+017.08E+017.59E+018.29E+01
Fe K[\beta _1]7.058E−033.75E+025.43E+016.00E+016.98E+017.40E+018.27E+018.86E+019.69E+01
Co K[\bar {\alpha }]6.925E−033.93E+025.72E+016.32E+017.35E+017.80E+018.71E+019.34E+011.02E+02
Mn K[\beta _1]6.490E−035.92E+016.84E+017.55E+018.78E+019.31E+011.04E+021.11E+021.22E+02
Fe K[\bar {\alpha }]6.400E−036.16E+017.10E+017.85E+019.13E+019.68E+011.08E+021.16E+021.27E+02
Cr K[\beta _1]5.947E−037.53E+018.69E+019.60E+011.12E+021.18E+021.32E+021.42E+021.55E+02
Mn K[\bar {\alpha }]5.895E−037.72E+018.90E+019.83E+011.14E+021.21E+021.35E+021.45E+021.58E+02
Cr K[\bar {\alpha }]5.412E−039.75E+011.13E+021.24E+021.44E+021.53E+021.71E+021.83E+021.99E+02
Ti K[\beta _1]4.932E−031.26E+021.45E+021.60E+021.86E+021.97E+022.20E+022.35E+022.56E+02
Ti K[\bar {\alpha }]4.509E−031.61E+021.85E+022.04E+022.37E+022.51E+022.80E+022.99E+023.26E+02

RadiationEnergy (MeV)3334353637383940
ArsenicSeleniumBromineKryptonRubidiumStrontiumYttriumZirconium
Ag K[\beta _1]2.494E−022.50E+012.65E+012.91E+013.07E+013.32E+013.56E+013.84E+014.07E+01
Pd K[\beta _1]2.382E−022.84E+013.00E+013.29E+013.48E+013.76E+014.03E+014.34E+014.60E+01
Rh K[\beta _1]2.272E−023.23E+013.41E+013.74E+013.95E+014.27E+014.57E+014.91E+015.20E+01
Ag K[\bar {\alpha }]2.210E−023.48E+013.68E+014.03E+014.25E+014.59E+014.91E+015.29E+015.59E+01
Pd K[\bar {\alpha }]2.112E−023.93E+014.16E+014.55E+014.80E+015.18E+015.54E+015.95E+016.29E+01
Rh K[\bar {\alpha }]2.017E−024.46E+014.71E+015.15E+015.43E+015.85E+016.25E+016.71E+016.25E+01
Mo K[\beta _1]1.961E−024.82E+015.08E+015.55E+015.84E+016.30E+016.72E+017.21E+017.61E+01
Mo K[\bar {\alpha }]1.744E−026.61E+016.95E+017.56E+017.93E+018.51E+019.06E+019.70E+011.63E+01
Zn K[\beta _1]9.572E−034.64E+014.97E+015.53E+015.92E+016.49E+017.06E+017.73E+018.35E+01
Cu K[\beta _1]8.905E−035.65E+016.05E+016.74E+017.21E+017.90E+018.59E+019.40E+011.01E+02
Zn K[\bar {\alpha }]8.631E−036.15E+016.59E+017.33E+017.85E+018.60E+019.35E+011.02E+021.10E+02
Ni K[\beta _1]8.265E−036.93E+017.42E+018.26E+018.83E+019.68E+011.05E+021.15E+021.24E+02
Cu K[\bar {\alpha }]8.041E−037.47E+018.00E+018.90E+019.52E+011.04E+021.13E+021.24E+021.39E+02
Co K[\beta _1]7.649E−038.55E+019.16E+011.02E+021.09E+021.19E+021.30E+021.42E+021.54E+02
Ni K[\bar {\alpha }]7.472E−039.11E+019.76E+011.09E+021.16E+021.27E+021.38E+021.51E+021.63E+02
Fe K[\beta _1]7.058E−031.06E+021.14E+021.27E+021.35E+021.48E+021.61E+021.76E+021.91E+02
Co K[\bar {\alpha }]6.925E−031.12E+021.20E+021.33E+021.42E+021.56E+021.70E+021.85E+022.00E+02
Mn K[\beta _1]6.490E−031.34E+021.43E+021.59E+021.70E+021.86E+022.02E+022.21E+022.38E+02
Fe K[\bar {\alpha }]6.400E−031.39E+021.49E+021.65E+021.76E+021.93E+022.10E+022.29E+022.47E+02
Cr K[\beta _1]5.947E−031.70E+021.82E+022.02E+022.15E+022.36E+022.56E+022.79E+023.00E+02
Mn K[\bar {\alpha }]5.895E−031.74E+021.86E+022.06E+022.20E+022.41E+022.62E+022.86E+023.08E+02
Cr K[\bar {\alpha }]5.412E−032.19E+022.34E+022.60E+022.77E+023.03E+023.28E+023.58E+023.86E+02
Ti K[\beta _1]4.932E−032.81E+023.00E+023.33E+023.55E+023.88E+024.21E+024.59E+024.93E+02
Ti K[\bar {\alpha }]4.509E−033.57E+023.81E+024.23E+024.50E+024.92E+025.32E+025.80E+026.22E+02

RadiationEnergy (MeV)4142434445464748
NiobiumMolybdenumTechnetiumRutheniumRhodiumPalladiumSilverCadmium
Ag K[\beta _1]2.494E−024.36E+015.25E+014.84E+015.06E+015.35E+015.55E+011.01E+011.06E+01
Pd K[\beta _1]2.382E−024.92E+016.03E+015.45E+015.69E+016.01E+011.06E+011.15E+011.20E+01
Rh K[\beta _1]2.272E−025.56E+016.80E+016.15E+017.00E+011.14E+011.21E+011.30E+011.36E+01
Ag K[\bar {\alpha }]2.210E−025.98E+017.20E+016.60E+011.14E+011.23E+011.30E+011.40E+011.46E+01
Pd K[\bar {\alpha }]2.112E−026.71E+017.71E+017.41E+011.29E+011.39E+011.47E+011.58E+011.66E+01
Rh K[\bar {\alpha }]2.017E−027.55E+017.95E+011.38E+011.47E+011.58E+011.67E+011.79E+011.88E+01
Mo K[\beta _1]1.961E−028.10E+011.38E+011.49E+011.58E+011.70E+011.80E+011.94E+012.02E+01
Mo K[\bar {\alpha }]1.744E−021.77E+011.88E+012.04E+012.17E+012.33E+012.47E+012.65E+012.78E+01
Zn K[\beta _1]9.572E−039.04E+019.65E+011.04E+021.10E+021.18E+021.25E+021.34E+021.40E+02
Cu K[\beta _1]8.905E−031.10E+021.17E+021.26E+021.34E+021.44E+021.51E+021.63E+021.69E+02
Zn K[\bar {\alpha }]8.631E−031.20E+021.27E+021.37E+021.46E+021.56E+021.65E+021.77E+021.84E+02
Ni K[\beta _1]8.265E−031.34E+021.43E+021.54E+021.63E+021.75E+021.85E+021.98E+022.07E+02
Cu K[\bar {\alpha }]8.041E−031.45E+021.54E+021.66E+021.76E+021.89E+021.99E+022.13E+022.22E+02
Co K[\beta _1]7.649E−031.66E+021.76E+021.90E+022.01E+022.16E+022.27E+022.43E+022.53E+02
Ni K[\bar {\alpha }]7.472E−031.76E+021.88E+022.02E+022.14E+022.29E+022.41E+022.59E+022.69E+02
Fe K[\beta _1]7.058E−032.05E+022.19E+022.35E+022.49E+022.67E+022.81E+023.01E+023.13E+02
Co K[\bar {\alpha }]6.925E−032.16E+022.30E+022.47E+022.62E+022.80E+022.95E+023.16E+023.29E+02
Mn K[\beta _1]6.490E−032.57E+022.73E+022.94E+023.11E+023.33E+023.50E+023.75E+023.90E+02
Fe K[\bar {\alpha }]6.400E−032.67E+022.84E+023.05E+023.23E+023.46E+023.63E+023.89E+024.05E+02
Cr K[\beta _1]5.947E−033.25E+023.45E+023.70E+023.92E+024.20E+024.41E+024.72E+024.90E+02
Mn K[\bar {\alpha }]5.895E 033.32E+023.53E+023.79E+024.01E+024.29E+024.51E+024.83E+025.02E+02
Cr K[\bar {\alpha }]5.412E−034.16E+024.42E+024.74E+025.01E+025.36E+025.63E+026.02E+026.26E+02
Ti K[\beta _1]4.932E−035.32E+025.65E+026.04E+026.39E+026.83E+027.16E+027.65E+027.95E+02
Ti K[\bar {\alpha }]4.509E−036.71E+027.12E+027.61E+028.04E+028.60E+029.01E+029.61E+029.95E+02

RadiationEnergy (MeV)4950515253545556
IndiumTinAntimonyTelluriumIodineXenonCaesiumBarium
Ag K[\beta _1]2.494E−021.13E+011.18E+011.25E+011.29E+011.40E+011.46E+011.56E+011.62E+01
Pd K[\beta _1]2.382E−021.27E+011.34E+011.41E+011.46E+011.59E+011.65E+011.76E+011.83E+01
Rh K[\beta _1]2.272E−021.45E+011.52E+011.60E+011.66E+011.80E+011.88E+012.00E+012.08E+01
Ag K[\bar {\alpha }]2.210E−021.56E+011.64E+011.73E+011.79E+011.94E+012.02E+012.15E+012.24E+01
Pd K[\bar {\alpha }]2.112E−021.76E+011.85E+011.96E+012.02E+012.19E+012.29E+012.43E+012.54E+01
Rh K[\bar {\alpha }]2.017E−022.00E+012.10E+012.22E+012.29E+012.18E+012.27E+012.42E+012.52E+01
Mo K[\beta _1]1.961E−022.16E+012.26E+012.39E+012.47E+012.68E+012.80E+012.98E+013.10E+01
Mo K[\bar {\alpha }]1.744E−022.95E+013.10E+013.27E+013.38E+013.67E+013.82E+014.07E+014.23E+01
Zn K[\beta _1]9.572E−031.48E+021.55E+021.64E+021.68E+021.82E+021.90E+022.01E+022.09E+02
Cu K[\beta _1]8.905E−031.80E+021.88E+021.98E+022.04E+022.20E+022.29E+022.43E+022.52E+02
Zn K[\bar {\alpha }]8.631E−031.95E+022.04E+022.15E+022.21E+022.39E+022.49E+022.63E+022.73E+02
Ni K[\beta _1]8.265E−032.19E+022.29E+022.41E+022.48E+022.68E+022.78E+022.95E+023.06E+02
Cu K[\bar {\alpha }]8.041E−032.36E+022.47E+022.59E+022.67E+022.88E+022.99E+023.17E+023.25E+02
Co K[\beta _1]7.649E−032.69E+022.81E+022.96E+023.04E+023.30E+023.43E+023.63E+023.76E+02
Ni K[\bar {\alpha }]7.472E−032.86E+022.99E+023.14E+023.23E+023.49E+023.62E+023.83E+023.96E+02
Fe K[\beta _1]7.058E−033.32E+023.47E+023.65E+023.74E+024.08E+024.22E+024.46E+024.61E+02
Co K[\bar {\alpha }]6.925E−033.49E+023.64E+023.83E+023.94E+024.25E+024.40E+024.65E+024.80E+02
Mn K[\beta _1]6.490E−034.13E+024.31E+024.54E+024.66E+025.03E+025.20E+025.49E+025.66E+02
Fe K[\bar {\alpha }]6.400E−034.28E+024.47E+024.71E+024.83E+025.22E+025.40E+025.69E+025.86E+02
Cr K[\beta _1]5.947E−035.19E+025.42E+025.70E+025.85E+026.31E+026.52E+026.86E+026.45E+02
Mn K[\bar {\alpha }]5.895E−035.31E+025.54E+025.82E+025.98E+026.45E+026.66E+027.00E+026.60E+02
Cr K[\bar {\alpha }]5.412E−036.63E+026.91E+027.23E+027.40E+027.96E+027.21E+027.60E+025.70E+02
Ti K[\beta _1]4.932E−038.41E+028.76E+029.15E+029.32E+021.00E+031.03E+032.60E+023.14E+02
Ti K[\bar {\alpha }]4.509E−031.05E+031.09E+039.91E+027.51E+022.83E+022.65E+023.30E+023.34E+02

RadiationEnergy (MeV)5758596061626364
LanthanumCeriumPraseodymiumNeodymiumPromethiumSamariumEuropiumGadolinium
Ag K[\beta _1]2.494E−021.72E+011.83E+011.95E+012.04E+012.17E+012.23E+012.35E+012.42E+01
Pd K[\beta _1]2.382E−021.95E+012.07E+012.20E+012.30E+012.45E+012.52E+012.66E+012.74E+01
Rh K[\beta _1]2.272E−022.21E+012.35E+012.50E+012.61E+012.78E+012.86E+013.01E+013.10E+01
Ag K[\bar {\alpha }]2.210E−022.38E+012.53E+012.69E+012.81E+012.99E+013.08E+013.24E+013.34E+01
Pd K[\bar {\alpha }]2.112E−022.69E+012.86E+013.04E+013.18E+013.38E+013.48E+013.66E+013.77E+01
Rh K[\bar {\alpha }]2.017E−023.05E+013.24E+013.45E+013.60E+013.83E+013.94E+014.15E+014.27E+01
Mo K[\beta _1]1.961E−023.29E+013.49E+013.72E+013.88E+014.13E+014.24E+014.47E+014.60E+01
Mo K[\bar {\alpha }]1.744E−024.49E+014.77E+015.07E+015.30E+015.63E+015.78E+016.09E+016.26E+01
Zn K[\beta _1]9.572E−032.21E+022.33E+022.47E+022.57E+022.73E+022.79E+022.93E+023.00E+02
Cu K[\beta _1]8.905E−032.66E+022.82E+022.99E+023.10E+023.28E+023.35E+023.52E+023.60E+02
Zn K[\bar {\alpha }]8.631E−032.89E+023.06E+023.24E+023.36E+023.55E+023.63E+023.80E+023.89E+02
Ni K[\beta _1]8.265E−033.24E+023.43E+023.63E+023.76E+023.97E+024.05E+024.24E+024.33E+02
Cu K[\bar {\alpha }]8.041E−033.48E+023.68E+023.90E+024.04E+024.26E+024.34E+024.34E+024.03E+02
Co K[\beta _1]7.649E−033.95E+024.17E+024.41E+024.57E+024.82E+023.54E+024.80E+023.35E+02
Ni K[\bar {\alpha }]7.472E−034.19E+024.42E+024.68E+024.84E+025.11E+023.71E+023.75E+023.56E+02
Fe K[\beta _1]7.058E−034.83E+025.10E+025.39E+024.92E+025.88E+021.63E+024.08E+021.53E+02
Co K[\bar {\alpha }]6.925E−035.07E+025.35E+025.65E+025.05E+024.00E+021.76E+024.19E+021.61E+02
Mn K[\beta _1]6.490E−035.97E+025.47E+026.16E+024.39E+024.68E+021.66E+021.95E+021.89E+02
Fe K[\bar {\alpha }]6.400E−036.18E+025.61E+024.48E+024.55E+021.94E+022.04E+022.03E+021.95E+02
Cr K[\beta _1]5.947E−037.44E+024.94E+021.88E+021.98E+022.32E+022.21E+022.44E+022.35E+02
Mn K[\bar {\alpha }]5.895E−037.60E+025.12E+021.93E+022.03E+022.37E+022.25E+022.49E+022.41E+02
Cr K[\bar {\alpha }]5.412E−032.25E+022.38E+022.38E+022.51E+022.94E+022.79E+023.09E+022.98E+02
Ti K[\beta _1]4.932E−032.84E+023.00E+023.00E+023.14E+023.69E+023.50E+023.90E+023.74E+02
Ti K[\bar {\alpha }]4.509E−033.55E+023.57E+023.75E+023.97E+024.62E+024.35E+024.88E+024.69E+02

RadiationEnergy (MeV)6566676869707172
TerbiumDysprosiumHolmiumErbiumThuliumYtterbiumLutetiumHafnium
Ag K[\beta _1]2.494E−022.55E+012.65E+012.77E+012.90E+013.04E+013.14E+013.28E+013.40E+01
Pd K[\beta _1]2.382E−022.88E+012.99E+013.13E+013.27E+013.43E+013.54E+013.71E+013.84E+01
Rh K[\beta _1]2.272E−023.26E+013.39E+013.54E+013.71E+013.89E+014.01E+014.20E+014.35E+01
Ag K[\bar {\alpha }]2.210E−023.51E+013.65E+013.81E+013.99E+014.18E+014.32E+014.51E+014.67E+01
Pd K[\bar {\alpha }]2.112E−023.96E+014.12E+014.30E+014.50E+014.71E+014.87E+015.09E+015.27E+01
Rh K[\bar {\alpha }]2.017E−024.49E+014.66E+014.87E+015.09E+015.33E+015.50E+015.75E+015.95E+01
Mo K[\beta _1]1.961E−024.83E+015.02E+015.24E+015.48E+015.74E+015.93E+016.19E+016.41E+01
Mo K[\bar {\alpha }]1.744E−026.58E+016.83E+017.13E+017.44E+017.79E+018.04E+018.40E+018.69E+01
Zn K[\beta _1]9.572E−033.14E+023.24E+023.36E+023.49E+023.65E+023.75E+023.91E+021.00E+02
Cu K[\beta _1]8.905E−033.76E+023.87E+024.02E+024.17E+021.08E+021.08E+021.21E+021.20E+02
Zn K[\bar {\alpha }]8.631E−034.06E+024.19E+023.98E+022.87E+021.17E+021.17E+021.31E+021.30E+02
Ni K[\beta _1]8.265E−034.52E+023.36E+024.44E+021.23E+021.31E+021.31E+021.46E+021.45E+02
Cu K[\bar {\alpha }]8.041E−033.21E+023.62E+021.29E+021.32E+021.40E+021.42E+021.56E+021.55E+02
Co K[\beta _1]7.649E−033.60E+021.38E+021.46E+021.49E+021.59E+021.59E+021.78E+021.76E+02
Ni K[\bar {\alpha }]7.472E−031.49E+021.46E+021.55E+021.58E+021.69E+021.69E+021.89E+021.87E+02
Fe K[\beta _1]7.058E−031.71E+021.68E+021.78E+021.82E+021.96E+021.96E+022.18E+022.16E+02
Co K[\bar {\alpha }]6.925E−031.80E+021.76E+021.87E+021.91E+022.06E+022.06E+022.29E+022.27E+02
Mn K[\beta _1]6.490E−032.11E+022.07E+022.20E+022.24E+022.43E+022.44E+022.70E+022.67E+02
Fe K[\bar {\alpha }]6.400E−032.19E+022.14E+022.28E+022.32E+022.53E+022.51E+022.80E+022.77E+02
Cr K[\beta _1]5.947E−032.63E+022.57E+022.72E+022.78E+023.05E+023.04E+023.39E+023.34E+02
Mn K[\bar {\alpha }]5.895E−032.69E+022.62E+022.80E+022.85E+023.12E+023.11E+023.47E+023.41E+02
Cr K[\bar {\alpha }]5.412E−033.32E+023.25E+023.47E+023.52E+023.86E+023.87E+024.31E+024.25E+02
Ti K[\beta _1]4.932E−034.19E+024.10E+024.38E+024.43E+024.94E+024.92E+025.47E+025.39E+02
Ti K[\bar {\alpha }]4.509E−035.24E+025.15E+025.47E+025.54E+026.21E+026.19E+026.88E+026.78E+02

RadiationEnergy (MeV)7374757677787980
TantalumTungstenRheniumOsmiumIridiumPlatinumGoldMercury
Ag K[\beta _1]2.494E−023.54E+013.68E+013.83E+013.95E+014.11E+014.26E+014.44E+014.58E+01
Pd K[\beta _1]2.382E−024.00E+014.15E+014.32E+014.45E+014.64E+014.80E+015.00E+015.16E+01
Rh K[\beta _1]2.272E−024.53E+014.70E+014.89E+015.04E+015.24E+015.43E+015.65E+015.83E+01
Ag K[\bar {\alpha }]2.210E−024.87E+015.05E+015.25E+015.41E+015.63E+015.83E+016.07E+016.26E+01
Pd K[\bar {\alpha }]2.112E−025.48E+015.69E+015.92E+016.10E+016.34E+016.57E+016.83E+017.04E+01
Rh K[\bar {\alpha }]2.017E−026.20E+016.43E+016.69E+016.89E+017.16E+017.41E+017.71E+017.95E+01
Mo K[\beta _1]1.961E−026.67E+016.92E+017.19E+017.41E+017.70E+017.97E+018.29E+018.54E+01
Mo K[\bar {\alpha }]1.744E−029.04E+019.38E+019.74E+011.00E+021.04E+021.07E+021.12E+021.15E+02
Zn K[\beta _1]9.572E−031.02E+021.08E+021.19E+021.18E+021.23E+021.21E+021.30E+021.16E+02
Cu K[\beta _1]8.905E−031.22E+021.30E+021.43E+021.42E+021.48E+021.45E+021.55E+021.41E+02
Zn K[\bar {\alpha }]8.631E−031.32E+021.41E+021.55E+021.54E+021.60E+021.57E+021.68E+021.54E+02
Ni K[\beta _1]8.265E−031.47E+021.57E+021.72E+021.71E+021.78E+021.75E+021.88E+021.74E+02
Cu K[\bar {\alpha }]8.041E−031.58E+021.68E+021.87E+021.84E+021.91E+021.88E+022.01E+021.88E+02
Co K[\beta _1]7.649E−031.79E+021.91E+022.09E+022.09E+022.16E+022.14E+022.29E+022.16E+02
Ni K[\bar {\alpha }]7.472E−031.90E+022.03E+022.22E+022.21E+022.30E+022.27E+022.43E+022.30E+02
Fe K[\beta _1]7.058E−032.20E+022.34E+022.57E+022.55E+022.65E+022.61E+022.79E+022.60E+02
Co K[\bar {\alpha }]6.925E−032.31E+022.46E+022.68E+022.68E+022.78E+022.76E+022.95E+022.73E+02
Mn K[\beta _1]6.490E−032.73E+022.88E+023.16E+023.14E+023.30E+023.25E+023.48E+023.27E+02
Fe K[\bar {\alpha }]6.400E−032.83E+023.01E+023.27E+023.27E+023.40E+023.57E+023.61E+023.39E+02
Cr K[\beta _1]5.947E−033.39E+023.61E+023.94E+023.92E+024.11E+024.23E+024.34E+024.16E+02
Mn K[\bar {\alpha }]5.895E−033.46E+023.69E+024.05E+024.03E+024.18E+024.34E+024.45E+024.27E+02
Cr K[\bar {\alpha }]5.412E−034.32E+024.57E+025.01E+024.99E+025.20E+025.41E+025.51E+025.41E+02
Ti K[\beta _1]4.932E−035.46E+025.79E+026.33E+026.31E+026.59E+026.83E+026.99E+026.99E+02
Ti K[\bar {\alpha }]4.509E−036.85E+027.25E+027.94E+027.92E+028.26E+028.19E+028.76E+028.97E+02

RadiationEnergy (MeV)8182838485868788
ThalliumLeadBismuthPoloniumAstatineRadonFranciumRadium
Ag K[\beta _1]2.494E−024.72E+014.88E+015.06E+015.30E+015.51E+015.45E+015.67E+015.84E+01
Pd K[\beta _1]2.382E−025.31E+015.49E+015.70E+015.96E+016.20E+016.12E+016.37E+016.56E+01
Rh K[\beta _1]2.272E−026.00E+016.20E+016.44E+016.73E+017.00E+016.90E+017.18E+017.40E+01
Ag K[\bar {\alpha }]2.210E−026.45E+016.66E+016.91E+017.23E+017.51E+017.21E+017.70E+017.93E+01
Pd K[\bar {\alpha }]2.112E−027.25E+017.49E+017.77E+018.12E+018.43E+018.32E+018.64E+018.90E+01
Rh K[\bar {\alpha }]2.017E−028.18E+018.45E+018.76E+019.15E+019.50E+019.36E+019.72E+011.00E+02
Mo K[\beta _1]1.961E−028.79E+019.08E+019.41E+019.83E+011.02E+021.01E+021.04E+021.08E+01
Mo K[\bar {\alpha }]1.744E−021.18E+021.22E+021.26E+021.32E+021.17E+021.08E+028.70E+018.80E+01
Zn K[\beta _1]9.572E−031.45E+021.51E+021.57E+021.63E+021.71E+021.71E+021.77E+021.75E+02
Cu K[\beta _1]8.905E−031.75E+021.81E+021.88E+021.96E+021.86E+022.05E+022.13E+022.10E+02
Zn K[\bar {\alpha }]8.631E−031.89E+021.96E+022.04E+022.12E+022.07E+022.23E+022.30E+022.28E+02
Ni K[\beta _1]8.265E−032.11E+022.16E+022.28E+022.37E+022.31E+022.49E+022.57E+022.54E+02
Cu K[\bar {\alpha }]8.041E−032.26E+022.35E+022.44E+022.54E+022.48E+022.67E+022.77E+022.73E+02
Co K[\beta _1]7.649E−032.57E+022.67E+022.76E+022.88E+022.82E+023.04E+023.12E+023.10E+02
Ni K[\bar {\alpha }]7.472E−032.71E+022.83E+022.95E+023.05E+022.99E+023.21E+023.32E+023.29E+02
Fe K[\beta _1]7.058E−033.14E+023.27E+023.39E+023.54E+023.45E+023.73E+023.84E+023.80E+02
Co K[\bar {\alpha }]6.925E−033.31E+023.43E+023.55E+023.70E+023.63E+023.92E+024.03E+023.98E+02
Mn K[\beta _1]6.490E−033.90E+024.06E+024.21E+024.35E+024.26E+024.60E+024.77E+024.70E+02
Fe K[\bar {\alpha }]6.400E−034.03E+024.20E+024.34E+024.52E+024.44E+024.77E+024.93E+024.87E+02
Cr K[\beta _1]5.947E−034.87E+025.07E+025.24E+025.44E+025.33E+025.76E+025.97E+025.85E+02
Mn K[\bar {\alpha }]5.895E−035.00E+025.18E+025.35E+025.58E+025.45E+025.89E+026.02E+025.99E+02
Cr K[\bar {\alpha }]5.412E−035.97E+026.43E+026.66E+026.91E+026.80E+027.34E+027.58E+027.43E+02
Ti K[\beta _1]4.932E−037.15E+028.15E+028.44E+028.30E+028.60E+029.32E+029.61E+029.41E+02
Ti K[\bar {\alpha }]4.509E−039.89E+021.03E+031.06E+031.10E+031.08E+031.18E+031.21E+031.33E+03

RadiationEnergy (MeV)8990919293949596
ActiniumThoriumProtactiniumUraniumNeptuniumPlutoniumAmericiumCurium
Ag K[\beta _1]2.494E−026.07E+016.19E+016.48E+016.55E+016.84E+017.05E+017.20E+017.35E+01
Pd K[\beta _1]2.382E−026.82E+016.95E+017.27E+017.35E+017.67E+017.91E+018.08E+018.24E+01
Rh K[\beta _1]2.272E−027.68E+017.82E+018.19E+018.27E+018.63E+018.89E+019.08E+016.00E+01
Ag K[\bar {\alpha }]2.210E−028.24E+018.39E+018.78E+018.86E+019.25E+015.60E+015.95E+016.43E+01
Pd K[\bar {\alpha }]2.112E−029.24E+019.41E+019.84E+019.93E+011.04E+021.07E+021.09E+021.11E+02
Rh K[\bar {\alpha }]2.017E−021.04E+021.06E+021.11E+021.12E+021.16E+021.20E+021.22E+021.10E+02
Mo K[\beta _1]1.961E−021.10E+029.87E+011.19E+027.49E+011.25E+021.29E+021.31E+021.34E+02
Mo K[\bar {\alpha }]1.744E−029.08E+019.65E+011.01E+021.02E+024.22E+013.99E+014.81E+014.90E+01
Zn K[\beta _1]9.572E−032.49E+021.70E+021.73E+021.85E+021.90E+021.80E+021.89E+021.94E+02
Cu K[\beta _1]8.905E−032.85E+022.19E+022.08E+022.22E+022.27E+022.16E+022.27E+022.32E+02
Zn K[\bar {\alpha }]8.631E−033.03E+022.55E+022.25E+022.40E+022.46E+022.34E+022.41E+022.51E+02
Ni K[\beta _1]8.265E−033.09E+022.85E+022.52E+022.68E+022.75E+022.62E+022.73E+022.80E+02
Cu K[\bar {\alpha }]8.041E−033.17E+023.06E+022.71E+022.88E+023.14E+022.80E+023.22E+023.38E+02
Co K[\beta _1]7.649E−033.81E+023.48E+023.06E+023.26E+023.35E+023.17E+023.33E+023.43E+02
Ni K[\bar {\alpha }]7.472E−033.99E+023.69E+023.25E+023.47E+023.55E+023.36E+023.52E+023.60E+02
Fe K[\beta _1]7.058E−034.44E+023.89E+023.75E+024.00E+024.10E+023.89E+024.07E+024.21E+02
Co K[\bar {\alpha }]6.925E−034.61E+024.06E+023.94E+024.20E+024.30E+024.08E+024.26E+024.37E+02
Mn K[\beta _1]6.490E−035.21E+024.46E+024.65E+024.96E+025.05E+024.08E+025.03E+025.15E+02
Fe K[\bar {\alpha }]6.400E−035.30E+024.85E+024.82E+025.28E+025.52E+024.98E+025.81E+025.90E+02
Cr K[\beta _1]5.947E−036.18E+025.09E+025.82E+026.17E+026.30E+026.00E+026.27E+026.40E+02
Mn K[\bar {\alpha }]5.895E−036.29E+026.23E+025.93E+026.32E+026.45E+026.12E+026.42E+026.55E+02
Cr K[\bar {\alpha }]5.412E−037.39E+027.68E+027.38E+027.66E+028.00E+027.60E+027.95E+028.12E+02
Ti K[\beta _1]4.932E−038.83E+029.78E+029.83E+029.66E+021.01E+039.62E+021.03E+031.03E+03
Ti K[\bar {\alpha }]4.509E−031.05E+031.23E+031.24E+031.23E+039.65E+029.00E+029.55E+029.84E+02

RadiationEnergy (MeV)9798
BerkeliumCalifornium
Ag K[\beta _1]2.494E−026.66E+017.35E+01
Pd K[\beta _1]2.382E−027.52E+018.24E+01
Rh K[\beta _1]2.272E−028.51E+019.26E+01
Ag K[\bar {\alpha }]2.210E−026.10E+016.92E+01
Pd K[\bar {\alpha }]2.112E−021.03E+021.11E+02
Rh K[\bar {\alpha }]2.017E−021.02E+021.25E+02
Mo K[\beta _1]1.961E−021.25E+021.34E+02
Mo K[\bar {\alpha }]1.744E−024.90E+015.00E+01
Zn K[\beta _1]9.572E−031.86E+022.08E+02
Cu K[\beta _1]8.905E−032.26E+022.49E+02
Zn K[\bar{\alpha}]8.631E−032.46E+022.70E+02
Ni K[\beta _1]8.265E−032.77E+023.01E+02
Cu K[\bar {\alpha }]8.041E−033.52E+023.60E+02
Co K[\beta _1]7.649E−033.57E+023.66E+02
Ni K[\bar {\alpha }]7.472E−033.62E+023.86E+02
Fe K[\beta _1]7.058E−034.22E+024.48E+02
Co K[\bar {\alpha }]6.925E−034.43E+024.69E+02
Mn K[\beta _1]6.490E−035.26E+025.52E+02
Fe K[\bar {\alpha }]6.400E−035.92E+026.07E+02
Cr K[\beta _1]5.947E−036.64E+026.87E+02
Mn K[\bar {\alpha }]5.895E−036.78E+027.03E+02
Cr K[\bar {\alpha }]5.412E−038.52E+028.71E+02
Ti K[\beta _1]4.932E−031.09E+031.10E+03
Ti K[\bar {\alpha }]4.509E−031.04E+031.05E+03

Users of these tables should be aware of three important facts.

  • (i) The values given in the tables are derived for the case of isolated atoms, and cooperative effects may become important in condensed phases (Section 4.2.3[link]).

  • (ii) The values are based solely on theoretical calculations.

  • (iii) The limits to the reliability of the data when compared with experimental values are shown in Fig. 4.2.4.4[link].

The linear attenuation coefficient [\mu_l] in units of cm−1 can be defined operationally as [\mu _l= \bigg({\rm ln} {I_o\over I}\bigg)\bigg/t \eqno (4.2.4.1)]from the exponential attenuation relationship [{I \over I_0} = \exp (- \mu_lt) \eqno (4.2.4.2)]in which an idealized plane-parallel slab of material is interposed normally into a parallel beam of monoenergetic X-rays initially of intensity [I_0], attenuated by the interposed slab to a reduced intensity I.

The linear attenuation coefficient [\mu_l] for multi-element substances may be obtained in two ways. Through the mass absorption coefficients, we have [\mu _l = \rho\textstyle \sum\limits_i g_i (\mu _m)_i, \eqno (4.2.4.3)]where [g_i] is the mass fraction of the element i for which the mass attenuation coefficient [(\mu_m)_i] is in units of cm2 g−1, and ρ is the density of the material in units of g cm−3. The summation is over all the constituent elements. The mass attenuation coefficient [\mu _m] is sometimes written as [(\mu _l/\rho).]

For a crystal with unit-cell volume [V_c], [\mu _l = {1 \over V_c} \sum _i \sigma _i, \eqno (4.2.4.4)]where the summation is over all the atoms in the cell. If [\sigma _i] is in barns/atom and [V_c] is in Å3, then [\mu_l] is in cm−1.

These tables list total interaction cross sections and mass attenuation coefficients for isolated atoms calculated for characteristic X-ray photon emissions ranging from Ti Kα to Ag Kβ.

The total interaction cross section is defined by [\sigma = \sigma _{\rm pe} + \sigma _R + \sigma _C, \eqno (4.2.4.5)]where [\sigma _{\rm pe}] is the photo-effect cross section; [\sigma _R] is the Rayleigh (unmodified, elastic) cross section; [\sigma _C] is the Compton (modified, inelastic) cross section.

The reader's attention is drawn to the fact that in the neighbourhood of an absorption edge for aggregations of atoms significant deviations may be found because of cooperative effects (XAFS and XANES). A discussion of these effects is given in Section 4.2.3[link].

4.2.4.2. Sources of information

| top | pdf |

4.2.4.2.1. Theoretical photo-effect data: σpe

| top | pdf |

Of the many theoretical data sets in existence, those of Storm & Israel (1970[link]), Cromer & Liberman (1970[link]), and Scofield (1973[link]) have often been used as bench marks against which both experimental and theoretical data have been compared. In particular, theoretical data produced using the S-matrix approach have been compared with these values. See, for example, Kissel, Roy & Pratt (1980[link]). Some indication of the extent to which agreement exists between the different theoretical data sets is given in §4.2.6.2.4[link] (Tables 4.2.6.3[link](b) and 4.2.6.5[link]). These tables show that the values of [f'(\omega,0)], which is proportional to σ, calculated using modern relativistic quantum mechanics, agree to better than 1%. It has also been demonstrated by Creagh & Hubbell (1987[link], 1990[link]) in their analysis of the results of the IUCr X-ray Attenuation Project that there appears to be no rational basis for preferring one of these data sets over the other.

These tables do not list separately photo-effect cross sections. However, should these be required, the data can be found using Table 4.2.6.8[link]. The cross section in barns/atom is related to [f'(\omega,0)] expressed in electrons/atom by σ = 5636λ[\,f'(\omega, 0),] where λ is expressed in ångströms.

The values for [\sigma_{\rm pe}] used in this compilation are derived from recent tabulations based on relativistic Hartree–Fock–Dirac–Slater calculations by Creagh. The extent to which this data set differs from other theoretical and experimental data sets has been discussed by Creagh (1990[link]).

4.2.4.2.2. Theoretical Rayleigh scattering data: σR

| top | pdf |

If each of the atoms gives rise to scattering in which momentum but not energy changes occur, and if each of the atoms can be considered to scatter as if it were an isolated atom, the cross section may be written as [\sigma _R = \pi r ^2_e \textstyle\int\limits^1_{-1} (1 + \cos ^2 \varphi) \,f^2(q,Z) \,{\rm d} (\cos \varphi), \eqno (4.2.4.6)]where

  • [r_e] is the classical radius of the electron;

  • [\varphi] is the angle of scattering ([=2 \theta] if [\theta] is the Bragg angle);

  • [2 \pi \,{\rm d} (\cos \varphi)] is the solid angle between cones with angles [\varphi] and [\varphi + {\rm d} \varphi];

  • [f(q,Z)] is the atomic scattering factor as defined by Cromer & Waber (1974[link]);

  • q is [ [\sin (\varphi /2) / \lambda],] the momentum transfer parameter. Here λ is expressed in ångströms.

Reliable tables of f(q, Z) exist and have been reviewed recently by Kane, Kissel, Pratt & Roy (1986[link]). The most recent schematic tabulations of f(q, Z) are those of Hubbell & Øverbø (1979[link]) and Schaupp et al. (1983[link]). The data used in these tables have been derived from the tabulation for q = 0.02 to 109 Å−1, for all Z's from 1 to 100 by Hubbell & Øverbø (1979[link]) based on the exact formula of Pirenne (1946[link]) for H, and relativistic calculations by Doyle & Turner (1968[link]), Cromer & Waber (1974[link]), Øverbø (1977[link], 1978[link]), and high-q extensions using the Bethe–Levinger expression in Levinger (1952[link]).

As mentioned in Creagh & Hubbell (1987[link]), the atoms in highly ordered single crystals do not scatter as though they are isolated atoms. Rather, cooperative effects become important. In this case, the Rayleigh scattering cross section must be replaced by two cross sections:

  • the Laue–Bragg cross section [\sigma _{\rm LB}],

  • and the thermal diffuse scattering cross section [\sigma _{\rm TD}].

That is, [\sigma _R] is replaced by [\sigma _{\rm LB} + \sigma _{\rm TD}].

These effects are discussed elsewhere (Subsection 4.2.3.2[link]). Briefly, [\sigma _{\rm LB} = (r^2_e \lambda ^2 / 2 NV_c) \textstyle\sum\limits _H [C_p md | F | {}^2 \exp (-2M)]_H. \eqno (4.2.4.7)]In equation (4.2.4.7)[link], which is due to De Marco & Suortti (1971[link]),

  • [C_p= \textstyle {1 \over 2}(1+ \cos ^2 \varphi)];

  • [d_H] is the spacing of the (hkl) planes in the crystal;

  • [m_H] is the multiplicity of the hkl Bragg reflection;

  • [F_H] is the geometrical structure factor for the crystal structure that contains N atoms in a cell of volume [V_c];

  • [\exp (-2M)_H] is the Debye–Waller temperature factor.

It is assumed that the total thermal diffuse scattering is equal to the scattering lost from Laue–Bragg scattering because of thermal vibrations. [\sigma _{\rm TD} = (r^2_e \lambda ^2/2NV_c) \textstyle\sum\limits _H \{ C_p md | F | {}^2 [1- \exp (-2 M)] \} _H. \eqno (4.2.4.8)]This equation is not in a convenient form for computation and the alternative formalism presented by Sano, Ohtaka & Ohtsuki (1969[link]) is often used in calculations. In this formalism, [\sigma _{\rm TD}= 2 \pi r ^2 _e \textstyle\int \limits^1 _{-1} C _p\,f^2 (q,Z) \{ 1 - \exp [-2M (q)] \} \,{\rm d} (\cos \varphi). \eqno (4.2.4.9)]

The values of f(q, Z) are those of Cromer & Waber (1974[link]).

Cross sections calculated using equation (4.2.4.8)[link] tend to oscillate at low energy and this corresponds to the inclusion of Bragg peaks in the summation or integration. Eventually, these oscillations abate and [\sigma _{\rm TD}] becomes a smoothly varying function of energy.

Creagh & Hubbell (1987[link]) and Creagh (1987a[link]) have stressed that, before cross sections are calculated for a given ensemble of atoms, care should be taken to ascertain whether single-atom or single-crystal scattering is appropriate for that ensemble.

4.2.4.2.3. Theoretical Compton scattering data: σC

| top | pdf |

The bound-electron Compton scattering cross section is given by [\eqalignno{ \sigma _C = {}&\pi r \,^2_e \textstyle\int\limits^1_{-1} [1 + k (1 - \cos \varphi)] ^{-2}\cr &\times \{ + \cos ^2 \varphi + k ^2 (1 - \cos \varphi) ^2\cr &\times [1 + k (1 - \cos \varphi)] ^{-1} \} I (q,z)\, {\rm d} (\cos \varphi). &(4.2.4.10)}]Here [k = \hbar \omega /mc^2] and [I (q, z)] is the incoherent scattering intensity expressed in electron units. The other symbols have the meanings defined in §§4.2.4.2.1[link] and 4.2.4.2.2[link].

Values of [\sigma _C] incorporated into the tables of total cross section σ have been computed using the incoherent scattering intensities from the tabulation by Hubbell et al. (1975[link]) based on the calculations by Cromer & Mann (1967[link]) and Cromer (1969[link]).

4.2.4.3. Comparison between theoretical and experimental data sets

| top | pdf |

Saloman & Hubbell (1986[link]) and Saloman et al. (1988[link]) have published an extensive comparison of the experimental database with the theoretical values of Scofield (1973[link], 1986[link]) for photon energies between 0.1 and 100 keV. Some examples taken from Saloman & Hubbell (1986[link]) are shown in Figs. 4.2.4.1[link] , 4.2.4.2[link] , and 4.2.4.3[link] .

[Figure 4.2.4.1]

Figure 4.2.4.1| top | pdf |

Agreement between theory and experiment for oxygen (Z = 8) in the `soft' X-ray region. The solid line is for the Scofield (1973[link]) values without renormalization and the dotted line is for the semi-empirical data of Henke et al. (1982[link]).

[Figure 4.2.4.2]

Figure 4.2.4.2| top | pdf |

The total cross section for silicon (Z = 14) compared with the unrenormalized Scofield values. The measured and theoretical attenuation coefficients show systematic differences of several percent for the photon energy range 10 to 100 keV.

[Figure 4.2.4.3]

Figure 4.2.4.3| top | pdf |

The total cross section for uranium (Z = 92): The theoretical values (solid line) are partially obscured by the high density of available measurements. Deviations of the measured values from the theoretical predictions are mostly of the order of 5%, although a few data sets deviate by more than 30%.

Comparisons between theory and experiment exist for about 80 elements and space does not permit reproduction of all the available information. This information has been summarized in Fig. 4.2.4.4[link]. Superimposed on the Periodic Table of the elements are two sets of data. The upper set corresponds to the average percent deviation between experiment and theory for the photon energy range 10 to 100 keV. The lower set corresponds to the average percent deviation between experiment and theory for the photon energy range 1 to 10 keV. An upwards pointing arrow [\uparrow] means that [(\sigma _{\exp}- \sigma _{\rm theor}) \,\gt \, 0] . No arrow implies that [(\sigma _{\exp} - \sigma _{\rm theor})=0.] A downwards pointing arrow [\downarrow] means that [(\sigma _{\exp} - \sigma _{\rm theor}) \lt 0.] An asterisk means no experimental data set was available.

[Figure 4.2.4.4]

Figure 4.2.4.4| top | pdf |

Comparison between this tabulation and experimental data contained in Saloman & Hubbell (1986[link]). The upper set corresponds to the average percent deviation between the experimental data and this tabulation for the energy range 10 to 100 keV. The lower set corresponds to the energy range 1 to 10 keV. For explanation of symbols see text.

For example: for tin (Z = 50), the experimental data are on average 5% higher than the theoretical predictions for the range of photon energies from 10 to 100 keV. For the range 1 to 10 keV, the experimental data are on average 7% higher than the theoretical predictions.

Fig. 4.2.4.4[link] is given as a rapid means of comparing theory and experiment. For more detailed information, see Saloman & Hubbell (1986[link]), Saloman et al. (1988[link]), and Creagh (1990[link]).

4.2.4.4. Uncertainty in the data tables

| top | pdf |

It is not possible to generalize on the accuracy of the experimental data sets. Creagh & Hubbell (1987[link]) have shown that many experiments for which the precision quoted by the author is high differ from other accurate measurements by a considerable amount. It must be stressed that the experimental apparatus has to be chosen so that it is appropriate for the atomic system being investigated. Details concerning the proper choice of measuring system are given in Section 4.2.3[link]. Within about 200 eV of an absorption edge, deviations of up to 200% may be observed between theory and experiment. This is the region in which XAFS and XANES oscillations occur.

With respect to the theoretical data: the detailed agreement between the several methods for calculating the photo-effect cross sections is quite remarkable and it is estimated that the reliability of these data is to within 2% for the energy range considered in this compilation. Some problems may exist, however, close to the absorption edges. Errors in the calculation of the Rayleigh and the Compton scattering cross sections are assessed to be of the order of 5%. Because the greater proportion of total attenuation is photoelectric, the accuracy of the total scattering cross section should be much better than 5% and usually close to 2%.

4.2.5. Filters and monochromators

| top | pdf |
D. C. Creaghb

4.2.5.1. Introduction

| top | pdf |

All sources of X-rays, whether they be produced by conventional sealed tubes, rotating-anode systems, or synchrotron-radiation sources, emit over a broad spectral range. In many cases, this spectral diversity is of concern, and techniques have been developed to minimize the problem. These techniques involve the use of filters, mirrors, and Laue and Bragg crystal monochromators, chosen so as to provide the best compromise between flux and spectral purity in a particular experiment. In other chapters, authors have discussed the use of techniques to improve the spectral purity of X-ray sources. This section does not purport to be a comprehensive exposition on the topic of filters and monochromators. Rather, it seeks to point the reader towards the information given elsewhere in this volume, and to add complementary information where necessary. A search of the Subject Index will find references to filters and monochromators that are not explicitly mentioned in the text of this section.

The ability to select photon energies, or bands of energies, depends on the scattering power of the atoms from which the monochromator is made and the arrangement of the atoms within the monochromator. The scattering powers of the atoms and their dependence on the energy of the incident photons were discussed in Sections 4.2.3[link] and 4.2.4[link] and are discussed more fully in Section 4.2.6[link]. In brief, the scattering power of the atom, or atomic scattering factor, is defined, for a given incident photon energy, as the ratio of the scattering power of the atom to that of a free Thomson electron. The scattering power is denoted by the symbol [f(\omega, \Delta)] and is a complex quantity, the real part of which, [f^{\prime }(\omega, \Delta)], is related to the elastic scattering cross section, and the imaginary part of which, [f^{\prime \prime }(\omega, \Delta)], is related directly to the photoelectric scattering cross section and therefore the linear attenuation coefficient [\mu _l].

At an interface between, say, air and the material from which the monochromator is made, reflection and refraction of the incident photons can occur, as dictated by Maxwell's equations. There is an associated refractive index n given by [ n=(1+\chi)^{1/2}, \eqno (4.2.5.1)]where [ \chi =-(r_e\lambda ^2/\pi)\textstyle \sum \limits _jN_j\;f_j(\omega, \Delta), \eqno (4.2.5.2)][r_e] is the classical radius of the electron, and [N_j] is the number density of atoms of type j.

An angle of total external reflection [\alpha _{c}] exists for the material, which is a function of the incident photon energy, since [f_{j}(\omega, \Delta)] is a function of photon energy. Thus, a polychromatic beam incident at the critical angle of one of the photon energies (E) will reflect totally those components having energies less than E, and transmit those components with energies greater than E. Fig. 4.2.5.1[link] shows calculations by Fukumachi, Nakano & Kawamura (1986[link]) for the reflectivity of single layers of aluminium, copper and platinum as a function of incident energy for a fixed angle of incidence (0.2°). For the aluminium specimen, the reflectivity curve shows the rapid decrease in reflectivity as the critical angle is exceeded. The reflectivity in this region varies as [E^{-2}]. The effect of increasing atomic number can be seen: the higher the atomic factor [f(\omega, \Delta)], the greater the energy that can be reflected from the surface. Also visible are the effects of the dispersion corrections [f^{\prime }(\omega, \Delta)] and [f^{\prime \prime }(\omega, \Delta)] on reflectivity. For copper, the K shell is excited, and for platinum the [L_{\rm I}], [L_{\rm II}] and [L_{\rm III}] shells are excited by the polychromatic beam.

[Figure 4.2.5.1]

Figure 4.2.5.1| top | pdf |

The variation of specular reflectivity with incident photon energy is shown for materials of different atomic number and a constant angle of incidence of 0.2°. (a) Aluminium: note the rapid decrease of reflectivity with energy. (b) Copper: the sudden decrease of reflectivity is due to the modification of the scattering-length density owing to absorption at the K-absorption edge. (c) Platinum: the three discontinuities in the reflectivity curve are due to absorption at the LI-, LII-, and LIII-absorption edges.

Interfaces can therefore be used to act as low-pass energy filters. The surface roughness and the existence of impurities and contaminants on the interface will, however, influence the characteristics of the reflecting surface, sometimes significantly.

4.2.5.2. Mirrors and capillaries

| top | pdf |

Whilst neither of these classes of X-ray optical device is strictly speaking a monochromator, they nevertheless form component parts of monochromator systems in the laboratory and at synchrotron-radiation sources.

4.2.5.2.1. Mirrors

| top | pdf |

In the laboratory, mirrors are used in conjunction with conventional sealed tubes and rotating-anode sources, the emission from which consists of Bremsstrahlung upon which is superimposed the characteristic spectrum of the anode material (Subsection 2.3.5.2[link] ). The shape of the Bremsstrahlung spectrum can be significantly modified by mirrors, and the intensity emitted at harmonics of the characteristic wavelength can be significantly reduced. More importantly, the mirrors can be fashioned into shapes that enable the emitted radiation to be brought to a focus. Ellipsoidal, logarithmic spiral, and toroidal mirrors have been manufactured commercially for use with laboratory X-ray sources. Since the X-rays are emitted isotropically from the anode surface, it is important to devise a mirror system that has a maximum angle of acceptance and a relatively long focal length.

At synchrotron-radiation sources, the high intensities that are generated over a very broad spectral range give rise to significant heat loading of subsequent monochromators and therefore degrade the performance of these elements. In many systems, mirrors are used as the first optical element in the monochromator, to reduce the heat load on the primary monochromator and to make it easier for the subsequent monochromators to reject harmonics of the chosen radiation. Shaped mirror geometries are often used to focus the beam in the horizontal plane (Subsection 2.2.7.3[link] ). A schematic diagram of the optical elements of a typical synchrotron-radiation beamline is shown in Fig. 4.2.5.2[link] . In this, the primary mirror acts as a thermal shunt for the subsequent monochromator, minimizes the high-energy component that may give rise to possible harmonic content in the final beam, and acts as a vertical collimator. The radii of curvature of mirrors can be changed using a mechanical four-point bending system (Oshima, Harada & Sakabe, 1986[link]). More recent advances in mirror technology enable the shape of the mirror to be changed through use of the piezoelectric effect (Sussini & Labergerie, 1995[link]).

[Figure 4.2.5.2]

Figure 4.2.5.2| top | pdf |

The use of mirrors in a typical synchrotron-radiation beamline. The X-rays are emitted tangentially to the orbit of the stored positron beam. They pass through a beam-defining slit onto a mirror that serves three purposes, viz energy discrimination, heat absorption, and focusing, by means of a mechanical four-point bending system. The beam then passes into a double-crystal monochromator, which selects the desired photon energy. The second element of this monochromator is capable of being bent sagittally using a mechanical four-point bending system to focus the beam in the horizontal plane. The beam is then refocused and redirected by a second mirror.

4.2.5.2.2. Capillaries

| top | pdf |

Capillaries, and bundles of capillaries, are finding increasing use in situations where a focused beam is required. The radiation is guided along the capillary by total external reflection, and the shape of the capillary determines the overall flux gain and the uniformity of the focused spot. Gains in flux of 100 and better have been reported. There is, however, a degradation in the angular divergence of the outgoing beam. For single capillaries, applications are laboratory-based protein crystallography, microtomography, X-ray microscopy, and micro-X-ray fluorescence spectroscopy. The design and construction of capillaries for use in the laboratory and at synchrotron-radiation sources has been discussed by Bilderback, Thiel, Pahl & Brister (1994[link]), Balaic & Nugent (1995[link]), Balaic, Nugent, Barnea, Garrett & Wilkins (1995[link]), Balaic et al. (1996[link]), and Engström, Rindby & Vincze (1996[link]). They are usually used after other monochromators in these applications and their role as a low-pass energy filter is not of much significance.

Bundles of capillaries are currently being produced commercially to produce focused beams (ellipsoidally shaped bundles) and half-ellipsoidal bundles are used to form beams of large cross section from conventional laboratory sources (Peele et al., 1996[link]; Kumakhov & Komarov, 1990[link]).

4.2.5.2.3. Quasi-Bragg reflectors

| top | pdf |

For one interface, the reflectivity (R) and the transmissivity (T) of the surface are determined by the Fresnel equations, viz: [ R=\left | (\theta _{1}-\theta _{2})/(\theta _{1}+\theta _{2})\right | ^{2}, \eqno (4.2.5.3)]and [ T=\left | 2\theta _{1}/(\theta _{1}+\theta_{2})\right | ^{2}, \eqno (4.2.5.4)]where [\theta _{1}] and [\theta _{2}] are the angles between the incident ray and the surface plane and the reflected ray and the surface plane, respectively.

If a succession of interfaces exists, the possibility of interference between successively reflected rays exists. Param­eters that define the position of the interference maxima, the line breadths of those maxima, and the line intensity depend inter alia on the regularity in layer thickness, the interface surface roughness, and the existence of surface tilts between successive interfaces. Algorithms for solving this type of problem are incorporated in software currently available from a number of commercial sources (Bede Scientific, Siemens, and Philips). The reflectivity profile of a system having a periodic layer structure is shown in Fig. 4.2.5.3[link] . This is the reflectivity profile for a multiple-quantum-well structure of alternating aluminium gallium arsenide and indium gallium arsenide layers (Holt, Brown, Creagh & Leon, 1997[link]). Note the interference maxima that are superimposed on the Fresnel reflectivity curve. From the full width at half-maximum of these interference lines, it can be inferred that the energy discrimination of the system, ΔE/E, is 2%. The energy range that can be reflected by such a multilayer system depends on the interlayer thickness: the higher the photon energy, the thinner the layer thickness.

[Figure 4.2.5.3]

Figure 4.2.5.3| top | pdf |

The reflectivity of a multiple-quantum-well device is shown. This consists of 40 alternating layers of AlGaAs and InGaAs. Shown also, but shifted downwards on the vertical scale for the purpose of clarity, is the theoretical prediction based on standard electromagnetic theory.

Commercially available multilayer mirrors exist, and hitherto they have been used as monochromators in the soft X-ray region in X-ray fluorescence spectrometers. These monochromators are typically made of alternating layers of tungsten and carbon, to maximize the difference in scattering-length density at the interfaces. Whilst the energy resolution of such systems is not especially good, these monochromators have a good angle of acceptance for the incident beam, and reasonably high photon fluxes can be achieved using conventional laboratory sources.

A recent development of this, the Goebel mirror system, is supplied as an accessory to a commercially available diffractometer (Siemens, 1996a[link],b[link],c[link]; OSMIC, 1996[link]). This system combines the focusing capacity of a curved mirror with the energy selectivity of the multilayer system. The spacing between layers in this class of mirror multilayers can be laterally graded to enhance the incident acceptance angle. These multilayers can be fixed to mirrors of any figure to a precision of 0.3′ and can therefore can be used to form parallel beams (parabolic optical elements) as well as focused beams (elliptical optical elements) of high quality.

4.2.5.3. Filters

| top | pdf |

It is usual to consider only the cases where a quasimonochromatic beam is to be extracted from a polychromatic beam. Before discussing this class of usage, mention must be made of two simple forms of filtering of radiation.

In the first, screening, a thin layer of absorbing foil is used to reduce the effect of specimen fluorescence on photon counting, film and imaging-plate detectors. A typical example is the use of aluminium foil in front of a Polaroid camera used in a Laue camera to reduce the K-shell fluorescence radiation from a transition-metal crystal when using a conventional sealed molybdenum X-ray source. A 0.1 mm thick foil will reduce the fluorescent radiation from the crystal by a factor of about five, and this radiation is emitted isotropically from the specimen. In contrast, the wanted Laue-reflected beams are emitted as a nearly parallel beam, and the signal-to-noise ratio in the resulting photograph is much increased.

The second case is the ultimate limiting case of filtering, shielding. If it is necessary to shield an object completely from a polychromatic incident beam, a sufficient thickness of absorbing material, calculated using the data in Section 4.2.4[link], to reduce the beam intensity to the level of the ambient background is inserted in the beam. [The details of how shielding systems are designed are given in reference works such as the Handbook of Radiation Measurement and Protection (Brodsky, 1982[link]).] In general, the use of an absorber of one atomic species will provide insufficient shielding. The use of composite absorbers is necessary to achieve a maximum of shielding for a minimum of weight. This is of utmost importance if one is designing, say, the shielding of an X-ray telescope to be carried in a rocket or a balloon (Grey, 1996[link]). To produce shielding that satisfies the requirements of minimum weight, good mechanical rigidity, and ability to be constructed to good levels of mechanical tolerance, shielding must be constructed using a number of layers of different absorbers, chosen such that the highest-energy radiation is just stopped in the first layer, the L-shell fluorescent radiation created in the absorption process is stopped in the second, and the lower-energy L and M fluorescent radiation is stopped by the next layer, and so on until the desired radiation level is reached.

In the usual case involving filters, the problem is one of removing as much as possible of the Bremsstrahlung radiation and unwanted characteristic radiation from the spectrum of a laboratory sealed tube or rotating-anode source whilst retaining as much of the wanted radiation as is possible. To give an example, a thin characteristic radiation filter of nickel of appropriate thickness almost completely eliminates the Bremsstrahlung and Kβ radiation from an X-ray source with a copper target, but reduces the intensity in the Cu Kα doublet by only about a factor of two. For many applications, this is all that is necessary to provide the required degree of monochromatization. If there is a problem with the residual Bremsstrahlung, this problem may be averted by making a second set of measurements with a different filter, one having an absorption edge at an energy a little shorter than that of the desired emission line. The difference between the two sets of measurements corresponds to a comparatively small energy range spanning the emission line. This balanced-filter method is more cumbersome than the single-filter method, but no special equipment or difficult adjustments are required. In general, if the required emission is from an element of atomic number Z, the first foil is made from material having atomic number Z − 1 and the second from atomic number Z + 1. A better balance can be achieved using three foils (Young, 1963[link]). The use of filters is discussed in more detail in §2.3.5.4.2[link] . Data for filters for the radiations in common use are given in Tables 2.3.5.2[link] and 2.3.5.3[link] . The information necessary for choosing filter materials and estimating their optimum thicknesses for other radiations is given in Sections 4.2.2[link], 4.2.3[link], and 4.2.4[link].

It should be remembered that filtration changes the wavelength of the emission line slightly, but this is only of significance for measurements of lattice parameters to high precision (Delf, 1961[link]).

4.2.5.4. Monochromators

| top | pdf |

4.2.5.4.1. Crystal monochromators

| top | pdf |

Even multifoil balanced filters transmit a wide range of photon energies. Strictly monochromatic radiation is impossible, since all atomic energy levels have a finite width, and emission from these levels therefore is spread over a finite energy range. The corresponding radiative line width is important for the correct evaluation of the dispersion corrections in the neighbourhood of absorption edges (§4.2.6.3.3.2[link]). Even Mössbauer lines, originating as they do from nuclear energy levels that are much narrower than atomic energy levels, have a finite line width. To achieve line widths comparable to these requires the use of monochromators using carefully selected single-crystal reflections.

Crystal monochromators make use of the periodicity of `perfect' crystals to select the desired photon energy from a range of photon energies. This is described by Bragg's law, [ 2d_{hkl}\sin \theta =n\lambda, \eqno (4.2.5.5)]where [d_{hkl}] is the spacing between the planes having Miller indices hkl, [\theta] is the angle of incidence, n is the order of a particular reflection (n = 1, 2, 3[\ldots]), and λ is the wavelength.

If there are wavelength components with values near λ/2, λ/3, [\ldots], these will be reflected as well as the wanted radiation, and harmonic contamination can result. This can be a difficulty in spectroscopic experiments, particularly XAFS, XANES and DAFS (Section 4.2.3[link]).

Equation (4.2.5.5)[link] neglects the effect of the refractive index of the material. This is usually omitted from Bragg's law, since it is of the order of 10−5 in magnitude. Because the refractive index is a strong function of wavelength, the angles at which the successive harmonics are reflected are slightly different from the Bragg angle of the fundamental. This fact can be used in multiple-reflection monochromators to minimize harmonic contamination.

As can be seen in Fig. 4.2.5.4[link] , each Bragg reflection has a finite line width, the Darwin width, arising from the interaction of the radiation with the periodic electron charge distribution. [See, for example, Warren (1968[link]) and Subsection 7.2.2.1[link] .] Each Bragg reflection therefore contains a spread of photon energies. The higher the Miller indices, the narrower the Darwin width becomes. Thus, for experiments involving the Mössbauer effect, extreme back-reflection geometry is used at the expense of photon flux.

[Figure 4.2.5.4]

Figure 4.2.5.4| top | pdf |

In (a), the schematic rocking curve for a silicon crystal in the neighbourhood of the 111 Bragg peak is shown. The full curve is due to the crystal set to the true Bragg angle, and the dotted curve corresponds to a surface tilted at an angle of 2′′ with respect to the beam prior to the acquisition of the rocking curve. Only the 111 and 333 reflections are shown for clarity. The 222 reflection is very weak because the geometrical structure factor is small. The separation of the 111 and 333 peaks occurs because the refractive index is different for these reflections. In a double-crystal monochromator, white radiation from the source will produce the scattered intensity given by the full curve. If that intensity distribution now falls on the second crystal, which is tilted with respect to the first, an angle of tilt can be found for which the Bragg condition is not fulfilled in the second crystal, and the 333 radiation cannot be reflected. The resultant reflected intensity is shown in (b). Note that this is an idealized case, and in practice the existence of tails in the reflectivity curve can allow the transmission of some harmonic radiation through the double-crystal monochromator.

If the beam propagates through the specimen, the geometry is referred to as transmission, or Laue, geometry. If the beam is reflected from the surface, the geometry is referred to as reflection, or Bragg, geometry. Bragg geometry is the most commonly used in the construction of crystal monochromators. Laue geometry has been used in only a relatively few applications until recently. The need to handle high photon fluxes with their associated high power load has led to the use of diamond crystals in Laue configurations as one of the first components of X-ray optical systems (Freund, 1993[link]). Phase plates can be created using the Laue geometry (Giles et al., 1994[link]). A schematic diagram of a system used at the European Synchrotron Radiation Facility is shown in Fig. 4.2.5.5[link] . Radiation from an insertion device falls on a Laue-geometry pre-monochromator and then passes through a channel-cut (multiple-reflection) monochromator. The strong linear polarization from the source and the monochromators can be changed into circular polarization by the asymmetric Laue-geometry polarizer and analysed by a similar Laue-geometry analysing crystal.

[Figure 4.2.5.5]

Figure 4.2.5.5| top | pdf |

A schematic diagram of a beamline designed to produce circularly polarized light from initially linearly polarized light using Laue-case reflections. Radiation from an insertion device is initially monochromated by a cooled diamond crystal, operating in Laue geometry. The outgoing radiation has linear polarization in the horizontal plane. It then passes through a silicon channel-cut monochromator and into a silicon crystal of a thickness chosen so as to produce equal amounts of radiation from the σ and π branches of the dispersion surface. These recombine to produce circularly polarized radiation at the exit surface of the crystal.

More will be said about polarization in §4.2.5.4.4[link] and Section 6.2.2[link] .

4.2.5.4.2. Laboratory monochromator systems

| top | pdf |

Many laboratories use powder diffractometers using the Bragg–Brentano configuration. For these, a sufficient degree of monochromatization is achieved through the use of a diffracted-beam monochromator consisting of a curved-graphite monochromator and a detector, both mounted on the 2θ arm of the diffractometer. Such a device rejects the unwanted Kβ radiation and fluorescence from the sample with little change in the magnitude of the Kα lines. Incident-beam monochromators are also used to produce closely monochromatic beams of the desired energy. Single-reflection monochromators used for the reduction of spectral energy spread are described in Subsections 2.2.7.2[link] and 2.3.5.4[link] .

For most applications, this simple means of monochromatization is adequate. Increasingly, however, more versatility and accuracy are being demanded of laboratory diffractometer systems. Increased angular accuracy in both the θ and 2θ axes, excellent monochromatization, and parallel-beam geometry are all demands of a user community using improved techniques of data collection and data analysis. The necessity to study thin films has generated a need for accurately collimated beams of small cross section, and there is a need to have well collimated and monochromatic beams for the study of rough surfaces. This, coupled with the need to analyse data using the Rietveld method (Young, 1993[link]), has caused a revolution in the design of commercial diffractometers, with the use of principles long since used in synchrotron-radiation research for the design of laboratory instruments. Monochromators of this type are briefly discussed in §4.2.5.4.3[link].

4.2.5.4.3. Multiple-reflection monochromators for use with laboratory and synchrotron-radiation sources

| top | pdf |

Single-reflection devices produce reflected beams with quite wide, quasi-Lorentzian, tails (Subsection 2.3.3.8[link] ), a situation that is not acceptable, for example, for the study of small-angle scattering (SAXS, Chapter 2.6[link] ). The effect of the tails can be reduced significantly through the use of multiple Bragg reflections.

The use of multiple Bragg reflections from a channel cut in a monolithic silicon crystal such that the channel lay parallel to the (111) planes of the crystal was shown by Bonse & Hart (1965b[link]) to remove the tails of reflections almost completely. This class of device, referred to as a (symmetrical) channel-cut crystal, is the most frequently used form of monochromator produced for modern X-ray laboratory diffractometers and beamlines at synchrotron-radiation sources (Figs. 4.2.5.2[link], 4.2.5.5[link]).

The use of symmetrical and asymmetrical Bragg reflections for the production of highly collimated monochromatic beams has been discussed by Beaumont & Hart (1973[link]). This paper contains descriptions of the configurations of channel-cut monochromators and combinations of channel-cut monochromators used in modern laboratory diffractometers produced by Philips, Siemens, and Bede Scientific. In another paper, Hart (1971[link]) discussed the whole gamut of Bragg reflecting X-ray optical devices. Hart & Rodriguez (1978[link]) extended this to include a class of device in which the second wafer of the channel-cut monochromator could be tilted with respect to the first (Fig. 4.2.5.6[link] ), thereby providing an offset of the crystal rocking curves with the consequent removal of most of the contaminant harmonic radiation (Fig. 4.2.5.4[link]). The version of monochromator shown here is designed to provide thermal stability for high incident-photon fluxes. Berman & Hart (1991[link]) have also devised a class of adaptive X-ray monochromators to be used at high thermal loads where thermal expansion can cause a significant degradation of the rocking curve, and therefore a significant loss of flux and spectral purity. The cooling of Bragg-geometry monochromators at high photon fluxes presents a difficult problem in design.

[Figure 4.2.5.6]

Figure 4.2.5.6| top | pdf |

A schematic diagram of a Hart-type tuneable channel-cut monochromator is shown. The monochromator is cut from a single piece of silicon. The reflecting surfaces lie parallel to the (111) planes. Cuts are made in the crystal block so as to form a lazy hinge, and the second wafer of the monochromator is able to be deflected by a force generated by a current in an electromagnet acting on an iron disc glued to the upper surface of the wafer. Cooling of the primary crystal of the monochromator is by a jet of water falling on the underside of the wafer. This type of system can tolerate incident-beam powers of 500 W mm−2 without significant change to the width of the reflectivity curve.

Kikuta & Kohra (1970[link]), Matsushita, Kikuta & Kohra (1971[link]) and Kikuta (1971[link]) have discussed in some detail the performance of asymmetrical channel-cut monochromators. These find application under circumstances in which beam widths need to be condensed or expanded in X-ray tomography or for micro X-ray fluorescence spectroscopy. Hashizume (1983[link]) has described the design of asymmetrical monolithic crystal monochromators for the elimination of harmonics from synchrotron-radiation beams.

Many installations use a system designed by the Kohzu Company as their primary monochromator. This is a separated element design in which the reference crystal is set on the axis of the monochromator and the first crystal is set so as to satisfy the Bragg condition in both elements. One element can be tilted slightly to reduce harmonic contamination. When the wavelength is changed (i.e. θ is changed), the position of the first wafer is changed either by mechanical linkages or by electronic positioning devices so as to maintain the position of the outgoing beam in the same place as it was initially. This design of a fixed-height, separated-element monochromator was due initially to Matsushita, Ishikawa & Oyanagi (1986[link]). More recent designs incorporate liquid-nitrogen cooling of the first crystal for use with high-power insertion devices at synchrotron-radiation sources. In many installations, the second crystal can be bent into a cylindrical shape to focus the beam in the horizontal plane. The design of such a sagittally focusing monochromator is discussed by Stephens, Eng & Tse (1992[link]). Creagh & Garrett (1995[link]) have described the properties of a monochromator based on a primary monochromator (Berman & Hart, 1991[link]) and a sagittally focusing second monochromator at the Australian National Beamline at the Photon Factory.

A recent innovation in X-ray optics has been made at the European Synchrotron Radiation Facility by the group led by Snigirev (1994[link]). This combines Bragg reflection of X-rays from a silicon crystal with Fresnel reflection from a linear zone-plate structure lithographically etched on its surface. Hanfland et al. (1994[link]) have reported the use of this class of reflecting optics for the focusing of 25 to 30 keV photon beams for high-pressure crystallography experiments (Fig. 4.2.5.7[link] ).

[Figure 4.2.5.7]

Figure 4.2.5.7| top | pdf |

A schematic diagram of the use of a Bragg–Fresnel lens to focus hard X-rays onto a high-pressure cell. The diameter of the sample in such a cell is typically 10 μm. The insert shows a scanning electron micrograph of the surface of the Bragg–Fresnel lens.

Further discussion on these monochromators is to be found in this volume in Subsection 2.2.7.2[link] , §2.3.5.4.1[link] , Chapter 2.7[link] , and Section 7.4.2[link] .

4.2.5.4.4. Polarization

| top | pdf |

All scattering of X-rays by atoms causes a probable change of polarization in the beam. Jennings (1981[link]) has discussed the effects of monochromators on the polarization state for conventional diffractometers of that era. For accurate Rietveld modelling or accurate charge-density studies, the theoretical scattered intensity must be known. This is not a problem at synchrotron-radiation sources, where the incident beam is initially almost completely linearly polarized in the plane of the orbit, and is subsequently made more linearly polarized through Bragg reflection in the monochromator systems. Rather, it is a problem in the laboratory-based systems where the source is in general a source of elliptical polarization. It is essential to determine the polarization for the particular monochromator and the source combined to determine the correct form of the polarization factor to use in the formulae used to calculate scattered intensity (Chapter 6.2[link] ).

4.2.6. X-ray dispersion corrections

| top | pdf |
D. C. Creaghb

The term `anomalous dispersion' is often used in the literature. It has been dropped here because there is nothing `anomalous' about these corrections. In fact, the scattering is totally predictable.

For many years after the theoretical prediction of the dispersion of X-rays by Waller (1928[link]), and the application of this theory to the case of hydrogen-like atoms by Hönl (1933a[link],b[link]), no real use was made by experimentalists of dispersion-correction effects in X-ray scattering experiments. The suggestion by Bijvoet, Peerdeman & Van Bommel (1951[link]) that dispersion effects might be used to resolve the phase problem in the solution of crystal structures stimulated interest in the practical usefulness of this hitherto neglected aspect of the scattering of photons by atoms. In one of the first texts to discuss the problem, James (1955[link]) collated experimental data and discussed both the classical and the non-relativistic theories of the anomalous scattering of X-rays. James's text remained the principal reference work until 1974, when an Inter-Congress Conference of the International Union of Crystallography dedicated to the discussion of the topic produced its proceedings (Ramaseshan & Abrahams, 1975[link]).

At that conference, reference was made to a theoretical data set calculated by Cromer & Liberman (1970[link]) using relativistic quantum mechanics. This data set was later used in IT IV (1974[link]) and has been used extensively by crystallographers for more than a decade.

The rapid development in computing techniques, improvements in materials of construction and experimental equipment, and the use of synchrotron-radiation sources for X-ray scattering experiments have led to the production of a number of reviews of both the theoretical and the experimental aspects of the anomalous scattering of X-rays. Review articles by Gavrila (1981[link]), Kissel, Pratt, Kane & Roy (1985[link]), and Creagh (1985[link]) discuss both the theoretical and the experimental techniques for the determination of the X-ray dispersion corrections. Creagh (1986[link]) has discussed the use of X-ray anomalous scattering for the characterization of materials, and a review by Helliwell (1984[link]) has described the anomalous scattering by atoms contained in proteins and its use for the solution of the structure of proteins. In a number of papers, Karle (1980[link], 1984a[link],b[link],c[link], 1985[link]) has recently shown how powerful dispersion techniques can be in the solution of crystal structures. Indeed, the high intensity afforded by synchrotron-radiation sources, together with improvements in specimen-handling techniques, has led to the general use of dispersion techniques for the solution of the phase problem in crystal structures. In particular, the MAD (multiple-wavelength anomalous-dispersion) technique is used extensively for the solution of such macromolecular crystal structures as proteins and the like. The origin of the technique lies in the Bijvoet relations, but the implementation and the development of the technique is due to Hendrickson (1994[link]).

In this section, a brief discussion of the physical principles underlying the theoretical tabulations of X-ray scattering will be given. This will be followed by a discussion of experimental techniques for the determination of the dispersion corrections. In the next section, theoretical and experimentally determined values for the dispersion corrections will be compared for a number of elements.

Currently, there is some discussion about the nature of the dispersion corrections: are they to be considered to exhibit tensor characteristics? It is clear that in all the theoretical calculations the atoms are considered to be isolated, and, therefore, if there is a tensor associated with the X-ray scattering, it must be associated with the reaction of the atoms with the polarization state of the incident radiation. Since the property of polarization of the X-ray beam is described by a first-rank tensor, it follows that the form factor must be described by a second-rank tensor (Templeton, 1994[link]). Either the detection system of the experimental equipment must be capable of resolving the change in the polarization states of the incident and scattered radiation, or the incident radiation must be plane polarized for this property to be observable. Except for certain diffractometers at synchrotron-radiation sources, or for specially designed conventional laboratory equipment, it is not possible to determine the polarization states before and after scattering.

To a very good approximation, therefore, one can describe the form factor as being made up of a number of separate components, the largest of which is a zeroth-rank tensor that corresponds to the conventionally accepted description of the form factor. The magnitudes of any of the higher-order tensor components are small compared with this term. Whether or not they are observable depends on the characteristics of the diffraction system used in the experiment. The form-factor formalism in its zeroth-order mode has been used extremely successfully for the solution of crystal structures, the description of wavefields in crystals, the determination of the distribution of electron density in crystal structures, etc. as has been shown by Creagh (1993[link]).

It must be stressed that all of the crystal structures solved so far have been solved using the conventionally accepted, zeroth-order, form-factor description of X-ray scattering. As well, all the data concerning the distribution of electron density within crystals have used this description.

Further discussion of this issue will be given in §4.2.6.3.3.4[link].

4.2.6.1. Definitions

| top | pdf |

4.2.6.1.1. Rayleigh scattering

| top | pdf |

When photons interact with atoms, a number of different scattering processes can occur. The dominant scattering mechanisms are: elastic scattering from the bound electrons (Rayleigh scattering); elastic scattering from the nucleus (nuclear Thomson scattering); virtual pair production in the field of the screened nucleus (Delbrück scattering); and inelastic scattering from the bound electrons (Compton scattering).

Of the elastic scattering processes, only Rayleigh scattering has a significant amplitude in the range of photon energies used by crystallographers ([\lt] 100 keV). Compton scattering will be discussed elsewhere (Section 4.2.4[link]).

The essential feature of Rayleigh scattering is that the internal energy of the atom remains unchanged in the interaction. The momentum [\hbar{\bf k}_i] and polarization [{\boldvarepsilon}_i] of the incident photon may be modified during the process to [\hbar{\bf k}_f] and [{\boldvarepsilon}_f] [(\hbar{\bf k}_i,{\boldvarepsilon}_i)+A\rightarrow A+(\hbar{\bf k}_f,{\boldvarepsilon}_f).]

4.2.6.1.2. Thomson scattering by a free electron

| top | pdf |

From classical electromagnetic theory, it can be shown that the fraction of incident intensity scattered by a free electron is, at a position r, [\varphi] from the scattering electron, [I/I_0=(r_e/r){}^2\textstyle{1\over2}(1+\cos^2\varphi), \eqno (4.2.6.1)]where [r_e] is the classical radius of the electron (= 2.817938 × 10−15 m). The factor [\textstyle{1\over2}(1+\cos^2\varphi)] arises from the assumption that the electromagnetic wave is initially unpolarized. Should the wave be polarized, the factor is necessarily different from that given in equation (4.2.6.1)[link].

Equation (4.2.6.1)[link] is the basis on which the scattering power of ensembles of electrons is compared.

4.2.6.1.3. Elastic scattering from electrons bound to atoms: the atomic scattering factor, the atomic form factor, and the dispersion corrections

| top | pdf |

In considering the interaction of a photon with electrons bound in an atom, one assumes that each electron scatters independently of its fellows, and that the total scattering power of the atom is the sum of the contributions from all the electrons in the atom. Assuming that one can define an electron density ρ(r) for an atom containing a single electron, one can show that the scattering power of that atom relative to the scattering power of a Thomson free electron is [f({\boldDelta})=\int\rho({\bf r})\exp[i({\bf k}_f-{\bf k}_i)\cdot{\bf r}]\,{\rm d}V, \eqno (4.2.6.2)]where [\eqalignno{{\boldDelta}&={\bf k}_f-{\bf k}_i\equiv\hbox{ change in photon momentum}\cr &=2|k|\sin(\theta/2),}]θ being the total angle of scattering of the photon.

The scattering power for the atom relative to a free electron is referred to as the atomic form factor or the atomic scattering factor of the atom.

The result (4.2.6.2)[link], which was derived using purely classical arguments, has been shown by Nelms & Oppenheimer (1955[link]) to be identical to the result gained by quantum mechanics. If it is assumed that the atom has spherical symmetry, [f(\Delta)=4\pi{\int\limits^\infty_0}\rho(r){\textstyle\sin\Delta r\over\textstyle\Delta r}r^2\,{\rm d}r. \eqno (4.2.6.3)]For an atom containing Z electrons, the atomic form factor becomes [f(\Delta)=4\pi{\sum^{n=Z}_{n=1}\int\limits^\infty_0}\rho_n(r){\textstyle\sin\Delta r\over\textstyle\Delta r}r^2\,{\rm d}r. \eqno (4.2.6.4)]Exact solutions for the form factor are difficult to obtain, and therefore approximations have to be made to enable equation (4.2.6.4)[link] to be evaluated. The two most commonly used approximations are the Thomas–Fermi (Thomas, 1927[link]; Fermi, 1928[link]) and the Hartree–Fock (Hartree, 1928[link]; Fock, 1930[link]) techniques.

In the Thomas–Fermi model, the atomic electrons are considered to be a degenerate gas obeying Fermi–Dirac statistics and the Pauli exclusion principle, the ground-state energy of the atom being equal to the zero-point energy of this gas. The average charge density can be written in terms of the radial potential function, V(r), which may then be substituted into Poisson's equation, [\nabla^2V(r)=\rho(r)/\varepsilon_0], which can then be solved for V(r) using the boundary conditions that [\lim_{r\rightarrow\infty}V(r)=0] and that [\lim_{r\rightarrow0}rV(r)=Ze].

The Thomas–Fermi charge distributions for different atoms are related to each other. If the form factor is known for a `standard' atom for which the atomic number is [Z_0] then, for an atom with atomic number Z, [f_Z({\boldDelta})=(Z/Z_0)\,f_0({\boldDelta}'). \eqno (4.2.6.5)]Here, [\Delta'=\Delta(Z/Z_0)^{1/3}.]

The most accurate calculations of wavefunctions of many-electron atoms have been made using the self-consistent-field (Hartree–Fock) method. In this independent-particle model, each electron is assumed to move in the field of the nucleus and in an average field due to the other electrons. With this approach, the charge distribution can be written as [\rho(r)=\textstyle\sum\limits^{n=Z}_{n=1}\rho_n(r)=\sum\limits^{n=Z}_{n=1}\Psi_n^*(r)\Psi_n(r), \eqno (4.2.6.6)]where [\rho_n(r)] is the charge-density distribution of the nth electron and [\Psi_n(r)] is its wavefunction. The technique has been extended to include the effects of both exchange and correlation. Tables of relativistic Hartree–Fock values have been given by Cromer & Waber (1974[link]). Their notation F(x, Z) is related to the notation used earlier as follows: [f_Z(\Delta)\equiv F(x,Z),]where [x={|k|\over2\pi}\sin(\varphi/2)={\Delta\over4\pi}.]

In the foregoing discussion, the fact that the electrons occupy definite energy levels within the atoms has been ignored: it has been assumed that the energy of the photon is very different from any of these energy levels. The theory for calculating the scattering power of an atom near a resonant energy level was supposed to be effectively the same as the well understood problem of the driven damped pendulum system. In this type of problem, the natural amplitude of the system was modified by a correction term (a real number) dependent on the proximity of the impressed frequency to the natural resonant frequency of the system and a loss term (an imaginary number) that was related to the damping factor for the resonant system. Thus the scattering power came to be written in the form [f=f_0+\Delta f'+i\Delta f'', \eqno (4.2.6.7)]where [f_0] is the atomic scattering factor remote from the resonant energy levels, [\Delta f'] is the real part of the anomalous-scattering factor, and [\Delta f''] is the imaginary part of the anomalous-scattering factor. The nomenclature of (4.2.6.7)[link] has been superseded, but one still encounters it occasionally in modern papers.

In what follows, a brief exposition of the various theories for the anomalous scattering of X-rays and descriptions of modern experimental techniques for their determination will be given. Comparisons will be made between the several theoretical and experimental results for a number of atomic species. From these comparisons, conclusions will be drawn as to the validity of the various theories and the relevance of certain experiments.

4.2.6.2. Theoretical approaches for the calculation of the dispersion corrections

| top | pdf |

All the theories that will be discussed here have the following assumptions in common: the elastic scattering is from an isolated neutral atom and that atom is spherically symmetrical. All but the most recent of the theoretical approaches neglect changes in polarization of the incident photon caused by the interaction of the photon with the atom. In the event, few experimental configurations are able to detect such changes in polarization, and the only observable for most experiments is the momentum change of the photon.

4.2.6.2.1. The classical approach

| top | pdf |

In the classical approach, electrons are thought of as occupying energy levels within the atom characterized by an angular frequency [\omega_n] and a damping factor [\kappa_n]. The forced vibration of an electron gives rise to a dipolar radiation field, when the atomic scattering factor can be shown to be [f={\omega^2\over\omega^2-\omega^2_n-i\kappa_n\omega}. \eqno (4.2.6.8)]If the probability that the electron is to be found in the nth orbit is [g_n], the real part of the atomic scattering factor may be written as [{\rm Re}(\,f)=\sum_ng_n+\sum_n\displaystyle{g_n\omega^2_n\over \omega^2-\omega^2_n}.\eqno (4.2.6.9)]The probability [ g_n] is referred to as the oscillator strength corresponding to the virtual oscillator having natural frequency [\omega_n]. Equation (4.2.6.9) may be written as[{\rm Re}(f) = f_0 + f^\prime, \eqno (4.2.6.10)]where [f_0] represents the sum of all the elements of the set of oscillator strengths and is unity for a single-electron atom. The second term may be written as [f'={\int\limits^\infty_{\omega_{\kappa_i}}}{\omega'^2({\rm d}g_\kappa/{\rm d}\omega')\over\omega^2-\omega'^2}\,{\rm d}\omega' \eqno (4.2.6.11)]if the atom is assumed to have an infinite number of energy states. For an atom containing κ electrons, it is assumed that the overall value of f′ is the coherent sum of the contribution of each individual electron, whence [f'={\sum_\kappa\int\limits^\infty_{\omega_\kappa}}{\omega^{\prime 2}({\rm d}g_\kappa/{\rm d}\omega')\over\omega^2-\omega^{\prime 2}}\,{\rm d}\omega' \eqno (4.2.6.12)]and the oscillator strength of the κth electron [g_\kappa={\int\limits^\infty_{\omega_\kappa} }\bigg[{{\rm d}g\over{\rm d}\omega}\bigg]_\kappa\,{\rm d}\omega]is not unity, but the total oscillator strength for the atom must be equal to the total number of electrons in the atom.

The imaginary part of the dispersion correction f′′ is associated with the damping of the incident wave by the bound electrons. It is therefore functionally related to the linear absorption coefficient, [\mu_l], which can be determined from experimental measurement of the decrease in intensity of the photon beam as it passes through a medium containing the atoms under investigation. It can be shown that the attenuation coefficient per atom [\mu_a] is related to the density of the oscillator states by [\mu_a={2\pi^2e^2\over\varepsilon_0mc}\bigg[{{\rm d}g\over{\rm d}\omega}\bigg], \eqno (4.2.6.13)]whence [f''={\pi\over2}\omega\bigg[{{\rm d}g_\kappa\over{\rm d}\omega}\bigg]. \eqno (4.2.6.14)]

An expression linking the real and imaginary parts of the dispersion corrections can now be written: [f'={2\over\pi}\sum_\kappa P\int\limits^\infty_{\omega_\kappa}\displaystyle{\omega'f''(\omega',0)\over\omega^2-\omega'^2}\,{\rm d}\omega'.\eqno (4.2.6.15)]This is referred to as the Kramers–Kronig transform. Note that the term involving the restoring force has been omitted from this equation.

Equations (4.2.6.12)[link], (4.2.6.14)[link], and (4.2.6.15)[link] are the fundamental equations of the classical theory of photon scattering, and it is to these equations that the predictions of other theories are compared.

4.2.6.2.2. Non-relativistic theories

| top | pdf |

The matrix element for Rayleigh scattering from an atom having a radially symmetric charge distribution may be written as [M=M_1({\boldvarepsilon}_i\cdot{\boldvarepsilon}^*_f)+M_2({\boldvarepsilon}_i\cdot{\boldkappa }_f)({\boldvarepsilon}^*_f\cdot{\boldkappa}_i),\eqno (4.2.6.16)]where [{\boldvarepsilon}_i] and [{\boldvarepsilon}_f] represent the initial and final states of the photon. The matrix element [M_1] represents scattering for polarizations [{\boldvarepsilon}_i] and [{\boldvarepsilon}_f] perpendicular to the plane of scattering and [M_2] represents scattering for polarization states lying in the plane of scattering.

Averaged over polarization states, the differential scattering cross section takes the form [{{\rm d}\sigma\over {\rm d}\Omega}={ r^2_e\over2}(|M_1|{}^2+|M_2|{}^2).\eqno (4.2.6.17)]Here σ is the photoelectric scattering cross section, which is related to the mass attenuation coefficient [\mu_m] by [\sigma=(M/N_A)\mu_m\times10^{-24},\eqno (4.2.6.18)]where M is the molecular weight and [N_A] is Avogadro's number.

Using the vector potential of the wavefield A, an expression for the perturbed Hamiltonian of a hydrogen-like atom coupled to the radiation field may be written as [\hat{\scr H}=\hat{\scr H}_0 - r_ec{\bf A}\cdot{\bf P}+{r^2_e\over 2}{\bf A}^2,\eqno (4.2.6.19)]where [\hat{\scr H}_0] is the Hamiltonian for the unperturbed atom and [{\bf P}=i\hbar\nabla].

After application of the second-order perturbation theory, the matrix element may be deduced to be [M=({\boldvarepsilon}_i\cdot{\boldvarepsilon}_f^*)f_0(\Delta)+{1\over m}\langle1|T_1|1\rangle+{1\over m}\langle1|T_2|1\rangle. \eqno (4.2.6.20)]In this equation, the initial and final wavefunctions are designated as [\langle1|] and [|1\rangle], respectively, and the terms [T_1] and [T_2] are given by [T_1={\boldvarepsilon}_f\cdot{\bf P}\exp(-i{\bf k}_f\cdot {\bf r}){1\over E_1-{\scr H}_0+\hbar\omega+i\xi}{\boldvarepsilon}_i\cdot{\bf P}\exp(-i{\bf k}_i\cdot{\bf r})]and [T_2={\boldvarepsilon}_i\cdot{\bf P}\exp(-i{\bf k}_i\cdot {\bf r}){1\over E_1-{\scr H}_0+\hbar\omega+i\xi}{\boldvarepsilon}_f\cdot{\bf P}\exp(-i{\bf k}_f\cdot{\bf r})]where ξ is an infinitesimal positive quantity.

The first term of equation (4.2.6.20)[link] corresponds to the atomic scattering factor and is identical to the value given by classical theory. The terms involving [T_1] and [T_2] correspond to the dispersion corrections. Equation (4.2.6.20)[link] contains no terms to account for radiation damping. More complete theories take the effect of the finite width of the radiating level into account.

It is necessary to realize that the atomic scattering factor depends on both the photon's frequency [\omega] and the momentum vector [{\boldDelta}]. To emphasize this dependence, equation (4.2.6.7)[link] is rewritten as [f(\omega,{\boldDelta})=f_0({\boldDelta})+f'(\omega,{\boldDelta})+if''(\omega,{\boldDelta}). \eqno (4.2.6.21)]In the dipole approximation, it can be shown that [f'(\omega,0)={\textstyle2\over\pi}P\int\limits^\infty_0\displaystyle{\omega'f''(\omega',0)\over \omega^2-\omega'^2}\,{\rm d}\omega', \eqno (4.2.6.22)]which may be compared with equation (4.2.6.15)[link] and [f''(\omega,0)={\omega\over4\pi r_ec}\sigma(\omega), \eqno (4.2.6.23)]which may be compared with equation (4.2.6.14)[link].

There is a direct correspondence between the predictions of the classical theory and the theory using second-order perturbation theory and non-relativistic quantum mechanics.

The extension of Hönl's (1933a[link],b[link]) study of the scattering of X-rays by the K shell of atoms to other electron shells has been presented by Wagenfeld (1975[link]).

In these calculations, the energy of the photon was assumed to be such that relativistic effects do not occur, nor do transitions within the discrete states of the atom occur. Transitions to continuum states do occur, and, using the analytical expressions for the wavefunctions of the hydrogen-like atom, analytical expressions may be developed for the photoelectric scattering cross sections. By expansion of the retardation factor [\exp(-i{\bf k}\cdot{\bf r})] as the power series [1-i{\bf k}\cdot{\bf r}-{1\over2}({\bf k}\cdot{\bf r}){}^2+\ldots], it is possible to determine dipolar, quadrupolar, and higher-order terms in the analytical expression for the photoelectric scattering cross section.

The values of the cross section so obtained were used to calculate the values of [f'(\omega,{\boldDelta})] using the Kramers–Kronig transform [equation (4.2.6.22)[link]] and [f''(\omega,{\boldDelta})] using equation (4.2.6.23)[link]. The work of Wagenfeld (1975[link]) predicts that the values of [f'(\omega,{\boldDelta})] and [f''(\omega,{\boldDelta})] are functions of [{\boldDelta}]. Whether or not this is a correct prediction will be discussed in Subsection 4.2.6.3[link].

Wang & Pratt (1983[link]) have drawn attention to the importance of bound–bound transitions in the dispersion relation for the calculation of forward-scattering amplitudes. Their inclusion is especially important for elements with small atomic numbers. In a later paper, Wang (1986[link]) has shown that, for silicon at the wavelengths of Mo Kα and Ag Kα1, values for [f'(\omega,0)] of 0.084 and 0.055, respectively, are obtained. These values should be compared with those listed in Table 4.2.6.4[link].

4.2.6.2.3. Relativistic theories

| top | pdf |

4.2.6.2.3.1. Cromer and Liberman: relativistic dipole approach

| top | pdf |

It is necessary to consider relativistic effects for atoms having all but the smallest atomic numbers. Cromer & Liberman (1970[link]) produced a set of tables based on a relativistic approach to the scattering of photons by isolated atoms that was later reproduced in IT IV (1974[link]). Subsequent experimental determinations drew attention to inaccuracies in these tables in the neighbourhood of absorption edges owing to the poor convergence of the Gaussian integration technique, which was used to evaluate the real part of the dispersion correction. In a later paper, Cromer & Liberman (1981[link]) recalculated 34 instances for which the incident radiation lay close to the absorption edges of atoms using a modified integration procedure. Care should be exercised when using the Cromer & Liberman computer program, especially for calculations of [f'(\omega,0)] for high atomic weight elements at low photon energies. As Creagh (1990[link]) and Chantler (1994[link]) have shown, incorrect values of [f'(\omega,0)] can be calculated because an insufficient number of values of [f''(\omega,0)] are calculated prior to performing the Kramers–Kronig transform. In a new tabulation, Chantler (1995[link]) presents the Cromer & Liberman data using a finer integrating grid. It should be noted that the relativistic correction is the same as that used in this tabulation.

These relativistic calculations are based on the scattering formula developed by Akhiezer & Berestetsky (1957[link]) for the scattering amplitude for photons by a bound electron, viz: [S_{i\rightarrow f}=-2\pi i\delta(\varepsilon_1+\hbar\omega_1-\varepsilon_2-\hbar\omega_2)\, \displaystyle{ 4\pi(e\hbar c){}^2\over 2mc^2\hbar(\omega_1\omega_2){}{^{1/2}}} f. \eqno (4.2.6.24)]Here the angular frequencies of the incident and scattered photons are [\omega_1] and [\omega_2], respectively, and the initial and final energy states of the atom are [\varepsilon_1] and [\varepsilon_2], respectively. The scattering factor f is a complicated expression that includes the initial and final polarization states of the photon, the Dirac velocity operator, and the phase factors [\exp(i{\bf k}_1\cdot{\bf r})] and [\exp(i{\bf k}_2\cdot{\bf r})] for the incident and scattered waves, respectively. Summation is over all positive and negative intermediate states except those positive energy states occupied by other atomic electrons. The form of this expression is not easily related to the form-factor formalism that is most widely used by crystallographers, and a number of manipulations of the formula for the scattering factor are necessary to relate it more directly to the crystallographic formalism. In doing so, a number of assumptions and simplifications were made. Cromer & Liberman restricted their study to coherent, forward scatter in which changes in photon polarization did not occur. With these approximations, and using the electrical dipole approximation [[\exp(i{\bf k}\cdot{\bf r})=1]], they were able to show that [f(\omega,0)=f(0)+f^+(\omega,0)+{\textstyle{5\over3}}{ E_{\rm tot}\over mc^2}+if''(\omega,0).\eqno (4.2.6.25)]In equation (4.2.6.25)[link], f(0) is the atomic form factor for the case of forward scatter [({\boldDelta}=0)], and the term [[+{5\over3}(E_{\rm tot}/mc^2)]] arises from the application of the dipole approximation to determine the contribution of bound electrons to the scattering process. The term [f''(\omega,0)] is related to the photoelectric scattering cross section expressed as a function of photon energy [\sigma(\hbar\omega)] by [f''(\omega,0)={mc \over 4\pi\hbar e^2}\hbar\omega\,\sigma(\hbar\omega) \eqno (4.2.6.26)]and [f^+(\omega,0)=\bigg({1\over 2\pi^2\hbar r_ec}\bigg)P{\int\limits^\infty_{mc^2}}{(\varepsilon^+-\varepsilon_1)\sigma(\varepsilon+-\varepsilon_1)\over (\hbar\omega)^2-(\varepsilon^+-\varepsilon_1)^2}{\,{\rm d}}\varepsilon^+.\eqno (4.2.6.27)]These equations may be compared with equations (4.2.6.23)[link] and (4.2.6.22)[link], respectively. But equation (4.2.6.25)[link] differs from equation (4.2.6.21)[link] by the term [{5\over3}(E_{\rm tot}/mc^2)], which is constant for each atomic species, and is related to the total Coulomb energy of the atom. Evidently, to keep the formalism the same, one must write[f'(\omega,0)=f^+(\omega,0)+\textstyle{5\over3}\displaystyle{ E_{\rm tot}\over mc^2}.\eqno (4.2.6.28)]In Table 4.2.6.1[link], values of [E_{\rm tot}/mc^2] are set out as a function of atomic number for elements ranging in atomic number from 3 to 98.

Table 4.2.6.1| top | pdf |
Values of Etot/mc2 listed as a function of atomic number Z

ZSymbol[E_{\rm tot}/mc^2]
3Li−0.0004
4Be−0.0006
5B−0.0012
6C−0.0018
7N−0.0030
8O−0.0042
9F−0.0054
10Ne−0.0066
11Na−0.0084
12Mg−0.0110
13Al−0.0125
14Si−0.0158
15P−0.0180
16S−0.0210
17Cl−0.0250
18Ar−0.0285
19K−0.0320
20Ca−0.0362
21Sc−0.0410
22Ti−0.0460
23V−0.0510
24Cr−0.0560
25Mn−0.0616
26Fe−0.0680
27Co−0.0740
28Ni−0.0815
29Cu−0.0878
30Zn−0.0960
31Ga−0.104
32Ge−0.114
33As−0.120
34Se−0.132
35Br−0.141
36Kr−0.150
37Rb−0.159
38Sr−0.171
39Y−0.180
40Zr−0.192
41Nb−0.204
42Mo−0.216
43Tc−0.228
44Ru−0.246
45Rh−0.258
46Pd−0.270
47Ag−0.285
48Cd−0.300
49 In−0.318
50Sn−0.330
51Sb−0.348
52Te−0.363
53I−0.384
54Xe−0.396
55Cs−0.414
56Ba−0.438
57La−0.456
58Ce−0.474
59Pr−0.492
60Nd−0.516
61Pm−0.534
62Sm−0.558
63Eu−0.582
64Gd−0.610
65Tb−0.624
66Dy−0.648
67Ho−0.672
68Er−0.696
69Tm−0.723
70Yb−0.750
71Lu−0.780
72Hf−0.804
73Ta−0.834
74W−0.864
75Re−0.900
76Os−0.919
77Ir−0.948
78Pt−0.984
79Au−1.014
80Hg−1.046
81Tl−1.080
82Pb−1.116
83Bi−1.149
84Po−1.189
85At−1.224
86Rn−1.260
87Fr−1.296
88Ra−1.332
89Ac−1.374
90Th−1.416
91Pa−1.458
92U−1.470
93Np−1.536
94Pu−1.584
95Am−1.626
96Cm−1.669
97Bk−1.716
98Cf−1.764

To develop their tables, Cromer & Liberman (1970[link]) used the Brysk & Zerby (1968[link]) computer code for the calculation of photoelectric cross sections, which was based on Dirac–Slater relativistic wavefunctions (Liberman, Waber & Cromer, 1965[link]). They employed a value for the exchange potential of 0.667[\rho({\bf r})^{1/3}] and experimental rather than computed values of the energy eigenvalues for the atoms.

The wide use of their tables by crystallographers inevitably meant that criticism of the accuracy of the tables was forthcoming on both theoretical and experimental grounds. Stibius-Jensen (1979[link]) drew attention to the fact that the use of the dipole approximation too early in the argument caused an error of [-{1\over2}Z(\hbar\omega/mc^2)^2] in the tabulated values. More recently, Cromer & Liberman (1981[link]) include this term in their calculations. Some experimental deficiencies of the tabulated values of [f'(\omega,0)] have been discussed by Cusatis & Hart (1977[link]), Hart & Siddons (1981[link]), Creagh (1980[link], 1984[link], 1985[link], 1986[link]), Deutsch & Hart (1982[link]), Dreier, Rabe, Malzfeldt & Niemann (1984[link]), Bonse & Hartmann-Lotsch (1984[link]), and Bonse & Henning (1986[link]).

In the latter two cases, the Kramers–Kronig transformation of photoelectric scattering results has been performed without taking into account the term that arises in the relativistic case for the total Coulomb energy of the atom. Although good agreement with the Cromer & Liberman tables is claimed, their failure to include this term is an implied criticism of the Cromer & Liberman tables. That this is unjustified can be seen by references to Fig. 4.2.6.2[link][link] taken from Bonse & Henning (1986[link]), which shows that their interferometer results [which measure [f'(\omega,0)] directly] and the Kramers–Kronig results differ from one another by [\sim E_{\rm tot}/mc^2] in the neighbourhood of the K-absorption edge of niobium in the compound lithium niobate.

[Figure 4.2.6.1]

Figure 4.2.6.1| top | pdf |

The relativistic correction in electrons per atom for: (a) the modified form-factor approach; (b) the relativistic multipole approach; (c) the relativistic dipole approach.

[Figure 4.2.6.2]

Figure 4.2.6.2| top | pdf |

Measured values of f′(ω, 0) at the K-edge of Nb in LiNbO3 and the Kramers–Kronig transformation of f′′(ω, 0). The curve is obtained by transformation and the points are measured by interferometry. For (a), the polarization of the incident radiation is parallel to the hexagonal c axis, and for (b) it is at right angles to the hexagonal c axis. After Bonse & Henning (1986[link]). Note that the distortion of the dispersion curve is due to X-ray absorption near-edge structure (XANES) effects (Section 4.2.4[link]).

Further theoretical objections have been made by Creagh (1984[link]) and Smith (1987[link]), who has shown that the Stibius-Jensen correction is not valid, and that, when higher-order multipolar expansions and retardation are considered, the total self-energy correction becomes [E_{\rm tot}/mc^2] rather than [{5\over3}E_{\rm tot}/mc^2]. Fig. 4.2.6.1[link] shows the variation of the self-energy correction with atomic number for the modified form factor (Creagh, 1984[link]; Smith, 1987[link]; Cromer & Liberman, 1970[link]).

For the imaginary part of the dispersion correction [f''(\omega,0)], which depends on the calculation of the photoelectric scattering cross section, better agreement is found between theoretical results and experimental data. Details of this comparison have been given elsewhere (Section 4.2.4[link]). Suffice it to say that Creagh & Hubbell (1990[link]), in reporting the results of the IUCr X-ray Attenuation Project, could find no rational basis for preferring the Scofield (1973[link]) Hartree–Fock calculations to the Cromer & Liberman (1970[link], 1981[link]) and Storm & Israel (1970[link]) Dirac–Hartree–Fock–Slater calculations.

Computer programs based on the Cromer & Liberman program (Cromer & Liberman, 1983[link]) are in use at all the major synchrotron-radiation laboratories. Many other laboratories have also acquired copies of their program. This program must be modified to remove the incorrect Stibius–Jensen correction term, and, as will be seen later, the energy term should be modified to be [E_{\rm tot}/mc^2].

4.2.6.2.3.2. The scattering matrix formalism

| top | pdf |

Kissel, Pratt & Roy (1980[link]) have developed a computer program based on the second-order S-matrix formalism suggested by Brown, Peierls & Woodward (1955[link]). Their aim was to provide a prescription for the accurate (∼1%) prediction of the total-atom Rayleigh scattering amplitudes.

Their model treats the elastic scattering as the sum of bound electron, nuclear, and Delbrück scattering cross sections, and treats the Rayleigh scattering by considering second-order, single-electron transitions from electrons bound in a relativistic, self-consistent, central potential. This potential was a Dirac–Hartree–Fock–Slater potential, and exchange was included by use of the Kohn & Sham (1965[link]) exchange model. They omitted radiative corrections.

In principle, the observables in an elastic scattering process are momentum [(\hbar{\bf k})] and polarization [{\boldvarepsilon}]. The complex polarization vectors [{\boldvarepsilon}] satisfy the conditions [{\boldvarepsilon}^*\cdot{\boldvarepsilon}=1'\semi\quad{\boldvarepsilon}\cdot{\bf k}=0. \eqno (4.2.6.29)]

In quantum mechanics, elastic scattering is described in terms of a differential scattering amplitude, M, which is related to the elastic cross section by equation (4.2.6.16)[link].

If polarization is not an observable, then the expression for the differential scattering cross section takes the form of equation (4.2.6.17)[link]. If polarization is taken into account, as may be the case when a polarizer is used on a beam scattered from a sample irradiated by the linearly polarized beam from a synchrotron-radiation source, the full equation, and not equation (4.2.6.17)[link], must be used to compute the differential scattering cross section.

The principle of causality implies that the forward-scattering amplitude [M(\omega,0)] should be analytic in the upper half of the [\omega] plane, and that the dispersion relation [{\rm Re}\ M(\omega,0)={2\omega^2\over\pi}{\int\limits^\infty_0}{{\rm Im}\ M(\omega',0)\over \omega'(\omega'^2-\omega^2)}{\rm d}\omega'\eqno (4.2.6.30)]should hold, with the consequence that [{\rm Re}\ M(\infty,0)=-{2\over\pi}{\int\limits^\infty_0}{{\rm Im}\ M(\omega',0)\over \omega'}{\rm d}\omega'.\eqno (4.2.6.31)]This may be rewritten as [M(\omega,0)-M(\infty,0)=f'(\omega,0)+if''(\omega,0),\eqno (4.2.6.32)]with the value of [f'(\omega,0)] defined by equation (4.2.6.15)[link]. Using the conservation of probability, [{\rm Im}\ M(\omega,0)={\omega\over4\pi r_ec}\sigma_{\rm tot},\eqno (4.2.6.33)]which is to be compared with equation (4.2.6.23)[link].

Starting with Furry's extension of the formalism of quantum mechanics proposed by Feynman and Dyson, the total Rayleigh amplitude may be written as [M_n= {\sum_p} \bigg[{\langle n|T_1^*|p\rangle\langle p|T_1|n\rangle\over E_n-E_p+\hbar\omega}+{\langle n|T_2|p\rangle\langle n|T_2^*|p\rangle\over E_n-E_p+\hbar\omega}\bigg], \eqno (4.2.6.34)]where [T_1={\boldalpha}\cdot{\boldepsilon}_i\cdot\exp(i{\bf k}_i\cdot{\bf r})]and [T_2={\boldalpha}\cdot{\boldepsilon}_f^*\cdot\exp(-i{\bf k}_f\cdot{\bf r}).]The [|p\rangle] are the complete set of bound and continuum states in the external field of the atomic potential. Singularities occur at all photon energies that correspond to transitions between bound [|n\rangle] and bound state [|p\rangle]. These singularities are removed if the finite widths of these states are considered, and the energies E are replaced by [iE\Gamma/2], where [\Gamma] is the total (radiative plus non-radiative) width of the state (Gavrila, 1981[link]). By use of the formalism suggested by Brown et al. (1955[link]), it is possible to reduce the numerical problems to one-dimensional radial integrals and differential equations. The required multipole expansions of [T_1] and the specification of the radial perturbed orbitals that are characterized by angular-momentum quantum numbers have been discussed by Kissel (1977[link]). Ultimately, all the angular dependence on the photon scattering angle is written in terms of the associated Legendre functions, and all the energy dependence is in terms of multipole amplitudes.

Solutions are not found for the inhomogeneous radial wave equations, and Kissel (1977[link]) expressed the solution as the linear sum of two solutions of the homogeneous equation, one of which was regular at the origin and the other regular at infinity.

Because excessive amounts of computer time are required to use these direct techniques for calculating the amplitudes from all the subshells, simpler methods are usually used for calculating outer-shell amplitudes. Kissel & Pratt (1985[link]) used estimates for outer-shell amplitudes based on the predictions of the modified form-factor approach. A tabulation of the modified relativistic form factors has been given by Schaupp, Schumacher, Smend, Rullhusen & Hubbell (1983[link]).

Because of the generality of their approach, the computer time required for the calculation of the scattering amplitudes for a particular energy is quite long, so that relatively few calculations have been made. Their approach, however, does not confine itself solely to the problem of forward scattering of photons as does the Cromer & Liberman (1970[link]) approach. Using their model, Kissel et al. (1980[link]) have been able to show that it is incorrect to assign a dependence of the dispersion corrections on the scattering vector [{\boldDelta}]. This is at variance with some established crystallographic practices, in which the dispersion corrections are accorded the same dependence on [{\boldDelta}] as [f_0({\boldDelta})], and also at variance with the predictions of Wagenfeld's (1975[link]) non-relativistic model.

4.2.6.2.4. Intercomparison of theories

| top | pdf |

A discussion of the validity of the non-relativistic dipole approximation for the calculation of forward Rayleigh scattering amplitudes has been given by Roy & Pratt (1982[link]). They compared their relativistic multipole calculations with the relativistic dipole approximation and with the non-relativistic dipole approximation for two elements, silver and lead. They concluded that a relativistic correction to the form factor of order [(Z\alpha)^2] persists in the high-energy limit, and that this constant correction accounts for much of the deviation from the non-relativistic dipole approximation at all energies above threshold. In addition, their results illustrate that cancelling occurs amongst the relativistic, retardation, and higher multipole contributions to the scattering amplitude. This implies that care must be taken in assessing where to terminate the series that describes the multipolarity of the scattering process.

In a later paper, Roy, Kissel & Pratt (1983[link]) discussed the elastic photon scattering for small momentum transfers and the validity of the form-factor theories. In this paper, which compares the relativistic modified form factor with experimental results for lead and a relativistic form factor and the tabulation by Hubbell, Veigele, Briggs, Brown, Cromer & Howerton (1975[link]), it is shown that the modified relativistic form-factor approach gives better agreement with experiment for high momentum transfers ([\lt] 104  Å−1) than the non-relativistic, form-factor theories.

Kissel et al. (1980[link]) used the S-matrix technique to calculate the real part of the forward-scattering amplitude [f'(\omega,0)] for the inert gases at the wavelength of Mo [K\alpha_1]. These values are compared with the predictions of the relativistic dipole theory (RDP) and the relativistic multipole theory (RMP) in Table 4.2.6.2[link]. In most cases, the agreement between the S matrix and the RMP theory is excellent, considering the differences in the methodology of the two sets of calculations. Table 4.2.6.3(a)[link] shows comparisons of the real part of the forward-scattering amplitude [f''(\omega,0)] calculated for the atoms aluminium, silicon, zinc, germanium, silver, samarium, tantalum and lead using the approach of Kissel et al. (1980[link]) with that of Cromer & Liberman (1970[link], 1981[link]), with tabulations by Wagenfeld (1975[link]), and with values taken from the tables in this section. Although reasonably satisfactory agreement exists between the relativistic values, large differences exist between the non-relativistic value (Wagenfeld, 1975[link]) and the relativistic values. The major difference between the relativistic values occurs because of differences in estimation of the self-consistent-field term, which is proportional to [E_{\rm tot}/mc^2]. The Cromer & Liberman (1970[link]) relativistic dipole value is [+{5\over3}(E_{\rm tot}/mc^2)], whereas the tabulation in this section uses the relativistic multipole value of [(+E_{\rm tot}/mc^2)]. This causes a vertical shift of the curve, but does not alter its shape. Should better estimates of the self-energy term be found, the correction is simply that of adding a constant to each value of [f'(\omega,0)] for each atomic species. There is a significant discrepancy between the Kissel et al. (1980[link]) result for 62Sm and the other theoretical values. This is the only major point of difference, however, and the results are better in accord with the relativistic multipole approach than with the relativistic dipole approach. Note that the relativistic multipole approach does not include the Stibius-Jensen correction, which alters the shape of the curve.

Table 4.2.6.2| top | pdf |
Comparison between the S-matrix calculations of Kissel (K) (1977[link]) and the form-factor calculations of Cromer & Liberman (C & L) (1970[link], 1981[link], 1983[link]) and Creagh & McAuley (C & M) for the noble gases and several common metals; f′(ω, 0) values are given for two frequently used photon energies

Energy (keV)ElementRDP (C & L)S matrix (K)RMP (C & M)
17.479 (Mo [\,K\alpha_1])Ne0.0210.0240.026
Ar0.1550.1700.174
Kr−0.652−0.478 −0.557
Xe−0.684 −0.416−0.428
22.613 (Ag [\,K\alpha_1])Al 0.0320.0390.041
Zn0.260 0.3230.324
Ta−0.937 −0.375−0.383
Pb −1.910 −1.034 −1.162

Table 4.2.6.3| top | pdf |
A comparison of the forward-scattering amplitudes computed using different theoretical approaches

(a) Real part. KPR (Kissel et al., 1980[link]); C & L (Cromer & Liberman, 1970[link], 1981[link]); W (Wagenfeld, 1975[link]); and C & M (this data set).

AtomRadiation[f'(\omega,0)]
KPRC & LWC & M
19701981
13AlCr [K\alpha_1]13.32013.32813.31613.37613.326
Cu [K\alpha_1]13.20913.20413.20313.23513.213
Ag [K\alpha_1]13.03913.03213.02013.07813.041
14SiCr [K\alpha_1] 14.33314.35414.44114.365
Cu [K\alpha_1] 14.24414.24214.28214.254
Ag [K\alpha_1] 14.04214.02914.07114.052
30ZnCr [K\alpha_1]29.16129.31629.314 29.383
Cu [K\alpha_1]28.36928.38828.383 28.451
Ag [K\alpha_1]30.32330.26030.232 30.324
32GeCr [K\alpha_1] 31.53831.53830.2031.614
Cu [K\alpha_1] 30.83730.83731.9230.911
Ag [K\alpha_1] 32.22832.22832.1432.302
47AgCu [K\alpha_1]47.07546.94046.936 47.131
62SmAg [K\alpha_1]58.30756.30456.299 56.676
73TaAg [K\alpha_1]72.62572.06371.994 72.617
82PbAg [K\alpha_1]80.96680.09080.012 80.832

(b) Imaginary part [f''(\omega,0)]. KPR (Kissel et al., 1980[link]); C & L (Cromer & Liberman, 1981[link]); W (Wagenfeld, 1975[link]); and C & M (this data set).

AtomRadiation[f'(\omega,0)]
KPRC & LWC & M
13AlCr [K\alpha_1]0.5140.522 0.512
Cu [K\alpha_1]0.2430.246 0.246
Ag [K\alpha_1]0.0310.031 0.031
14SiCr [K\alpha_1] 0.6940.700.692
Cu [K\alpha_1] 0.3300.330.330
Ag [K\alpha_1] 0.0430.0470.043
30ZnCr [K\alpha_1]1.3701.373 1.371
Cu [K\alpha_1]0.6780.678 0.678
Ag [K\alpha_1]0.9320.938 0.938
32GeCr [K\alpha_1] 1.7861.841.784
Cu [K\alpha_1] 0.8860.870.886
Ag [K\alpha_1] 1.1901.231.190
47AgCu [K\alpha_1]4.2424.282 4.282
62SmCu [K\alpha_1]12.1612.218 12.218
73TaAg [K\alpha_1]4.4034.399 4.399
82PbAg [K\alpha_1]6.9376.929 6.929

In §4.2.6.3.3[link], some examples are given to illustrate the extent to which predictions of these theories agree with experimental data for [f'(\omega,0)].

That there is little to choose between the different theoretical approaches where the calculation of [f''(\omega,0)] is concerned is illustrated in Table 4.2.6.3(b)[link]. In most cases, the agreement between the scattering matrix, relativistic dipole, and relativistic multipole values is within 1%. In contrast, there are some significant differences between the relativistic and the non-relativistic values of [f''(\omega,0)]. The extent of the discrepancies is greater the higher the atomic number, as one might expect from the assumptions made in the formulation of the non-relativistic model. Some detailed comparisons of theoretical and experimental data for linear attenuation coefficients [proportional to [f''(\omega,0)]] have been given by Creagh & Hubbell (1987[link]) for silicon, and for copper and carbon by Gerward (1982[link], 1983[link]). These tend to confirm the assertion that, at the 1% level of accuracy, there is little to choose between the various relativistic models for computing scattering cross sections.

Further discussion of this is given in §4.2.6.3.3[link].

4.2.6.3. Modern experimental techniques

| top | pdf |

The atomic scattering factor enters directly into expressions for such macroscopic material properties as the refractive index, n, and the linear attenuation coefficient, [\mu_l]. The refractive index depends on the dielectric susceptibility χ through [n=(1+\chi)^{1/2},\eqno (4.2.6.35)]where [\chi=-{r_e\lambda^2\over \pi}\sum_j N_j\,f_j(\omega,{\boldDelta})\eqno (4.2.6.36)]and [N_j] is the number density of atoms of type j.

The imaginary part of the dispersion correction [f''(\omega,{\boldDelta})] for the case where [{\boldDelta}=0] is related to the atomic scattering cross section through equation (4.2.6.23)[link].

Experimental techniques that measure refractive indices or X-ray attenuation coefficients to determine the dispersion corrections involve measurements for which the scattering vector, [{\boldDelta}], is zero or close to it. Data from these experiments may be compared directly with data sets such as Cromer & Liberman (1970[link], 1981[link]).

Other techniques measure the intensities of Bragg reflections from crystalline materials or the variation of intensities within one particular Laue reflection (Pendellösung). For these cases, [{\boldDelta}=g_{hkl}], the reciprocal-lattice vector for the reflection or reflections measured. These techniques can be compared only indirectly with existing relativistic tabulations, since these have been developed for the [{\boldDelta}=0] case. Data are available for elements having atomic numbers less than 20 in the non-relativistic case (Wagenfeld, 1975[link]).

The following sections will discuss some modern techniques for the measurement of dispersion corrections, and an intercomparison will be made between experimental data and theoretical calculations for a representative selection of atoms and at two extremes of photon energies: near to and remote from an absorption edge of those atoms.

4.2.6.3.1. Determination of the real part of the dispersion correction: [f'(\omega,0)]

| top | pdf |

X-ray interferometer techniques are now used extensively for the measurement of the refractive index of materials and hence [f'(\omega,0)]. All the interferometers are transmission-geometry LLL devices (Bonse & Hart, 1965b[link], 1966a[link],b[link],c[link],d[link], 1970[link]), and initially they were used to measure the X-ray refractive indices of such materials as the alkali halides, beryllium and silicon using the characteristic radiation emitted by sealed X-ray tubes. Measurements were made for such characteristic emissions as Ag [K\alpha_1], Mo [K\alpha_1], Cu [K\alpha_1] and Cr [K\alpha_1] by a variety of authors (Creagh & Hart, 1970[link]; Creagh, 1970[link]; Bonse & Hellkötter, 1969[link]; Bonse & Materlik, 1972[link]).

The ready availability of synchrotron-radiation sources led to the adaptation of the simple LLL interferometers to use this new radiation source. Bonse & Materlik (1975[link]) reported measurements at DESY, Hamburg, made with a temporary adaptation of a diffraction-beam line. Recent advances in X-ray interferometry have led to the establishment of a permanent interferometer station at DESY (Bonse, Hartmann-Lotsch & Lotsch, 1983b[link]). This, and many of the earlier interferometers invented by Bonse, makes its phase measurements by the rotation of a phase-shifting plate in the beams emanating from the first wafer of the interferometer.

In contrast, the LLL interferometer designed by Hart (1968[link]) uses the movement of the position of lattice planes in the third wafer of the interferometer relative to the standing-wave field formed by the recombination of two of the diffracted beams within the interferometer. Measurements made with and without the specimen in position enabled both the refractive index and the linear attenuation coefficient to be determined. The use of energy-dispersive detection meant that these parameters could be determined for harmonics of the fundamental frequency to which the interferometer was tuned (Cusatis & Hart, 1975[link], 1977[link]). Subsequently, measurements have been made by Siddons & Hart (1983[link]) and Hart & Siddons (1981[link]) for zirconium, niobium, nickel, and molybdenum. Hart (1985[link]) planned to provide detailed dispersion curves for a large number of elements capable of being rolled into thin foils.

Both types of interferometers have yielded data of high quality, and accuracies better than 0.2 electrons have been claimed for measurements of [f'(\omega,0)] in the neighbourhood of the K- and L-absorption edges of a number of elements. The energy window has been claimed to be as low as 0.3 eV in width. However, on the basis of the measured values, it would seem that the width of the energy window is more likely to be about 2 eV for a primary wavelength of 5 keV.

Apparently, the ångström-ruler design is the better of the two interferometer types, since the interferometer to be mounted at the EU storage ring is to be of this type (Buras & Tazzari, 1985[link]).

Interferometers of this type have the advantage of enabling direct measurements of both refractive index and linear attenuation coefficients to be made. The determination of the energy scale and the assessment of the energy bandpass of such a system are two factors that may influence the accuracy of this type of interferometer.

One of the oldest techniques for determining refractive indices derives from measurement of the deviation produced when a prism of the material under investigation is placed in the photon beam. Recently, a number of groups have used this technique to determine the X-ray refractive index, and hence [f'(\omega,0)].

Deutsch & Hart (1984a[link],b[link]) have designed a novel double-crystal transmission spectrometer for which they were able to detect to high accuracy the angular rotation of one element with respect to the other by reference to the Pendellösung maxima that are observed in the wave field of the primary wafer. In this second paper, data gained for beryllium and lithium fluoride wedges are discussed.

Several Japanese groups have used more conventional monochromator systems having Bragg-reflecting optics to determine the refractive indices of a number of materials. Hosoya, Kawamure, Hunter & Hakano (1978; cited by Bonse & Hartmann-Lotsch, 1984[link]) made determinations of [f'(\omega,0)] in the region of the K-absorption edge for copper. More recently, Ishida & Katoh (1982[link]) have described the use of a multiple-reflection diffractometer for the determination of X-ray refractive indices. Later, Katoh et al. (1985a[link],b[link]) described its use for the measurement of [f'(\omega,0)] for lithium fluoride and potassium chloride at a wavelength near that of Mo [K\alpha_1] and for germanium in the neighbourhood of its K-absorption edge.

Measurements of the linear attenuation coefficient [\mu_l] over an extended energy range can be used as a basis for the determination of the real part of the dispersion correction [f'(\omega,0)] because of the Kramers–Kronig relation, which links [f'(\omega,0)] and [f''(\omega,0)]. However, as Creagh (1980[link]) has pointed out, even if the integration can be performed accurately [implying the knowledge of [f''(\omega,0)] over several decades of photon energies and the exact energy at which the absorption edge occurs], there will still be some ambiguity in the result because there still has to be the inclusion of the appropriate relativistic correction term.

The experimental procedures that must be adopted to ensure that the linear attenuation coefficients are measured correctly have been given in Subsection 4.2.3.2[link]. One other problem that must be addressed is the accuracy to which the photon energy can be measured. Accuracy in the energy scale becomes paramount in the neighbourhood of an absorption edge where large variations in [f'(\omega,0)] occur for very small changes in photon energy [\hbar\omega].

Despite these difficulties, Creagh (1977[link], 1978[link], 1982[link]) has used the technique to determine [f'(\omega,0)] and [f''(\omega,0)] for several alkali halides and Gerward, Thuesen, Stibius-Jensen & Alstrup (1979[link]) used the technique to measure these dispersion corrections for germanium. More recently, the technique has been used by Dreier et al. (1984[link]) to determine [f'(\omega,0)] and [f''(\omega,0)] for a number of transition metals and rare-earth atoms. The experimental configuration used by them was a conventional XAFS system. Similar techniques have been used by Fuoss & Bienenstock (1981[link]) to study a variety of amorphous materials in the region of an absorption edge.

Henke et al. (1982[link]) used the Kramers–Kronig relation to compute the real part of the dispersion correction for most of the atoms in the Periodic Table, given their measured scattering cross sections. This data set was computed specifically for the soft X-ray region [(\hbar\omega\lt1.5\,{\rm keV})].

Linear attenuation coefficient measurements yield [f'(\omega,0)] directly and [f''(\omega,0)] indirectly through use of the Kramers–Kronig integral. Data from these experiments do not have the reliability of those from refractive-index measurements because of the uncertainty in knowing the correct value for the relativistic correction term.

None of the previous techniques is useful for small photon energies. These photons would experience considerable attenuation in traversing both the specimen and the experimental apparatus. For small photon energies or large atomic numbers, reflection techniques are used, the most commonly used technique being that of total external reflection. As Henke et al. (1982[link]) have shown, when reflection occurs at a smooth (vacuum–material) interface, the refractive index of the reflecting material can be written as a single complex constant, and measurement of the angle of total external reflection may be related directly to the refractive index and therefore to [f'(\omega,{\boldDelta})]. Because the X-ray refractive indices of materials are only slightly less than unity, the scattering wavevector [{\boldDelta}] is small, and the scattering angle is only a few degrees in magnitude. Assuming that there is not a strong dependence of [f'(\omega,{\boldDelta})] with [{\boldDelta}], one may consider that this technique provides an estimate of [f'(\omega,0)] for a photon energy range that cannot be surveyed using more precise techniques. A recent review of the use of reflectometers to determine [f'(\omega,0)] has been given by Lengeler (1994[link]).

4.2.6.3.2. Determination of the real part of the dispersion correction: [f'(\omega,{\boldDelta})]

| top | pdf |

This classification includes those experiments in which measurements of the geometrical structure factors [F_{hkl}] for various Bragg reflections are undertaken. Into this category fall those techniques for which the period of standing-wave fields (Pendellösung) and reflectivity of perfect crystals in Laue or Bragg reflection are measured. Also included are those techniques from which the atomic scattering factors are inferred from measurements of Bijvoet- or Friedel-pair intensity ratios for noncentrosymmetric crystal structures.

4.2.6.3.2.1. Measurements using the dynamical theory of X-ray diffraction

| top | pdf |

The development of the dynamical theory of X-ray diffraction (see, for example, Part 5[link] in IT B, 2001[link]) and recent advances in techniques for crystal growth have enabled experimentalists to determine the geometrical structure factor [F_{hkl}] for a variety of materials by measuring the spacing between minima in the internal standing wave fields within the crystal (Pendellösung).

Two classes of Pendellösung experiment exist: those for which the ratio [(\lambda/\cos\theta)] is kept constant and the thickness of the samples varies; and those for which the specimen thickness remains constant and [(\lambda/\cos\theta)] is allowed to vary.

Of the many experiments performed using the former technique, measurements by Aldred & Hart (1973a[link],b[link]) for silicon are thought to be the most accurate determinations of the atomic form factor [f(\omega,{\boldDelta})] for that material. From these data, Price, Maslen & Mair (1978[link]) were able to refine values of [f'(\omega,{\boldDelta})] for a number of photon energies. Recently, Deutsch & Hart (1985[link]) were able to extend the determination of the form factor to higher values of momentum transfer [(\hbar{\boldDelta})]. This technique requires for its success the availability of large, strain-free crystals, which limits the range of materials that can be investigated.

A number of experimentalists have attempted to measure Pendellösung fringes for parallel-sided specimens illuminated by white radiation, usually from synchrotron-radiation sources. [See, for example, Hashimoto, Kozaki & Ohkawa (1965[link]) and Aristov, Shmytko & Shulakov (1977[link]).] A technique in which the Pendellösung fringes are detected using a solid-state detector has been reported by Takama, Kobayashi & Sato (1982[link]). Using this technique, Takama and his co-workers have reported measurements for silicon (Takama, Iwasaki & Sato, 1980[link]), germanium (Takama & Sato, 1984[link]), copper (Takama & Sato, 1982[link]), and aluminium (Takama, Kobayashi & Sato, 1982[link]). A feature of this technique is that it can be used with small crystals, in contrast to the first technique in this section. However, it does not have the precision of that technique.

Another technique using the dynamical theory of X-ray diffraction determines the integrated reflectivity for a Bragg-case reflection that uses the expression for integrated reflectivity given by Zachariasen (1945[link]). Using this approach, Freund (1975[link]) determined the value of the atomic scattering factor [f(\omega,{\bf g}_{222})] for copper. Measurements of intensity are difficult to make, and this method is not capable of yielding results having the precisions of the Pendellösung techniques.

4.2.6.3.2.2. Friedel- and Bijvoet-pair techniques

| top | pdf |

The Bijvoet-pair technique (Bijvoet et al., 1951[link]) is used extensively by crystallographers to assist in the resolution of the phase problem in the solution of crystal structures. Measurements of as many as several hundred values for the diffracted intensities [I_{hkl}] for a crystal may be made. When these are analysed, the Cole & Stemple (1962[link]) observation that the ratio of the intensities scattered in the Bijvoet or Friedel pair is independent of the state of the crystal is assumed to hold. This is a necessary assumption since in a large number of structure analyses radiation damage occurs during the course of an experiment.

For simple crystal structures, Hosoya (1975[link]) has outlined a number of ways in which values of [f'(\omega,{\bf g}_{hkl})] and [f''(\omega,{\bf g}_{hkl})] may be extracted from the Friedel-pair ratios. Measurements of these corrections for atoms such as gallium, indium, arsenic and selenium have been made.

In more complicated crystal structures for which the positional parameters are known, attempts have been made to determine the anomalous-scattering corrections by least-squares-refinement techniques. Measurements of these corrections for a number of atoms have been made, inter alia, by Engel & Sturm (1975[link]), Templeton & Templeton (1978[link]), Philips, Templeton, Templeton & Hodgson (1978[link]), Templeton, Templeton, Philips & Hodgson (1980[link]), Philips & Hodgson (1985[link]), and Chapuis, Templeton & Templeton (1985[link]). There are a number of problems with this approach, not the least of which are the requirement to measure intensities accurately for a large period of time and the assumption that specimen perfection does not affect the intensity ratio. Also, factors such as crystal shape and primary and secondary extinction may adversely affect the ability to measure intensity ratios correctly. One problem that has to be addressed in this type of determination is the fact that [f'(\omega,0)] and [f''(\omega,0)] are related to one another, and cannot be refined separately.

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)].

4.2.6.4. Table of wavelengths, energies, and linewidths used in compiling the tables of the dispersion corrections

| top | pdf |

Table 4.2.6.7[link] lists the characteristic emission wavelengths that are commonly used by crystallographers in their experiments. Also included are the emission energies (since many systems use energy rather than wavelength discrimination) and the line widths (full width at half-maximum) of these lines (Agarwal, 1979[link]; Stearns, 1984[link]; Deutsch & Hart, 1984a[link],b[link]).

Table 4.2.6.7| top | pdf |
Lists of wavelengths, energies, and linewidths used in compiling the table of dispersion corrections (a) Agarwal (1979[link]); (b) Deutsch & Hart (1982[link])

RadiationWavelength (Å)Energy (keV)Linewidth (eV)
79Au [K\alpha_1]0.180195 68.803 46 (a)
74W [K\alpha_1]0.209010 59.318 43 (a)
73Ta [K\alpha_1]0.215947 57.412 42 (a)
47Ag [K\alpha_1]0.559360 22.165 7 (a)
42Mo [K\alpha_1]0.709260 17.480 4 (a)
29Cu [K\alpha_1]1.540520 8.04792 2.61 (b)
27Co [K\alpha_1]1.788965 6.9302 1.8
26Fe [K\alpha_1]1.93597 6.4040 1.6
24Cr [K\alpha_1]2.289620 5.4149 1.5
22Ti [K\alpha_1]2.748510 4.51081.4

4.2.6.5. Tables of the dispersion corrections for forward scattering, averaged polarization using the relativistic multipole approach

| top | pdf |

See Subsection 4.2.6.3[link] for comments on the accuracy of these tables. Note also that in the neighbourhood of absorption edges the values for condensed matter may be significantly different from the values in the tables due to XAFS and XANES effects. The values in Table 4.2.6.8[link] are for scattering by isolated atoms.

Table 4.2.6.8| top | pdf |
Dispersion corrections for forward scattering

Wavelength (Å)2.7485102.2896201.9359701.7889651.5405200.7092600.5593600.2159470.2090100.180195
Li f′ =0.00350.00230.00150.00130.0008−0.0003−0.0004−0.0006−0.0006−0.0006
f′′=0.00130.00080.00060.00050.00030.00010.00000.00000.00000.0000
Be f′ =0.01170.00830.00600.00520.00380.00050.0001−0.0005−0.0005−0.0005
f′′=0.00500.00330.00230.00190.00140.00020.00010.00000.00000.0000
B f′ =0.02630.01900.01400.01210.00900.00130.0004−0.0009−0.0009−0.0010
f′′=0.01390.00940.00650.00550.00390.00070.00040.00000.00000.0000
C f′ =0.04900.03640.02730.02370.01810.00330.0015−0.0012−0.0013−0.0014
f′′=0.03130.02130.01480.01250.00910.00160.00090.00010.00010.0001
N f′ =0.08070.06060.04610.04030.03110.00610.0030−0.0020−0.0020−0.0023
f′′=0.06060.04160.02930.02480.01800.00330.00190.00020.00020.0001
O f′ =0.12130.09280.07160.06300.04920.01060.0056−0.0025−0.0026−0.0030
f′′=0.10570.07310.05180.04400.03220.00600.00360.00040.00040.0003
F f′ =0.17000.13240.10370.09200.07270.01710.0096−0.0027−0.0028−0.0034
f′′=0.17100.11920.08510.07250.05340.01030.00610.00070.00070.0005
Ne f′ =0.22570.17930.14260.12730.10190.02590.0152−0.0025−0.0028−0.0037
f′′=0.26210.18370.13180.11260.08330.01640.00980.00120.00110.0008
Na f′ =0.28010.22950.18570.16700.13530.03620.0218−0.0028−0.0031−0.0044
f′′=0.38290.26990.19570.16670.12390.02490.01500.00190.00170.0012
Mg f′ =0.32990.27780.23090.20940.17190.04860.0298−0.0030−0.0034−0.0052
f′′=0.53650.38120.27650.23730.17710.03630.02200.00280.00260.0018
Al f′ =0.37600.32600.27740.25510.21300.06450.0406−0.0020−0.0026−0.0050
f′′=0.72870.52120.38070.32760.24550.05140.03130.00400.00370.0027
Si f′ =0.39210.36470.32090.29790.25410.08170.0522−0.0017−0.0025−0.0055
f′′=0.96190.69210.50810.43840.33020.07040.04310.00560.00520.0038
P f′ =0.38210.38980.35920.33880.29550.10230.0667−0.0002−0.0012−0.0050
f′′=1.24230.89840.66280.57310.43350.09420.05800.00770.00710.0052
S f′ =0.31670.38990.38480.37060.33310.12460.08260.00150.0003−0.0045
f′′=1.56651.14100.84570.73290.55670.12340.07630.01030.00960.0069
Cl f′ =0.18320.35080.39200.38920.36390.14840.09980.00320.0017−0.0042
f′′=1.93841.42221.05960.92020.70180.15850.09840.01340.01250.0091
Ar f′ =−0.06560.26090.36960.38800.38430.17430.11910.00590.0041−0.0030
f′′=2.36701.74581.30871.13880.87170.20030.12490.01740.01620.0118
K f′ =−0.50830.09140.30680.35320.38680.20090.13990.00890.0067−0.0017
f′′=2.84372.10891.58881.38651.06570.24940.15620.02190.02040.0149
Ca f′ =−1.3666−0.19870.18670.27820.36410.22620.16110.01220.0097−0.0002
f′′=3.36942.51381.90320.66481.28550.30640.19260.02730.02550.0187
Sc f′ =−5.4265−0.6935−0.01200.14740.31190.25190.18290.01590.01300.0015
f′′=4.00172.96462.25571.97741.53310.37160.23480.03380.03150.0231
Ti f′ =−2.2250−1.6394−0.3318−0.06170.21910.27760.20600.02120.01790.0047
f′′=0.52643.45382.64252.32131.80690.44570.28300.04140.03870.0284
V f′ =−1.6269−4.4818−0.8645−0.38710.06870.30050.22760.02590.02210.0070
f′′=0.63400.45753.06442.69942.10970.52940.33760.05000.04680.0344
Cr f′ =−1.2999−2.1308−1.9210−0.9524−0.16350.32090.24960.03140.02720.0101
f′′=0.75690.54683.52513.11302.44390.62360.39920.05990.05610.0413
Mn f′ =−1.0732−1.5980−3.5716−2.0793−0.52990.33680.27040.03770.03300.0139
f′′=0.89560.64790.47983.55462.80520.72830.46810.07120.06660.0492
Fe f′ =−0.8901−1.2935−2.0554−3.3307−1.13360.34630.28860.04380.03860.0173
f′′=1.05210.76200.56490.49013.19740.84440.54480.08400.07870.0582
Co f′ =−0.7307−1.0738−1.5743−2.0230−2.36530.34940.30500.05120.04540.0219
f′′=1.22720.88970.66020.57313.61430.97210.62960.09840.09210.0682
Ni f′ =−0.5921−0.9005−1.2894−1.5664−3.00290.33930.31470.05630.05000.0244
f′′=1.42401.03310.76710.66620.50911.11240.72320.11460.10740.0796
Cu f′ =−0.4430−0.7338−1.0699−1.2789−1.96460.32010.32400.06470.05790.0298
f′′=1.64271.19300.88640.77000.58881.26510.82570.13260.12420.0922
Zn f′ =−0.3524−0.6166−0.9134−1.0843−1.54910.28390.32420.07220.06480.0344
f′′=1.88611.37121.01930.88570.67781.43010.93750.15260.14300.1063
Ga f′ =−0.2524−0.4989−0.7701−0.9200−1.28460.23070.31790.08000.07210.0393
f′′=2.15181.56741.16631.01380.77631.60831.05890.17450.16360.1218
Ge f′ =−0.1549−0.3858−0.6412−0.7781−1.08850.15470.30160.08800.07960.0445
f′′=2.44451.78411.32911.15570.88551.80011.19030.19870.18630.1389
As f′ =−0.0687−0.2871−0.5260−0.6523−0.93000.04990.27580.09620.08730.0501
f′′=2.76272.01941.50691.31091.00512.00581.33140.22520.21120.1576
Se f′ =0.0052−0.1919−0.4179−0.5390−0.7943−0.09290.23670.10470.09540.0560
f′′=3.11312.27841.70271.48211.13722.22591.48310.25430.23860.1782
Br f′ =0.0592−0.1095−0.3244−0.4363−0.6763−0.29010.18110.11060.10260.0613
f′′=3.49012.55781.91401.66731.28052.45951.64520.28580.26820.2006
Kr f′ =0.1009−0.0316−0.2303−0.3390−0.5657−0.55740.10670.11800.10820.0668
f′′=3.90832.86692.14721.87131.43852.70791.81920.31970.30030.2251
Rb f′ =0.10560.0247−0.1516−0.2535−0.4688−0.93930.00680.12470.11460.0717
f′′=4.35053.19542.39602.08931.60792.96762.00250.35610.33460.2514
Sr f′ =0.12200.1037−0.0489−0.1448−0.3528−1.5307−0.11720.13210.12190.0769
f′′=4.89463.60292.70602.36141.82003.24982.20250.39640.37260.2805
Y f′ =0.06540.12630.0138−0.0720−0.2670−2.7962−0.28790.13800.12780.0819
f′′=5.41983.99643.00542.62412.02443.56672.40990.43900.41280.3112
Zr f′ =−0.03040.13380.0659−0.0066−0.1862−2.9673−0.53640.14310.13290.0863
f′′=5.98184.42263.33012.90862.24490.55972.61410.48520.45620.3443
Nb f′ =−0.16590.12110.10720.0496−0.1121−2.0727−0.82820.14710.13710.0905
f′′=6.58034.87613.67683.21332.48260.62152.84040.53420.50250.3797
Mo f′ =−0.34870.08010.13010.0904−0.0483−1.6832−1.27030.14870.13910.0934
f′′=7.20475.34844.03883.53262.73390.68573.09780.58620.55170.4177
Tc f′ =−0.6073−0.00250.13140.11640.0057−1.4390−2.00870.14960.14060.0960
f′′=7.87395.85974.43313.87993.00490.75933.34900.64240.60470.4582
Ru f′ =−0.9294−0.10910.12200.13310.0552−1.2594−5.36300.14910.14090.0981
f′′=8.59886.40694.85404.25093.29600.83633.65060.70160.66070.5014
Rh f′ =−1.3551−0.26300.08610.13050.0927−1.1178−2.52800.14450.13730.0970
f′′=9.35046.98205.29854.64323.60450.91870.59640.76390.71950.5469
Pd f′=−1.9086−0.46400.02790.11280.1215−0.9988−1.95560.13870.13270.0959
f′′=10.14417.59385.77195.06133.93371.00720.65460.83020.78220.5955
Ag f′ =−2.5003−0.7387−0.07000.06340.1306−0.8971−1.64730.12950.12510.0928
f′′=10.99168.23586.27095.50274.28201.10150.71670.90010.84840.6469
Cd f′ =−3.5070−1.1086−0.2163−0.02140.1185−0.8075−1.43960.11710.11470.0881
f′′=11.90198.91746.80175.97284.65331.20240.78320.97410.91850.7013
In f′ =−5.1325−1.5975−0.4165−0.14730.0822−0.7276−1.28430.10130.10120.0816
f′′=12.63109.62907.35946.46745.04491.31000.85421.05190.99220.7587
Sn f′ =−7.5862−2.2019−0.6686−0.30970.0259−0.6537−1.15870.08090.08390.0728
f′′=13.516810.37427.94736.98965.45911.42460.92991.13371.06970.8192
Sb f′ =−9.2145−3.0637−0.9868−0.5189−0.0562−0.5866−1.05470.05590.06190.0613
f′′=12.766111.10268.56207.53675.89461.54611.01041.21961.15120.8830
Te f′ =−11.6068−4.2407−1.4022−0.7914−0.1759−0.5308−0.97100.02160.03160.0435
f′′=−10.101311.80799.20678.11136.35311.67511.09601.30951.23660.9499
I f′ =−13.9940−5.63531.90321.1275−0.3257−0.4742−0.8919−0.0146−0.00010.0259
f′′=3.407112.61569.88528.71596.83621.81191.18681.40371.32591.0201
Xe f′ =−9.6593−8.1899−2.6313−1.5532−0.5179−0.4205−0.8200−0.0565−0.03670.0057
f′′=3.706311.740710.57769.35857.35001.95781.28381.50231.41951.0938
Cs f′ =−8.1342−10.3310−3.5831−2.1433−0.7457−0.3680−0.7527−0.1070−0.08090.0194
f′′=4.073212.855111.290210.04547.90522.11921.39161.60581.51791.1714
Ba f′ =−7.2079−11.0454−4.6472−2.7946−1.0456−0.3244−0.6940−0.1670−0.1335−0.0494
f′′=4.411010.091912.000310.70918.46172.28191.50041.71271.61941.2517
La f′ =−6.5722−12.8190−6.3557−3.6566−1.4094−0.2871−0.6411−0.2363−0.1940−0.0835
f′′=4.75873.564812.892711.43369.03762.45231.61481.82381.72501.3353
Ce f′ =−6.0641−9.3304−8.0962−4.8792−1.8482−0.2486−0.5890−0.3159−0.2633−0.1222
f′′=5.13013.843311.873412.13509.65962.63311.73581.93981.83531.4227
Pr f′ =−5.6727−7.9841−10.9279−6.7923−2.4164−0.2180−0.5424−0.4096−0.3443−0.1666
f′′=5.50914.13049.239412.865310.28202.82141.86242.05991.94961.5136
Nd f′ =−5.3510−7.1451−10.5249−8.1618−3.1807−0.1943−0.5012−0.5194−0.4389−0.2183
f′′=5.90054.42789.981411.912110.90793.01791.99502.18432.06791.6077
Pm f′ =−5.0783−6.5334−13.2062−10.0720−4.0598−0.1753−0.4626−0.6447−0.5499−0.2776
f′′=6.31444.74223.62789.232411.55233.22492.13472.31432.19061.7056
Sm f′ =−4.8443−6.0570−9.3497−10.2609−5.3236−0.1638−0.4287−0.7989−0.6734−0.3455
f′′=6.75245.07443.88399.941212.21783.44182.28152.45102.31971.8069
Eu f′ =−4.6288−5.6630−7.9854−13.5405−8.9294−0.1578−0.3977−0.9903−0.8137−0.4235
f′′=7.20355.41784.14983.655011.18573.66822.43512.58962.45261.9120
Gd f′ =−4.5094−5.3778−7.1681−9.3863−8.8380−0.1653−0.3741−1.2279−1.0234−0.5140
f′′=7.67085.77564.42803.901611.91573.90352.59542.73042.58782.0202
Tb f′ =−4.3489−5.0951−6.5583−8.0413−9.1472−0.1723−0.3496−1.5334−1.2583−0.6165
f′′=8.18826.16674.72924.16749.18914.15372.76542.87972.73102.1330
Dy f′ =−4.1616−4.8149−6.0597−7.1503−9.8046−0.1892−0.3302−1.9594−1.5632−0.7322
f′′=8.69456.55275.02804.43209.84774.40982.94043.02742.87332.2494
Ho f′ =−4.0280−4.5887−5.6628−6.5338−14.9734−0.2175−0.3168−2.6705−1.9886−0.8709
f′′=9.23026.96195.34514.71293.70464.67833.12413.17993.02182.3711
Er f′ =−3.9471−4.4106−5.3448−6.0673−9.4367−0.2586−0.3091−5.5645−2.6932−1.0386
f′′=9.79217.39105.67765.00743.93804.95763.31580.61673.16952.4949
Tm f′ =−3.9079−4.2698−5.0823−5.6969−8.0393−0.3139−0.3084−2.8957−5.6057−1.2397
f′′=10.37637.83856.02495.31514.18215.24833.51550.65690.61922.6240
Yb f′ =−3.8890−4.1523−4.8591−5.3940−7.2108−0.3850−0.3157−2.4144−2.9190−1.4909
f′′=10.97428.29696.38135.63094.43295.54863.72290.69940.65922.7538
Lu f′ =−3.9056−4.0630−4.6707−5.1360−6.6179−0.4720−0.3299−2.1535−2.4402−1.8184
f′′=11.57878.76496.74845.95744.69375.85843.93770.74360.70102.8890
Hf f′ =−4.0452−4.0564−4.4593−4.9466−6.1794−0.5830−0.3548−1.9785−2.1778−2.2909
f′′=12.25469.28327.15186.31504.97766.18524.16430.79050.74543.0246
Ta f′ =−4.0905−3.9860−4.3912−4.7389−5.7959−0.7052−0.3831−1.8534−2.0068−3.1639
f′′=12.94799.81717.56866.68505.27186.52274.39920.83920.79153.1610
W f′ =−4.1530−3.9270−4.2486−4.5529−5.4734−0.8490−0.4201−1.7565−1.8819−3.8673
f′′=13.664310.36968.00057.06885.57746.87224.64300.89050.83880.6433
Re f′ =−4.2681−3.9052−4.1390−4.4020−5.2083−1.0185−0.4693−1.6799−1.7868−2.8429
f′′=14.393110.93468.44357.46315.89237.23104.89440.94410.89070.6827
Os f′ =−4.4183−3.9016−4.0478−4.2711−4.9801−1.2165−0.5280−1.6170−1.7107−2.4688
f′′=15.155311.52518.90677.87536.22167.60305.15581.00010.94370.7238
Ir f′ =−4.5860−3.9049−3.9606−4.1463−4.7710−1.4442−0.5977−1.5648−1.6486−2.2499
f′′=15.955812.14539.39238.30746.56677.98875.42691.05890.99930.7669
Pt f′ =−4.8057−3.9435−3.8977−4.0461−4.5932−1.7033−0.6812−1.5228−1.5998−2.1036
f′′=16.787012.79109.89858.75786.92648.39055.70811.11931.05650.8116
Au f′ =−5.0625−3.9908−3.8356−3.9461−4.4197−2.0133−0.7638−1.4693−1.5404−1.9775
f′′=17.640013.455110.42029.22227.29808.80225.99781.18331.11710.8589
Hg f′ =−5.4327−4.1029−3.8228−3.8921−4.2923−2.3894−0.8801−1.4389−1.5055−1.8958
f′′=18.524114.147310.96509.70767.68499.22666.29891.24831.17960.9080
Tl f′ =−5.8163−4.2233−3.8103−3.8340−4.1627−2.8358−1.0117−1.4111−1.4740−1.8288
f′′=19.437814.864311.530010.21088.09009.66886.60901.31891.24560.9594
Pb f′ =−6.4779−4.4167−3.8519−3.8236−4.0753−3.3944−1.1676−1.3897−1.4497−1.7773
f′′=20.333615.598712.110610.72928.506010.11116.92871.39091.31371.0127
Bi f′ =−7.0419−4.6533−3.9228−3.8408−4.0111−4.1077−1.3494−1.3721−1.4290−1.7346
f′′=21.219616.344812.701711.25758.931010.25667.25661.46611.38511.0685
Po f′ =−7.7195−4.9604−4.0267−3.8855−3.9670−5.1210−1.5613−1.3584−1.4133−1.7005
f′′=22.197417.141013.332911.82099.383411.04967.59861.54431.45921.1266
At f′ =−8.5994−5.3399−4.1781−3.9706−3.9588−7.9122−1.8039−1.3540−1.4066−1.6784
f′′=23.221317.939013.970912.39159.84339.97777.95091.62601.53671.1876
Rn f′ =−10.2749−5.7275−4.3331−4.0549−3.9487−8.0659−2.0847−1.3475−1.3982−1.6571
f′′=24.261318.772014.631312.981510.318110.45808.31121.71031.61671.2504
Fr f′ =−10.8938−6.2180−4.5387−4.1818−3.9689−7.2224−2.4129−1.3404−1.3892−1.6367
f′′=24.304119.600915.301613.582510.80387.78478.68391.79861.70041.3162
Ra f′ =−12.3462−6.7502−4.7764−4.3309−4.0088−6.7704−2.8081−1.3462−1.3931−1.6299
f′′=25.537420.438915.977814.190211.29698.14359.06141.88911.78631.3840
Ac f′ =−12.3496−7.4161−5.0617−4.5270−4.0794−6.8494−3.2784−1.3473−1.3922−1.6190
f′′=25.136321.305316.668714.809611.79948.51789.45021.98451.87701.4553
Th f′ =−13.6049−8.2118−5.3692−4.7310−4.1491−7.2400−3.8533−1.3524−1.3955−1.6136
f′′=26.251122.224817.401815.464212.32968.89799.84032.08191.96951.5284
Pa f′ =−14.4639−9.4459−5.7337−4.9639−4.2473−8.0334−4.6067−1.3672−1.4083−1.6170
f′′=27.447523.154818.140616.129512.86819.280710.24132.18352.06611.6047
U f−12.3528−9.9362−6.1485−5.2392−4.3638−9.6767−5.7225−1.3792−1.4184−1.6188
f′′30.172523.123918.872816.795213.40909.664610.64282.28761.16501.6831
Np f−17.4143−11.1080−6.6136−5.5633−4.5053−11.4937−6.9995−1.3941−1.4312−1.6231
f′′31.740524.116819.637917.483713.96664.41939.58762.39582.26791.7648
Pu f−18.0862−11.4073−6.9721−5.8130−4.6563−9.4100−13.5905−1.4180−1.4527−1.6351
f′′33.896323.296020.154817.957914.37294.30566.94682.49792.36521.8430
Am f−19.7042−11.7097−7.7881−6.2920−4.8483−7.8986−6.7022−1.4359−1.4684−1.6424
f′′37.371624.571521.173818.861815.08774.51257.31082.62182.48291.9358
Cm f−24.9307−10.4100−8.6102−6.7506−5.0611−7.3248−6.2891−1.4655−1.4952−1.6592
f′′41.485225.811521.888019.511915.63554.69807.60442.74212.59742.0271
Bk f−32.8492−9.2185−9.3381−7.4293−5.3481−6.8498−6.3438−1.4932−1.5203−1.6746
f′′32.542129.302821.951420.358116.31904.90867.94772.86532.71472.1208
Cf f−23.6520−23.5202−9.7799−7.8616−5.5545−6.6561−6.4144−1.5323−1.5562−1.6984
f′′21.933431.299922.485820.853616.74285.07858.19302.98072.82502.2102

Acknowledgements

The efforts described in Section 4.2.2 owe their inception to the encouragement of the late A. J. C. Wilson, who persistently communicated the need for an updated wavelength resource for the crystallographic community. The larger effort evolved at NIST with the support of the Standard Reference Data Program as established with the help of the late Jean Gallagher, and sustained by the program's current Director, John Rumble. Early phases of the preparation of the material in this section benefited from the efforts of John Schweppe. Cedric Powell supplied valuable advice in the area of electron binding energies. RDD, EGK, PI and EL are particularly grateful to the Editor, E. Prince, for his help and patience in the development of these wavelength tables. Richard Deslattes died between the first publication of this article and this revision. This work would not have been possible without his dedication to this project over more than a decade. The earlier wavelength table of the late J. A. Bearden, under whom one of the present authors (RDD) studied, was a significant influence on this project.

References

First citation Agarwal, B. K. (1979). X-ray spectroscopy, p. 215. Berlin: Springer-Verlag.Google Scholar
First citation Akhiezer, A. I. & Berestetsky, V. B. (1957). Quantum electrodynamics. TESE, Oak Ridge, Tennessee, USA.Google Scholar
First citation Aldred, P. J. E. & Hart, M. (1973a). The electron distribution in silicon. I. Experiment. Proc. R. Soc. London Ser. A, 332, 233–238.Google Scholar
First citation Aldred, P. J. E. & Hart, M. (1973b). The electron distribution in silicon. II. Theoretical interpretation. Proc. R. Soc. London Ser. A, 332, 239–254.Google Scholar
First citation Allen, S. J. M. (1935). X-rays in theory and experiment, edited by A. H. Compton & S. K. Allison, pp. 799–806. New York: Van Nostrand.Google Scholar
First citation Allen, S. J. M. (1969). Handbook of chemistry and physics, edited by R. C. Weast, pp. E143–E144. Cleveland: The Chemical Rubber Co.Google Scholar
First citation Alvarez, L. W., Crawford, F. S. & Stevenson, M. L. (1958). Elastic scattering of 1.6 MeV gamma rays from H, Li, C and Al nuclei. Phys. Rev. 112, 1267–1273.Google Scholar
First citation Aristov, V. V., Shmytko, I. M. & Shulakov, E. V. (1977). Dynamical contrast of the topographic image of a crystal with continuous X-radiation. Acta Cryst. A33, 412–418.Google Scholar
First citation Arndt, U. W. (1992). X-ray wavelengths. International Tables for Crystallography, Vol. C, edited by A. J. C. Wilson, pp. 176–182. Dordrecht: Kluwer Academic Publishers.Google Scholar
First citation Azaroff, L. V. & Pease, D. M. (1974). X-ray absorption spectra. X-ray spectroscopy, edited by L. V. Azaroff, Chap. 6, pp. 284–337. New York: McGraw-Hill.Google Scholar
First citation Bailey, R. L. (1978). The design and operation of magnetic liquid shaft seals. In Thermomechanics of magnetic fluids, edited by B. Berkovsky. London: Hemisphere.Google Scholar
First citation Balaic, D. X., Barnea, Z., Nugent, K. A., Garrett, R. F., Varghese, J. N. & Wilkins, S. W. (1996). Protein crystal diffraction patterns using a capillary-focused synchrotron X-ray beam. J. Synchrotron Rad. 3, 289–295.Google Scholar
First citation Balaic, D. X. & Nugent, K. A. (1995). The X-ray optics of tapered capillaries. Appl. Opt. 34, 7263–7272.Google Scholar
First citation Balaic, D. X., Nugent, K. A., Barnea, Z., Garrett, R. & Wilkins, S. W. (1995). Focusing of X-rays by total external reflection from a paraboloidally tapered glass capillary. J. Synchrotron Rad. 2, 296–299.Google Scholar
First citation Band, I. M., Kharitonov, Yu. I. & Trzhaskovskaya, M. B. (1979). Photoionization cross sections and photoelectron angular distributions for X-ray line energies in the range 0.132–4.509 keV. At. Data Nucl. Data Tables, 23, 443–505.Google Scholar
First citation Barreau, G. H., Börner, H. G., Egidy, T. V. & Hoff, R. W. (1982). Precision measurement of X-ray energies, natural widths and intensities in the actinide region. Z. Phys. A308, 209–213.Google Scholar
First citation Basile, G., Bergamin, A., Cavagnero, G., Mana, G., Vittone, E. & Zosi, G. (1994). Measurement of the silicon (220) lattice spacing. Phys. Rev. Lett. 72, 3133–3136.Google Scholar
First citation Basile, G., Bergamin, A., Cavagnero, G., Mana, G., Vittone, E. & Zosi, G. (1995). The (220) lattice spacing of silicon. IEEE Trans. Instrum. Meas. 44, 526–529.Google Scholar
First citation Bearden, J. A. (1960). Lead K absorption edge for μ-meson mass determination. Phys. Rev. Lett. 4, 240–241.Google Scholar
First citation Bearden, J. A. (1965). Selection of W Kα1 as the X-ray wavelength standard. Phys. Rev. B, 137, 455–461.Google Scholar
First citation Bearden, J. A. (1967). X-ray wavelengths. Rev. Mod. Phys. 39, 78–100.Google Scholar
First citation Bearden, J. A., Henins, A., Marzolf, J. G., Sauder, W. C. & Thomsen, J. S. (1964). Precision redetermination of standard reference wavelengths for X-ray spectroscopy. Phys. Rev. A, 135, 899–910.Google Scholar
First citation Bearden, J. A., Thomsen, J. S., Burr, A. F., Yap, F. Y., Huffman, F. N., Henins, A. & Matthews, G. D. (1964). Remeasurement of selected X-ray lines and reevaluation of all published wavelengths on a consistent and absolute scale. Baltimore, Maryland: The Johns Hopkins University.Google Scholar
First citation Beaumont, J. H. & Hart, M. (1973). Multiple-Bragg reflection monochromators for synchrotron radiation. J. Phys. E, 7, 823–829.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 a 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. B48, 17–21.Google Scholar
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 Berman, L. E. & Hart, M. (1991). Adaptive crystal optics for high power synchrotron sources. Nucl. Instrum. Methods, A302, 558–562.Google Scholar
First citation Bertin, E. P. (1975). Principles and practice of X-ray spectrometric analysis. New York: Plenum.Google Scholar
First citation Beyer, H., Indelicato, P., Finlayson, H., Liesen, D. & Deslattes, R. D. (1991). Measurement of the 1s Lamb-shift in hydrogenlike nickel. Phys. Rev. A, 43, 223–227.Google Scholar
First citation Bianconi, A., Incoccia, L. & Stipcich, S. (1983). Editors. EXAFS and near edge structure. Berlin: Springer.Google Scholar
First citation Bijvoet, J. M., Peerdeman, A. F. & Van Bommel, A. J. (1951). Determination of the absolute configuration of optically active compounds by means of X-rays. Nature (London), 168, 271.Google Scholar
First citation Bilderback, D. H., Thiel, D. J., Pahl, R. & Brister, K. E. (1994). X-ray applications with glass-capillary optics. J. Synchrotron Rad. 1, 37–42.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 Blundell, S. A. (1993a). Ab initio calculations of QED effects of Li-like, Na-like and Cu-like ions. Phys. Scr. T46, 144–149.Google Scholar
First citation Blundell, S. A. (1993b). Calculations of the screened self-energy and vacuum polarization in Li-like, Na-like and Cu-like ions. Phys. Rev. A, 47, 1790–1803.Google Scholar
First citation Blundell, S. A., Johnson, W. R. & Sapirstein, J. (1990). Improved many-body perturbation theory calculations of the n = 2 states of lithiumlike uranium. Phys. Rev. A, 41, 1698–1700.Google Scholar
First citation Blundell, S. A., Mohr, P. J., Johnson, W. R. & Sapirstein, J. (1993). Evaluation of two-photon exchange graphs for highly charged heliumlike ions. Phys. Rev. A, 48, 2615–2626.Google Scholar
First citation Bonse, U. & Hart, M. (1965a). An X-ray interferometer. Appl. Phys. Lett. 6, 155–156.Google Scholar
First citation Bonse, U. & Hart, M. (1965b). Tailless X-ray single-crystal reflection curves obtained by multiple reflection. Appl. Phys. Lett. 7, 238–240.Google Scholar
First citation Bonse, U. & Hart, M. (1966a). A Laue-case X-ray interferometer. Z. Phys. 188, 154–162.Google Scholar
First citation Bonse, U. & Hart, M. (1966b). An X-ray interferometer with Bragg-case beam splitting. Z. Phys. 194, 1–17.Google Scholar
First citation Bonse, U. & Hart, M. (1966c). Moire patterns of atomic planes obtained by X-ray interferometry. Z. Phys. 190, 455–467.Google Scholar
First citation Bonse, U. & Hart, M. (1966d). Small single X-ray scattering by spherical particles. Z. Phys. 189, 151–165.Google Scholar
First citation Bonse, U. & Hart, M. (1970). Interferometry with X-rays. Phys. Today, 23(8), 26–31.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., Hartmann-Lotsch, I. & Lotsch, H. (1983b). The X-ray interferometer for high resolution measurement of anomalous dispersion at HASYLAB. Nucl. Instrum. Methods, 208, 603–604.Google Scholar
First citation Bonse, U. & Hellkötter, H. (1969). Interferometrische Messung des Brechungsindex für Röntgenstrahlen. Z. Phys. 223, 345–352.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 Bonse, U. & Materlik, G. (1975). Dispersion correction measurements by X-ray interferometry. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 107–109. Copenhagen: Munksgaard.Google Scholar
First citation Borchert, G. L. (1976). Precise energies of K-Röntgen-lines of Tm, Th, U and Pu. Z. Naturforsch. Teil A, 31, 102–104.Google Scholar
First citation Borchert, G. L., Hansen, P. G., Jonson, B., Ravn, H. L. & Desclaux, J. P. (1980). Comparison of the K X-ray energy ratios of high Z and low Z elements with relativistic SCF DF calculations. Atomic masses and fundamental constants, edited by J. E. Nolen & W. Benenson, Vol. 6, pp. 189–201. New York: Plenum Press.Google Scholar
First citation Bowen, D. K., Stock, S. R., Davies, S. T., Pantos, E., Birnbaum, H. R. & Chen, H. (1984). Topographic EXAFS. Nature (London), 309, 336–338.Google Scholar
First citation Brodsky, A. (1982). Editor. Handbook of radiation measurement and protection, Vols. 1 and 2. Florida: CRC Press.Google Scholar
First citation Brown, G. E., Peierls, R. E. & Woodward, J. B. (1955). The coherent scattering of γ-rays by K electrons in heavy atoms: method. Proc. R. Soc. London Ser. A, 227, 51–63.Google Scholar
First citation Brysk, H. & Zerby, C. D. (1968). Photoelectric cross sections in the keV range. Phys. Rev. 171, 292–298.Google Scholar
First citation Bunker, G., Hasnain, S. S. & Sayers, D. E. (1990). Report of the International Workshops on Standards and Criteria in XAFS. In X-ray absorption spectroscopy, edited by S. S. Hasnain. London: Ellis Horwood.Google Scholar
First citation Buras, B. (1985). The European Synchrotron Radiation Project. Nucl. Sci. Appl. 2, 127–143.Google Scholar
First citation Buras, B. & Marr, G. V. (1979). Editors. European Synchrotron Radiation Facility. Suppl. III: Instrumentation. Strasbourg: ESF.Google Scholar
First citation Buras, B. & Tazzari, S. (1984). Editors. European Synchrotron Radiation Facility. Geneva: ESRP c/o CERN.Google Scholar
First citation Buras, B. & Tazzari, S. (1985). The European Synchrotron Radiation Facility. Report of the ESRP, pp. 6–32. European Synchrotron Radiation Project, CERN, LEP Division, Geneva, Switzerland.Google Scholar
First citation Burr, A. F. (1996). Personal communication.Google Scholar
First citation Byer, R. L., Kuhn, K., Reed, M. & Trail, J. (1983). Progress in high peak and average power lasers for soft X-ray production. Proc. SPIE, 448, 2–7.Google Scholar
First citation Caballero, A., Villain, F., Dexpert, H., Le Peltier, F. & Lynch, J. (1993). Characterization by in situ EXAFS spectroscopy of Pt/Al2O3 and PtRe/Al2O3 catalysts under reaction conditions. Jpn. J. Appl. Phys. 32, Suppl. 32–2, 439–441.Google Scholar
First citation Cardona, M. & Ley, L. (1978). Photoemission in solids I. Top. Appl. Phys. 26, 265–276.Google Scholar
First citation Castaing, R. & Descamps, J. (1955). Sur les bases physiques de l'analyse ponctuelle par spectrographie X. J. Phys. Radium, 16, 304–317.Google Scholar
First citation Cauchois, Y. & Hulubei, H. (1947). Tables de constantes et données numeriques. I. Longueurs d'onde des emissions X et des discontinuités d'absorption X. Paris: Herman.Google Scholar
First citation Cauchois, Y. & Senemaud, C. (1978). Tables internationales de constantes selectionnées. 18. Longeurs d'onde des emissions X et des discontinuités d'absorption X. London: Pergamon Press.Google Scholar
First citation Chantler, C. T. (1994). Towards improved form factor tables. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 61–79. Amsterdam: North Holland.Google Scholar
First citation Chantler, C. T. (1995). Theoretical form factor, attenuation and scattering tabulation for Z = 1–92 from 1–10 eV to E = 0.4–1.0 MeV. J. Phys. Chem. Ref. Data, 24, 71–643.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 Chipman, D. R. (1969). Conversion of relative intensities to an absolute scale. Acta Cryst. A25, 209–214.Google Scholar
First citation Citrin, P. H., Eisenberger, P. & Hewitt, R. (1978). Extended X-ray absorption fine structure of surface atoms on single-crystal substrates: iodine absorbed on Ag(111). Phys. Rev. Lett. 41, 309–312.Google Scholar
First citation Clay, R. E. (1934). A 5 kW X-ray generator with a spinning target. Proc. Phys. Soc. London, 46, 703–712.Google Scholar
First citation Cohen, E. R. & Taylor, B. N. (1973). The 1973 least-squares adjustment of the fundamental constants. J. Phys. Chem. Ref. Data, 2, 663–734.Google Scholar
First citation Cohen, E. R. & Taylor, B. N. (1987). The 1986 adjustment of the fundamental physical constants. Rev. Mod. Phys. 89, 1121–1148.Google Scholar
First citation Cole, H. & Stemple, N. R. (1962). Effect of crystal perfection and polarity on absorption edges seen in Bragg diffraction. J. Appl. Phys. 33, 2227–2233.Google Scholar
First citation Collins, C. B., Davanloo, F. & Bowen, T. S. (1986). Flash X-ray source of intense nanosecond pulses produced at high repetition rates. Rev. Sci. Instrum. 57, 863–865.Google Scholar
First citation Compton, A. H. & Allison, S. K. (1935). X-rays in theory and experiment. New York: Van Nostrand.Google Scholar
First citation Cooper, M. J. (1985). Compton scattering and electron momentum determination. Rep. Prog. Phys. 48, 415–481.Google Scholar
First citation Cosslett, V. E. & Nixon, W. C. (1951). X-ray shadow microscope. Nature (London), 168, 24–25.Google Scholar
First citation Cosslett, V. E. & Nixon, W. C. (1960). X-ray microscopy, pp. 217–222. Cambridge University Press.Google Scholar
First citation Creagh, D. C. (1970). On the measurement of X-ray dispersion corrections using X-ray interferometers. Aust. J. Phys. 23, 99–103.Google Scholar
First citation Creagh, D. C. (1977). Determination of the mass attenuation coefficients and the anomalous dispersion corrections for calcium at X-ray wavelengths from I Kα1 to Cu Kα1. Phys. Status Solidi A, 39, 705–715.Google Scholar
First citation Creagh, D. C. (1978). A new device for the precise measurement of X-ray attenuation coefficients and dispersion corrections. Advances in X-ray analysis, Vol. 21, edited by C. S. Barrett & D. E. Leyden, pp. 149–153. New York: Plenum.Google Scholar
First citation Creagh, D. C. (1980). X-ray interferometer measurements of the anomalous dispersion correction f′(ω, 0) for some low Z elements. Phys. Lett. A, 77, 129–132.Google Scholar
First citation Creagh, D. C. (1982). On the use of photoelectric attenuation coefficients for the determination of the anomalous dispersion coefficient f′ for X-rays. Phys. Lett. A, 103, 52–56.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. (1986). The X-ray anomalous dispersion of materials. In Recent advances in X-ray characterization of materials, edited by P. Krishna, Chap. 7. Oxford: Pergamon Press.Google Scholar
First citation Creagh, D. C. (1987a). The resolution of discrepancies in tables of photon attenuation coefficients. Nucl. Instrum. Methods, A255, 1–16.Google Scholar
First citation Creagh, D. C. (1987b). The X-ray anomalous dispersion corrections and their use for the characterization of materials. In Progress in crystal growth and characterization, Vol. 14, edited by P. Krishna, Chap. 7, pp. 1–46. Oxford: Pergamon Press.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. & Garrett, R. F. (1995). Testing of a sagittal focusing monochromator at BL 20B at the Photon Factory. Access to major facilities program, edited by J. W. Boldeman, pp. 251–252. Sydney: ANSTO.Google Scholar
First citation Creagh, D. C. & Hart, M. (1970). X-ray interferometric measurements of the forward scattering amplitude of lithium fluoride. Phys. Status Solidi, 37, 753–758.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. (1969). Anomalous dispersion corrections computed from self consistent field relativistic Dirac–Slater wavefunctions. J. Chem. Phys. 50, 4857–4859.Google Scholar
First citation Cromer, D. T. & Liberman, D. (1970). Relativistic calculation of anomalous scattering factors for X-rays. J. Chem. Phys. 53, 1891–1898.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 Cromer, D. T. & Liberman, D. A. (1983). Calculation of anomalous scattering factors at arbitrary wavelengths. J. Appl. Cryst. 16, 437.Google Scholar
First citation Cromer, D. T. & Mann, J. B. (1967). Compton scattering factors for spherically symmetric free atoms. J. Chem. Phys. 47, 1892–1893.Google Scholar
First citation Cromer, D. T. & Waber, J. T. (1974). Atomic scattering factors for X-rays. International tables for X-ray crystallography, Vol. IV, edited by J. A. Ibers & W. C. Hamilton, Chap. 2.2, pp. 71–147. Birmingham: Kynoch Press.Google Scholar
First citation Crozier, E. D. & Seary, A. J. (1980). Asymmetric effects in the extended X-ray absorption fine structure. Analysis of solid and liquid zinc. Can. J. Phys. 58, 1388–1399.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 Cusatis, C. & Hart, M. (1977). The anomalous dispersion corrections for zirconium. Proc. R. Soc. London Ser. A, 354, 291–302.Google Scholar
First citation De Marco, J. J. & Suortti, P. (1971). Effect of scattering on the attenuation of X-rays. Phys. Rev. B, 4, 1028–1033.Google Scholar
First citation Delf, B. W. (1961). The effect of absorption in the β-filter on the mean wavelength of X-ray emission lines. Proc. Phys. Soc. London, 78, 305–306.Google Scholar
First citation Desclaux, J. P. (1975). A multiconfiguration relativistic Dirac–Fock program. Comput. Phys. Commun. 9, 31–45.Google Scholar
First citation Desclaux, J. P. (1993). Relativistic multiconfiguration Dirac–Fock package. Methods and techniques in computational chemistry - 94, Vol. A, edited by E. Clementi. Cagliary: STEF.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 Deslattes, R. D. & Kessler, E. G. Jr (1985). Experimental evaluation of inner-vacancy level energies for comparison with theory. Atomic inner-shell physics, edited by B. Crasemann, pp. 181–235. New York: Plenum.Google Scholar
First citation Deslattes, R. D., Tanaka, M., Greene, G. L., Henins, A. & Kessler, E. G. (1987). Remeasurement of a silicon lattice period. IEEE Trans. Instrum. Meas. IM-36, 166.Google Scholar
First citation Deutsch, M. & Hart, M. (1982). Wavelength energy shape and structure of the Cu Kα1 X-ray emission line. Phys. Rev. B, 26, 5550–5567.Google Scholar
First citation Deutsch, M. & Hart, M. (1984a). Noninterferometric measurement of the X-ray refractive index of beryllium. Phys. Rev. B, 30, 643–646.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 Deutsch, M. & Hart, M. (1985). A new approach to the measurement of X-ray structure amplitudes determined by the Pendellösung method. Acta Cryst. A41, 48–55.Google Scholar
First citation Doyle, P. A. & Turner, P. S. (1968). Relativistic Hartree–Fock X-ray and electron scattering factors. Acta Cryst. A24, 390–397.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 Durham, P. J. (1983). Multiple scattering calculations of XANES. EXAFS and near edge structure, edited by A. Bianconi, L. Incoccia & S. Stipcich, pp. 37–43. Berlin: Springer.Google Scholar
First citation Dyson, N. A. (1973). X-rays in atomic and nuclear physics. London: Longman.Google Scholar
First citation Ehrenberg, W. & Spear, W. E. (1951). An electrostatic focusing system and its application to a fine focus X-ray tube. Proc. Phys. Soc. London Sect. B, 64, 67–75.Google Scholar
First citation Engel, D. H. & Sturm, M. (1975). Experimental determination of f′′ for heavy atoms. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 93–100. Copenhagen: Munksgaard.Google Scholar
First citation Engström, P., Rindby, A. & Vincze, L. (1996). Capillary optics. ESRF Newsletter, 26, 30–31.Google Scholar
First citation Farge, Y. & Duke, P. J. (1979). Editors. European Synchrotron Radiation Facility. Suppl. 1: The scientific case. Strasbourg: ESF.Google Scholar
First citation Fermi, E. (1928). Eine statistiche Methode zur Bestimmung einiger geschaften des Atoms und ihre Anwendung auf die Theorie des periodische Systems der Elemente. Z. Phys. 48, 73–79.Google Scholar
First citation Fiorito, R. B., Rule, D. W., Piestrup, M. A., Li, Q., Ho, A. H. & Maruyama, X. K. (1993). Parametric X-ray generation from moderate-energy electron beams. Nucl. Instrum. Methods, B79, 758–761.Google Scholar
First citation Fock, V. (1930). Näherungsmethode zur Lösung des quantenmechanischen Mehrkörperproblems. Z. Phys. 61, 126–148.Google Scholar
First citation Forsyth, J. M. & Frankel, R. D. (1980). Flash X-ray diffraction from biological specimens using a laser-produced plasma source. Report No. 106. Laboratory for Laser Energetics, University of Rochester, USA.Google Scholar
First citation Forsyth, J. M. & Frankel, R. D. (1984). Experimental facility for nanosecond time-resolved low-angle X-ray diffraction experiments using a laser-produced plasma source. Rev. Sci. Instrum. 55, 1235–1242.Google Scholar
First citation Fourme, R. (1992). Sources X intenses et cristallographie biologique. Ann. Phys. (Leipzig), 17, 247–255.Google Scholar
First citation Frankel, R. D. & Forsyth, J. M. (1979). Nanosecond exposure X-ray diffraction patterns from biological specimens using a laser plasma source. Science, 204, 622–624.Google Scholar
First citation Frankel, R. D. & Forsyth, J. M. (1985). Time-resolved X-ray diffraction study of photo stimulated purple membrane. Biophys. J. 47, 387–393.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 Freund, A. K. (1993). Thin is beautiful. ESRF Newsletter, 19, 11–13.Google Scholar
First citation Fricke, H. (1920). The K-characteristic absorption frequencies for the chemical elements magnesium to chromium. Phys. Rev. 16, 202–212.Google Scholar
First citation Fuggle, J. C., Burr, A. F., Watson, L. M., Fabian, D. J. & Lang, W. (1974). X-ray photoelectron studies of thorium and uranium. J. Phys. F, 4, 335–342.Google Scholar
First citation Fujimoto, Z., Fujii, K., Tanaka, M. & Nakayama, K. (1997). Lattice measurement in silicon. IEEE Trans. Instrum. Meas. 50, 123–124.Google Scholar
First citation Fukumachi, T., Nakano, Y. & Kawamura, T. (1986). Energy dependence of X-ray reflectivity from multi-layer mirrors. X-ray instrumentation for the Photon Factory: dynamic analyses of micro-structures in matter, edited by S. Hosoya, Y. Iitaka & H. Hashizume, pp. 25–34. Japan: Photon Factory.Google Scholar
First citation Fuoss, P. H. & Bienenstock, A. (1981). X-ray anomalous scattering factors – measurements and applications. Inner-shell and X-ray physics of atoms and solids, edited by D. J. Fabian, A. Kleinpoppen & L. M. Watson, pp. 875–884. New York: Plenum.Google Scholar
First citation Fuoss, P. H., Eisenberger, P., Warburton, W. K. & Bienenstock, A. (1981). Application of differential anomalous X-ray scattering to structural studies of amorphous materials. Phys. Rev. Lett. 46, 1537–1540.Google Scholar
First citation Gavrila, M. (1981). Photon atom elastic scattering in inner-shell and X-ray physics of atoms and solids, edited by B. Crasemann. New York: Plenum.Google Scholar
First citation Gerstenberg, H. & Hubbell, J. H. (1982). Comparison of experimental with theoretical photon attenuation cross sections between 10 eV and 100 GeV. Nuclear data for science and technology, edited by K. H. Bockhoff, pp. 1007–1009. Amsterdam: North-Holland.Google Scholar
First citation Gerward, L. (1981). X-ray attenuation coefficients and atomic photoelectric absorption cross sections of silicon. J. Phys. B, 14, 3389–3395.Google Scholar
First citation Gerward, L. (1982). X-ray attenuation coefficients of copper in the energy range 5 to 50 keV. Z. Naturforsch. Teil A, 37, 451–459.Google Scholar
First citation Gerward, L. (1983). X-ray attenuation coefficients of carbon in the energy range 5 to 20 keV. Acta Cryst. A39, 322–325.Google Scholar
First citation Gerward, L. (1986). Empirical absorption equations for use in X-ray spectrometric analysis. X-ray Spectrom. 15, 29–33.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 Giles, C., Vettier, C., de Bergevin, F., Grubel, G., Goulon, J. & Grossi, F. (1994). X-ray polarimetry with phase plates. ESRF Newsletter, 21, 16–17.Google Scholar
First citation Godwin, R. P. (1968). Synchrotron radiation light source. Springer Tracts Mod. Phys. 51, 1–69.Google Scholar
First citation Goldberg, M. (1961). Intensités relatives des raies X du spectre L excité par bombardement électronique des éléments lourds. J. Phys. Radium, 22, 743–748.Google Scholar
First citation Goldsztaub, S. (1947). Tube à rayons X de grande brilliance à foyer ponctuel. C. R. Acad. Sci. 224, 458–459.Google Scholar
First citation Green, M. (1963). The target absorption correction in X-ray microanalysis. X-ray optics and X-ray microanalysis, edited by H. Pattee, V. E. Cosslett & A. Engstrom, pp. 361–377. London: Academic Press.Google Scholar
First citation Green, M. & Cosslett, V. E. (1968). Measurement of K, L and M shell X-ray production efficiencies. Br. J. Appl. Phys. Ser. 2, 1, 425–436.Google Scholar
First citation Grey, D. (1996). Instrumentation developments and observations in hard X-ray astronomy. PhD thesis. The University of New South Wales, Australia.Google Scholar
First citation Grimvall, G. & Persson, E. (1969). Absorption of X-rays in germanium. Acta Cryst. A25, 417–422.Google Scholar
First citation Guo, C.-L. & Wu, Y.-Q. (1985). Empirical relationship between the characteristic X-ray intensity and the incident electron energy. Kexue Tongbao, 30, 1621–1627.Google Scholar
First citation Gurman, S. J. (1988). The small atom approximation in EXAFS and surface EXAFS. J. Phys. C, 21, 3699–3717.Google Scholar
First citation Gurman, S. J. (1995). Interpretation of EXAFS data. J. Synchrotron Rad. 2, 56–63.Google Scholar
First citation Hanfland, M., Häusermann, D., Snigirev, A., Snigireva, I., Ahahama, Y. & McMahon, M. (1994). Bragg–Fresnel lens for high pressure studies. ESRF Newsletter, 22, 8–9.Google Scholar
First citation Hart, M. (1968). An ångström ruler. J. Phys. D, 1, 1405–1408.Google Scholar
First citation Hart, M. (1971). Bragg reflection X-ray optics. Rep. Prog. Phys. 34, 435–490.Google Scholar
First citation Hart, M. (1985). Private communication.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. & Rodrigues, A. R. D. (1978). Harmonic-free single-crystal monochromators for neutrons and X-rays. J. Appl. Cryst. 11, 248–253.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 Hartree, D. R. (1928). The wavemechanics of an atom with a non-Coulomb central field. Proc. Cambridge Philos. Soc. 24, 89–132.Google Scholar
First citation Härtwig, J., Hölzer, G., Wolf, J. & Förster, E. (1993). Remeasurement of the profile of the characteristic Cu Kα emission line with high precision and accuracy. J. Appl. Cryst. 26, 539–548.Google Scholar
First citation Hashimoto, H., Kozaki, S. & Ohkawa, T. (1965). Observations of Pendellösung fringes and images of dislocations by X-ray shadow micrographs of silicon crystals. Appl. Phys. Lett. 6, 16–17.Google Scholar
First citation Hashizume, H. (1983). Asymmetrically grooved monolithic crystal monochromators for suppression of harmonics in synchrotron X-radiation. J. Appl. Cryst. 16, 420–427.Google Scholar
First citation Hasnain, S. S. (1990). X-ray absorption fine structure. London: Ellis Horwood.Google Scholar
First citation Hastings, J. B., Eisenberger, P., Lengeler, B. & Perlman, M. L. (1975). Local structure determination at high dilution: internal oxidation at 75 ppm Fe in Cu. Phys. Rev. Lett. 43, 1807–1810.Google Scholar
First citation Heinrich, K. F. J. (1966). X-ray absorption uncertainty. The electron microprobe, edited by T. D. McKinley, K. F. J. Heinrich & D. B. Wittrey, pp. 296–377. New York: John Wiley.Google Scholar
First citation Helliwell, J. R. (1984). Synchrotron X-radiation protein crystallography. Rep. Prog. Phys. 47, 1403–1497.Google Scholar
First citation Hendrickson, W. A. (1994). MAD phasing for macromolecular structures. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 159–173. Amsterdam: North Holland.Google Scholar
First citation Henins, A. (1971). Ruled grating measurements of the Al Kα1,2 wavelength. Precision measurements and fundamental constants, edited by D. N. Langenberg & B. N. Taylor, pp. 255–258. Natl Bur. Stand. (US) Spec. Publ. No. 343. Gaithersburg, Maryland: National Bureau of Standards.Google Scholar
First citation Henke, B. L., Lee, P., Tanaka, T. J., Shimambukuro, R. L. & Fujikawa, B. K. (1982). Low-energy X-ray interaction coefficients: photoabsorption, scattering and reflection. At. Data Nucl. Data Tables, 27, 1–144.Google Scholar
First citation Hertz, G. (1920). Über die Absorptionsgrenzen in den L-Serie. Z. Phys. 3, 19–27.Google Scholar
First citation Hida, M., Wada, N., Maeda, H., Hikaru, T., Tsu, Y. & Kamino, N. (1985). An EXAFS investigation on the lattice relaxation of Ni fine particles prepared by gas evaporation. Jpn. J. Appl. Phys. 24, L3–L5.Google Scholar
First citation Hofmann, A. (1978). Quasi-monochromatic synchrotron radiation from undulators. Nucl. Instrum. Methods, 152, 17–21.Google Scholar
First citation Holt, S. A., Brown, A. S., Creagh, D. C. & Leon, R. (1997). The application of grazing incidence X-ray diffraction and specular reflectivity to the structural investigation of multiple quantum well and quantum dot semiconductor devices. J. Synchrotron Rad. 4, 169–174.Google Scholar
First citation Hölzer, G., Fritsch, M., Deutsch, M., Härtwig, J. & Förster, E. (1997). 1,2 and Kβ1,3 X-ray emission lines of the 3d transition metals. Phys. Rev. A, 56, 4554–4568.Google Scholar
First citation Honkimaki, V., Sleight, J. & Suortti, P. (1990). Characteristic X-ray flux from sealed Cr, Cu, Mo, Ag and W tubes. J. Appl. Cryst. 23, 412–417.Google Scholar
First citation Hönl, H. (1933a). Zur Dispersions theorie der Röntgenstrahlen. Z. Phys. 84, 1–16.Google Scholar
First citation Hönl, H. (1933b). Atomfactor für Röntgenstrahlen als Problem der Dispersionstheorie (K-Schale). Ann. Phys. (Leipzig), 18, 625–657.Google Scholar
First citation Hosoya, S. (1975). Anomalous scattering measurements and amplitude and phase determinations with continuous X-rays. Anomalous scattering, edited by S. Ramaseshan & S. C. Abrahams, pp. 275–287. Copenhagen: Munksgaard.Google Scholar
First citation Hubbell, J. H. (1969). Phonon cross sections, attenuation coefficients, and energy absorption coefficients from 10 keV to 100 GeV. Report NRDS-NBS29. National Institute of Standards and Technology, Gaithersburg, MD, USA.Google Scholar
First citation Hubbell, J. H., Gerstenberg, H. M. & Saloman, E. B. (1986). Bibliography of photon total cross section (attenuation coefficient) measurements 10 eV to 13.5 GeV. Report NBSIR 86-3461. National Institute of Standards and Technology, Gaithersburg, MD, USA.Google Scholar
First citation Hubbell, J. H., McMaster, W. H., Del Grande, N. K. & Mallett, J. H. (1974). X-ray cross sections and attenuation coefficients. International tables for X-ray crystallography, Vol. IV, edited by J. A. Ibers & W. C. Hamilton, pp. 47–70. Birmingham: Kynoch Press.Google Scholar
First citation Hubbell, J. H. & Øverbø, I. (1979). Relativistic atomic form factors and photon coherent scattering cross sections. J. Phys. Chem. Ref. Data, 8, 69–105.Google Scholar
First citation Hubbell, J. H., Veigele, W. J., Briggs, E. A., Brown, R. T., Cromer, D. T. & Howerton, R. J. (1975). Atomic form factors, incoherent scattering functions and photon scattering cross sections. J. Phys. Chem. Ref. Data, 4, 471–538.Google Scholar
First citation Huke, K. & Kobayakawa, M. (1989). World-wide census of SR facilities. Rev. Sci. Instrum. 60, 2548–2561.Google Scholar
First citation Indelicato, P. (1990). Kα transitions in few-electron ions and in atoms. X-ray and inner-shell processes, edited by T. A. Carlson, M. O. Krause & S. T. Manson, pp. 591–601. New York: American Institute of Physics.Google Scholar
First citation Indelicato, P. & Desclaux, J. P. (1990). Multiconfiguration Dirac–Fock calculations of transition energies with QED corrections in three-electron ions. Phys. Rev. A, 42, 5139–5149.Google Scholar
First citation Indelicato, P., Gorceix, O. & Desclaux, J. P. (1987). MCDF studies of two electron ions II: radiative corrections and comparison with experiment. J. Phys. B, 20, 651.Google Scholar
First citation Indelicato, P. & Lindroth, E. (1992). Relativistic effects, correlation, and QED corrections on Kα transitions in medium to very heavy atoms. Phys. Rev. A, 46, 2426–2436.Google Scholar
First citation Indelicato, P. & Lindroth, E. (1996). Current status of the relativistic theory of inner hole states in heavy atoms. Comments At. Mol. Phys. 32, 197–208.Google Scholar
First citation Indelicato, P. & Mohr, P. J. (1990). Electron screening correction to the self energy in high-Z atoms. 12th International Conference on Atomic Physics, Ann Arbor, Michigan, USA.Google Scholar
First citation Indelicato, P. & Mohr, P. J. (1991). Quantum electrodynamic effects in atomic structure. Theor. Chim. Acta, 80, 207–214.Google Scholar
First citation International Tables for Crystallography (2001). Vol. B. Dordrecht: Kluwer Academic Publishers.Google Scholar
First citation International Tables for X-ray Crystallography (1974). Vol. IV. Birmingham: Kynoch Press.Google Scholar
First citation Ishida, K. & Katoh, H. (1982). Use of multiple reflection diffraction to measure X-ray refractive index. Jpn. J. Appl. Phys. 21, 1109.Google Scholar
First citation Ishimura, T., Shiraiwa, Y. & Sawada, M. (1957). The input power limit of the cylindrical rotating anode of an X-ray tube. J. Phys. Soc. Jpn, 12, 1064–1070.Google Scholar
First citation Jackson, D. F. & Hawkes, D. J. (1981). X-ray attenuation coefficients of elements and mixtures. Phys. Rep. 70, 169–233.Google Scholar
First citation James, R. W. (1955). The optical principles of the diffraction of X-rays. Ithaca: Cornell University Press.Google Scholar
First citation Jenkins, R., Manne, R., Robin, J. & Senemaud, C. (1991). Nomenclature, symbols, units and their usage in spectrochemical analysis. VIII. Nomenclature system for X-ray spectroscopy. Pure Appl. Chem. 63, 735–746.Google Scholar
First citation Jennings, L. D. (1981). Extinction, polarization and crystal monochromators. Acta Cryst. A37, 584–593.Google Scholar
First citation Jennings, L. D. (1984). The polarization ratio of crystal monochromators. Acta Cryst. A40, 12–16.Google Scholar
First citation Johnson, W. R. & Soff, G. (1985). The Lamb shift in hydrogenlike atoms, 1 ≤ Z ≤ 110. At. Data Nucl. Data Tables, 33, 405.Google Scholar
First citation Joy, D. C. & Maher, D. (1985). Quantitative electron energy loss spectroscopy: an introduction to the power of Kevex Elstar software. Kevex Analyst, 10, 6–9.Google Scholar
First citation Kane, P. P., Kissel, L., Pratt, R. H. & Roy, S. C. (1986). Elastic scattering of γ-rays and X-rays by atoms. Phys. Rep. 150, 75–159.Google Scholar
First citation Karle, J. (1980). Anomalous dispersion and the use of triplet phase invariants. Int. J. Quantum Chem. 7, 357–367.Google Scholar
First citation Karle, J. (1984a). Rules for evaluating triplet phase invariants by use of anomalous dispersion data. Acta Cryst. A40, 366–379.Google Scholar
First citation Karle, J. (1984b). The relative scaling of multiple wavelength anomalous dispersion data. Acta Cryst. A40, 1–4.Google Scholar
First citation Karle, J. (1984c). Triplet phase invariants from an exact algebraic analysis of anomalous dispersion. Acta Cryst. A40, 526–532.Google Scholar
First citation Karle, J. (1985). Many algebraic formulas for the evaluation of triplet phase invariants from isomorphous replacement and anomalous dispersion data. Acta Cryst. A41, 182–189.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 Kessler, E. G. Jr, Deslattes, R. D. & Henins, A. (1979). Wavelength of the W Kα1 X-ray line. Phys. Rev. A, 19, 215–218.Google Scholar
First citation Kikuta, S. (1971). X-ray crystal monochromators using successive asymmetric diffractions and their applications to measurements of diffraction curves. II. Type 1 collimator. J. Phys. Soc. Jpn, 30, 222–227.Google Scholar
First citation Kikuta, S. & Kohra, K. (1970). X-ray crystal collimators using successive asymmetric diffractions and their applications to measurements of diffraction curves. I. General considerations on collimators. J. Phys. Soc. Jpn, 29, 1322–1328.Google Scholar
First citation Kim, Y. K., Baik, D. H., Indelicato, P. & Desclaux, J. P. (1991). Resonance transition energies of Li-, Na-, and Cu-like ions. Phys. Rev. A, 44, 148–166.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 Kirkpatrick, P. & Wiedmann, L. (1945). Theoretical continuous X-ray energy and polarization. Phys. Rev. 67, 321–339.Google Scholar
First citation Kissel, L. (1977). Rayleigh scattering: elastic scattering by bound electrons. PhD thesis, University of Pittsburgh, PA, USA.Google Scholar
First citation Kissel, L. & Pratt, R. H. (1985). Rayleigh scattering: elastic photon scattering by bound electrons. In Atomic inner-shell physics, edited by B. Crasemann. New York: Plenum.Google Scholar
First citation Kissel, L., Pratt, R. H., Kane, P. P. & Roy, S. C. (1985). Elastic scattering of X-rays and γ-rays by atoms. Sandia Report SANDE5-0501J. Sandia, Albuquerque, NM, USA.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 Koch, B., MacGillavry, C. H. & Milledge, H. J. (1962). Absorption. International tables for X-ray crystallography, Vol. III, Section 2.2, pp. 157–192. Birmingham: Kynoch Press.Google Scholar
First citation Koch, E. E. (1983). Editor. Handbook on synchrotron radiation. Amsterdam: North-Holland.Google Scholar
First citation Kohn, W. & Sham, L. S. (1965). Self consistent equations including exchange and correlation effects. Phys. Rev. A, 140, 1133–1138.Google Scholar
First citation Kostarev, A. L. (1941). Theory of the fine structure of X-ray absorption spectra in solids. Zh. Eksper. Teor. Fiz. 11, 60–73.Google Scholar
First citation Kostarev, A. L. (1949). Elucidation of the super-fine structure of X-ray absorption spectra in solids. Zh. Eksper. Teor. Fiz. 19, 413–420.Google Scholar
First citation Kraft, S., Stümpel, J., Becker, P. & Kuetgens, U. (1996). High resolution X-ray absorption spectroscopy with absolute energy calibration for the determination of absorption edge energies. Rev. Sci. Instrum. 67, 681–687.Google Scholar
First citation Kramers, H. A. (1923). On the theory of X-ray absorption and of the continuous X-ray spectrum. Philos. Mag. 46, 836–871.Google Scholar
First citation Kronig, R. de L. (1932a). Zur Theorie der Feinstruktur in den Röntgenabsorptionspecktren II. Z. Phys. 75, 191–210.Google Scholar
First citation Kronig, R. de L. (1932b). Zur Theorie der Feinstruktur in den Röntgenabsorptionspecktren III. Z. Phys. 75, 468–475.Google Scholar
First citation Kulenkampff, H. & Schmidt, L. (1943). Die Energieverteilung im Spektrum der Röntgen Bremsstrahlung. Ann. Phys. (Leipzig), 43, 494–512.Google Scholar
First citation Kulipanov, G. N. & Skrinskii, A. N. (1977). Utilization of synchrotron radiation: current status and prospects. Usp. Fiz. Nauk, 122, 369–418. English translation: Sov. Phys. Usp. 20, 559–586.Google Scholar
First citation Kumakhov, M. A. & Komarov, F. F. (1990). Multiple reflection from surface X-ray optics. Phys. Rep. 191(5), 289–350.Google Scholar
First citation Kunz, C. (1979). Editor. Synchrotron radiation. Berlin: Springer.Google Scholar
First citation Kuroda, H., Ohta, T., Murata, T., Udagawa, Y. & Nomura, M. (1992). XAFS VII: Proceedings of the Seventh Conference on X-ray Absorption Fine Structure. Jpn. J. Appl. Phys. 32, Suppl. 32-2.Google Scholar
First citation Kusev, S. V., Raiko, V. I. & Skuratowski, I. Ya. (1992). The status of beam lines for macromolecular crystallography at the Siberia-2 storage ring. Rev. Sci. Instrum. 63, 1055–1057.Google Scholar
First citation Kutzler, F. W., Natoli, C. R., Misemer, D. K., Doniach, S. & Hodgson, K. O. (1981). Use of one electron theory for the interpretation of near edge structure in K-shell X-ray absorption spectra of transition metal complexes. J. Chem. Phys. 73, 3274–3288.Google Scholar
First citation Laclare, J. L. (1994). Target specifications and performance of the ESRF source. J. Synchrotron Rad. 1, 12–18.Google Scholar
First citation Lea, K. (1978). Highlights of synchrotron radiation. Phys. Rep. 43, 337–375.Google Scholar
First citation Leapman, R. D. & Cosslett, V. E. (1976). Extended fine structure above the X-ray edge in electron energy loss spectra. J. Phys. D, 9, L29–L31.Google Scholar
First citation Lebugle, A., Axelsson, U., Nyholm, R. & Mårtensson, N. (1981). Experimental L and M core level binding energies for the metals 22Ti to 30Zn. Phys Scr. 23, 825–827.Google Scholar
First citation Lee, P. A. (1981). Theory of extended X-ray absorption fine structure. EXAFS spectroscopy: techniques and applications, edited by B. K. Teo & D. C. Joy, Chap. 2, pp. 5–13. New York: Plenum.Google Scholar
First citation Lee, P. A. & Beni, G. (1977). New method for the calculation of atomic phase shifts: application to extended X-ray absorption fine structure (EXAFS) in molecules and crystals. Phys. Rev. B, 15, 2862–2883.Google Scholar
First citation Lee, P. A., Citrin, P. H., Eisenberger, P. & Kincaid, B. M. (1981). Extended X-ray absorption fine structure – its strengths and limitations as a structural tool. Rev. Mod. Phys. 53, 769–787.Google Scholar
First citation Lengeler, B. (1994). Experimental determination of the dispersion correction f′(E) to the atomic scattering factor. Resonant anomalous X-ray scattering, edited by G. Materlik, C. J. Sparks & K. Fischer, pp. 35–69. Amsterdam: North Holland.Google Scholar
First citation Lengeler, B., Materlik, G. & Müller, J. E. (1983). Near edge structure in cerium and cerium compounds. EXAFS and near edge structure, edited by A. Bianconi, L. Incoccia & S. Stipcich, pp. 150–153. Berlin: Springer.Google Scholar
First citation Lereboures, B., Dürr, J., d'Huysser, A., Bonelle, J. P. & Lenglet, M. (1980). Phys. Status Solidi, 62, K175.Google Scholar
First citation Leroux, J. & Thinh, T. P. (1977). Revised tables of X-ray mass attenuation coefficients. Quebec: Corporation Scientifique Claisse, Inc.Google Scholar
First citation Levinger, J. S. (1952). Small angle coherent scattering of gammas by bound electrons. Phys. Rev. 87, 656–662.Google Scholar
First citation Liberman, D., Waber, J. T. & Cromer, D. T. (1965). Self-consistent field Dirac–Slater wavefunctions for atoms and ions. I. Comparison with previous calculations. Phys. Rev. A, 137, 27–34.Google Scholar
First citation Liebhafsky, H. A., Pfeiffer, H. G., Winslow, E. H. & Zemany, P. D. (1960). X-ray absorption and emission analytical chemistry, pp. 313–317. New York: John Wiley.Google Scholar
First citation Lindgren, I., Persson, H., Salomonson, S. & Labzowsky, L. (1995). Full QED calculations of two-photon exchange for heliumlike systems: analysis in the Coulomb and Feynman gauge. Phys. Rev. A, 51, 1167–1195.Google Scholar
First citation Lindroth, E. & Indelicato, P. (1993). Inner shell transitions in heavy atoms. Phys. Scr. T46, 139–143.Google Scholar
First citation Lindroth, E. & Indelicato, P. (1994). High precision calculations of inner shell transitions in heavy elements. Nucl. Instrum. Methods, B87, 222–226.Google Scholar
First citation Lum, G. K., Wiegand, C. E., Kessler, E. G. Jr, Deslattes, R. D., Jacobs, L., Schwitz, W. & Seki, R. (1981). Kaonic mass by critical absorption of kaonic atom X-rays. Phys. Rev. D, 23, 2522–2532.Google Scholar
First citation Lytle, F. W., Sayers, D. E. & Stern, E. A. (1989). Report of the International Workshops on Standards and Criteria in XAFS. In X-ray absorption spectroscopy. Physica (Utrecht), B158, 701–722.Google Scholar
First citation Lytle, F. W., Stern, E. A. & Sayers, D. E. (1975). Extended X-ray-absorption fine-structure technique II. Experimental practice and selected results. Phys. Rev. B, 11, 4825–4835.Google Scholar
First citation McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969/1970). Compilation of X-ray cross sections. Report UCRL-50174. (Section I, 1970; Section II, 1969; Section III, 1969; Section IV, 1969.) Lawrence Livermore National Laboratory, Livermore, CA, USA.Google Scholar
First citation Marcus, M., Powers, L. S., Storm, A. R., Kincaid, B. M. & Chance, B. (1980). Curved-crystal (LiF) X-ray focusing array for fluorescence EXAFS in dilute samples. Rev. Sci. Instrum. 51, 1023–1029.Google Scholar
First citation Martens, G. & Rabe, P. (1980). EXAFS studies on superficial regions by means of total reflection. Phys. Status Solidi A, 58, 415–425.Google Scholar
First citation Maruyama, X. K., Di Nova, K., Snyder, D., Piestrup, M. A., Li, Q., Fiorito, R. B. & Rule, D. W. (1993). A compact tunable X-ray source based on parametric X-ray generation by moderate-energy linacs. Proc. 1993 Particle Accelerator Conference. IEEE, 2, 1620–1622.Google Scholar
First citation Materlik, G., Bedzyk, M. J. & Frahm, A. (1984). Report SR-84-07. DESY, Hamburg, Germany.Google Scholar
First citation Matsushita, T., Ishikawa, T. & Oyanagi, H. (1986). Sagitally focusing double-crystal monochromator with constant exit height at the Photon Factory. Nucl. Instrum. Methods, A246, 377–379.Google Scholar
First citation Matsushita, T., Kikuta, S. & Kohra, K. (1971). X-ray crystal monochromators using successive asymmetric diffractions and their applications to measurements of diffraction curves. III. Type I1 collimators. J. Phys. Soc. Jpn, 30, 1136–1144.Google Scholar
First citation Metchnik, V. & Tomlin, S. G. (1963). On the absolute intensity of emission of characteristic X radiation. Proc. Phys. Soc. London, 81, 956–964.Google Scholar
First citation Mohr, P. J. (1974a). Numerical evaluation of the 1S1/2 state radiative level shift. Ann. Phys. (Leipzig), 88, 52–87.Google Scholar
First citation Mohr, P. J. (1974b). Self-energy radiative corrections in hydrogen-like systems. Ann. Phys. (Leipzig), 88, 26–51.Google Scholar
First citation Mohr, P. J. (1975). Lamb shift in a strong Coulomb potential. Phys. Rev. Lett. 34, 1050–1052.Google Scholar
First citation Mohr, P. J. (1982). Self-energy of the n = 2 states in a strong Coulomb field. Phys. Rev. A, 26, 2338–2354.Google Scholar
First citation Mohr, P. J. (1992). Self-energy correction to one-electron energy levels in a strong Coulomb field. Phys. Rev. A, 46, 4421–4424.Google Scholar
First citation Mohr, P. J. & Soff, G. (1993). Nuclear size correction to the electron self-energy. Phys. Rev. Lett. 70, 158–161.Google Scholar
First citation Mohr, P. J. & Taylor, B. N. (2000). CODATA recommended values of the fundamental physical constants: 1998. Rev. Mod. Phys. 72, 351–495.Google Scholar
First citation Montenegro, E. C., Baptista, G. B. & Duarte, P. W. E. P. (1978). K and L X-rays mass attenuation coefficients for low-Z materials. At. Data Nucl. Data Tables, 22, 131–177.Google Scholar
First citation Mooney, T., Lindroth, E., Indelicato, P., Kessler, E. & Deslattes, R. D. (1992). Precision measurements of K and L transitions in xenon: experiment and theory for the K, L and M levels. Phys. Rev. A, 45, 1531–1543.Google Scholar
First citation Mooney, T. M. (1996). Personal communication.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 Müller, A. (1927). Input limits of X-ray generators. Proc. R. Soc. London Ser. A, 117, 30–42.Google Scholar
First citation Müller, A. (1929). A spinning-target X-ray generator and its input limit. Proc. R. Soc. London Ser. A, 125, 507–516.Google Scholar
First citation Müller, A. (1931). Further estimates of the input limits of X-ray generators. Proc. R. Soc. London Ser. A, 132, 646–649.Google Scholar
First citation Mustre de Leon, J., Stern, E. A., Sayers, D. E., Ma, Y. & Rehr, J. J. (1988). XAFS V: Proceedings of the Fifth International Conference on X-ray Absorption Fine Structure. Amsterdam: North-Holland.Google Scholar
First citation Nagel, D. J. (1980). Comparison of X-ray sources for exposure of photo resists. Ann. NY Acad. Sci. 342, 235–247.Google Scholar
First citation Natoli, C. R. (1990). Multichannel multiple scattering theory with general potentials. Phys. Rev. B, 42, 1944–1968.Google Scholar
First citation Natoli, C. R., Misemer, D. K., Doniach, S. & Kutzler, F. W. (1980). First principles calculation of X-ray absorption edge structure in molecular clusters. Phys. Rev. A, 22, 1104–1108.Google Scholar
First citation Nelms, A. T. & Oppenheimer, I. (1955). J. Res. Natl Bur. Stand. 55, 53–62.Google Scholar
First citation Nordfors, B. (1960). The statistical error in X-ray absorption measurements. Ark. Fys. 18, 37–47.Google Scholar
First citation Norman, D., Durham, P. J. & Pendry, J. B. (1983). The absorption of oxygen on nickel (001) studied by XANES. EXAFS and near edge structure, edited by A. Bianconi, L. Incoccia & S. Stipcich, pp. 144–146. Berlin: Springer.Google Scholar
First citation Nyholm, R., Berndtsson, A. & Mårtensson, N. (1980). Core level binding energies for the elements Hf to Bi (Z = 72–83). J. Phys. C, 13, L1091–L1096.Google Scholar
First citation Nyholm, R. & Mårtensson, N. (1980). Core level binding energies for the elements Zr–Te (Z = 40–52). J. Phys. C, 13, L279-L284.Google Scholar
First citation Oosterkamp, W. J. (1948a). The heat dissipation in the anode of an X-ray tube. Philips Res. 3, 49–59.Google Scholar
First citation Oosterkamp, W. J. (1948b). The heat dissipation in the anode of an X-ray tube. Philips Res. 3, 161–173.Google Scholar
First citation Oosterkamp, W. J. (1948c). The heat dissipation in the anode of an X-ray tube. Philips Res. 3, 303–317.Google Scholar
First citation Oshima, K., Harada, J. & Sakabe, N. (1986). Curved crystal monochromator. X-ray instrumentation for the Photon Factory: dynamic analyses of micro-structures in matter, edited by S. Hosoya, Y. Iitaka & H. Hashizume, pp. 35–41. Japan: Photon Factory.Google Scholar
First citation OSMIC (1996). Catalogue. A new family of collimating and focusing optics for X-ray analysis. OSMIC, Michigan, USA.Google Scholar
First citation Øverbø, I. (1977). The Coulomb correction to electron pair production by intermediate-energy photons. Phys. Lett. B, 71, 412–414.Google Scholar
First citation Øverbø, I. (1978). Large-q form factors for light atoms. Phys. Scr. 17, 547–549.Google Scholar
First citation Oyanagi, H., Ihara, H., Matsushita, T., Hirabayashi, M., Terada, N., Tokumoto, M., Senzaki, K., Kimura, T. & Yao, T. (1987). Short range order in high Tc superconductors BaxY1−xCuO3−y and SrxLa2−xCuO4–7. Jpn. J. Appl. Phys. 26, L828–L831.Google Scholar
First citation Oyanagi, H., Martini, M., Saito, M. & Haga, K. (1995). Nineteen element high purity Ge solid state detector array for fluorescence X-ray absorption fine structure studies. Submitted to Rev. Sci. Instrum.Google Scholar
First citation Oyanagi, H., Matsushita, T., Tanoue, H., Ishiguro, T. & Kohra, K. (1985). Fluorescence-detected X-ray absorption spectroscopy applied to structural characterization of very thin films: ion-beam-induced modification of thin Ni layers on Si (100). Jpn. J. Appl. Phys. 24, 610–619.Google Scholar
First citation Oyanagi, H., Takeda, T., Matsushita, T., Ishiguro, T. & Sasaki, A. (1986). Local structure in InGaAsP I quaternary alloys. J. Phys. (Paris), 28, Suppl. 12, C8, 423–426.Google Scholar
First citation Pantos, E. (1982). The SRS program library documentation. Daresbury: SRC.Google Scholar
First citation Papatzacos, P. & Mort, K. (1975). Delbrück scattering calculations. Phys. Rev. D, 12, 206–221.Google Scholar
First citation Parratt, L. G. (1959). Electronic band structure of solids by X-ray spectroscopy. Rev. Mod. Phys. 31, 616–645.Google Scholar
First citation Parratt, L. G., Porteus, I. O., Schnopper, H. W. & Watanabe, T. (1959). X-ray absorption coefficients and geometrical collimation of the beam. Rev. Sci. Instrum. 30, 344–347.Google Scholar
First citation Peele, A. G., Nugent, K. A., Rode, A. V., Gabel, K., Richardson, M. C. M., Strack, R. & Siegmund, W. (1996). X-ray focusing with lobster-eye optics: a comparison of theory with experiment. Appl. Opt. 35, 4420–4425.Google Scholar
First citation Pendry, J. B. (1983). The transition region between XANES and EXAFS. EXAFS and near edge structure, edited by A. Bianconi, L. Incoccia & S. Stipcich, pp. 4–10. Berlin: Springer.Google Scholar
First citation Petiau, J. & Calas, G. (1983). EXAFS for inorganic systems, pp. 127–129. Daresbury: SRC.Google Scholar
First citation Philips, J. C. & Hodgson, K. O. (1985). Single-crystal X-ray diffraction and anomalous scattering using synchrotron radiation. Synchrotron radiation research, edited by H. Winick & S. Doniach, pp. 565–604. New York: Plenum.Google Scholar
First citation Philips, J. C., Templeton, D. H., Templeton, L. K. & Hodgson, K. O. (1978). L-III edge anomalous X-ray scattering by cesium measured with synchrotron radiation. Science, 201, 257–259.Google Scholar
First citation Phillips, W. C. (1985). X-ray sources. Methods Enzymol. 114, 300–316.Google Scholar
First citation Piestrup, M. A., Boyers, D. G., Pincus, C. I., Harris, J. L., Maruyama, X. K., Bergstrom, J. C., Caplan, H. S., Silzer, R. M. & Skopik, D. M. (1991). Quasimonochromatic X-ray source using photoabsorption-edge transition radiation. Phys. Rev. A, 43, 3653–3661.Google Scholar
First citation Piestrup, M. A., Moran, M. J., Boyers, D. G., Pincus, C. I., Kephart, J. O., Gearhart, R. A. & Maruyama, X. K. (1991). Generation of hard X-rays from transition radiation using high-density foils and moderate-energy electrons. Phys. Rev. A, 43, 2387–2396.Google Scholar
First citation Pirenne, M. H. (1946). The diffraction of X-rays and electrons by free molecules. Cambridge University Press.Google Scholar
First citation Plechaty, E. F., Cullen, E. E. & Howerton, R. J. (1981). Tables and graphs of photon-interaction cross sections from 0.1 keV to 100 MeV. Derived from the LLL Evaluated-Nuclear-Data Library. Report UCRL-50400, Vol. 6, Rev. 3. Lawrence Livermore National Laboratory, Livermore, CA, USA.Google Scholar
First citation Powell, C. J. (1995). Elemental binding energies for X-ray photoelectron spectroscopy. Appl. Surf. Sci. 89, 141–149.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 Ramaseshan, S. & Abrahams, S. C. (1975). Anomalous scattering. Copenhagen: Munksgaard.Google Scholar
First citation Reed, S. J. B. (1975). Electron microprobe analysis. Cambridge University Press.Google Scholar
First citation Rehr, J. J. & Albers, R. C. (1990). Phys. Rev. B, 41, 8139–8144.Google Scholar
First citation Rieck, C. D. (1962). Tables relating to the production, wavelengths, and intensities of X-rays. International tables for X-ray crystallography, Vol. III, edited by C. H. MacGillavry & G. D. Rieck, pp. 59–72. Birmingham: Kynoch Press.Google Scholar
First citation Riggs, P. J., Mei, R., Yocum, C. F. & Penner-Hahn, J. E. (1993). The characterization of the Mn site in the photosynthetic oxygen evolving complexes: the effect of hydroxylamine and hydroquinone on the XAFS. Jpn. J. Appl. Phys. 32, Suppl. 32–2, 527–529.Google Scholar
First citation Roy, S. C., Kissel, L. & Pratt, R. H. (1983). Elastic photon scattering at small momentum transfer and validity of form-factor theories. Phys. Rev. A, 27, 285–290.Google Scholar
First citation Roy, S. C. & Pratt, R. H. (1982). Validity of non relativistic dipole approximation for forward Rayleigh scattering. Phys. Rev. A, 26, 651–653.Google Scholar
First citation Rudakov, L. I., Baigarin, K. A., Kalinin, Y. G., Korolev, V. D. & Kumachov, M. A. (1991). Pulsed-plasma-based X-ray source and new X-ray optics. Phys. Fluids B (Plasma Phys.), 3, 2414–2419.Google Scholar
First citation Sakurai, K. (1993). High-intensity X-ray line-focal spot for laboratory EXAFS measurements. Rev. Sci. Instrum. 64, 267–268.Google Scholar
First citation Sakurai, K. & Sakurai, H. (1994). Comment on high intensity low tube-voltage X-ray source for laboratory EXAFS measurements. Rev. Sci. Instrum. 65, 2417–2418.Google Scholar
First citation Saloman, E. B. & Hubbell, J. H. (1986). X-ray attenuation coefficients (total cross sections): comparison of the experimental data base with the recommended values of Henke and the theoretical values of Scofield for energies between 0.1–100 keV. Report NBSIR 86-3431. National Institute of Standards and Technology, Gaithersburg, MD, USA.Google Scholar
First citation Saloman, E. B. & Hubbell, J. H. (1987). Critical analysis of soft X-ray cross section data. Nucl. Instrum. Methods, A255, 38–42.Google Scholar
First citation Saloman, E. B., Hubbell, J. H. & Scofield, J. H. (1988). At. Data Nucl. Data Tables, 38, 1–197.Google Scholar
First citation Sano, H., Ohtaka, K. & Ohtsuki, Y. H. (1969). Normal and abnormal absorption coefficients of X-rays. J. Phys. Soc. Jpn, 27, 1254–1261.Google Scholar
First citation Sawada, M., Tsutsumi, K., Shiraiwa, T., Ishimura, T. & Obashi, M. (1959). Some contributions to the X-ray spectroscopy of solid state. Annu. Rep. Sci. Works Osaka Univ. 7, R–87.Google Scholar
First citation Sayers, D. E., Lytle, F. W. & Stern, E. A. (1970). Point scattering theory of X-ray K absorption fine structure. Adv. X-ray Anal. 13, 248–256.Google Scholar
First citation Schaupp, D., Schumacher, M., Smend, F., Rullhusen, P. & Hubbell, J. H. (1983). Small-angle Rayleigh scattering of photons at high energies: tabulations of relativistic HFS modified atomic form factors. J. Phys. Chem. Ref. Data, 12, 467–512.Google Scholar
First citation Schmidt, V. V. (1961a). Contribution to the theory of the temperature dependence of the fine structure of X-ray absorption spectra. Bull. Acad. Sci. USSR Phys. Ser. 25, 988–993.Google Scholar
First citation Schmidt, V. V. (1961b). On the effect of the temperature dependence of the fine structure of X-ray absorption spectra. J. Exp. Theor. Phys. 12, 886–890.Google Scholar
First citation Schmidt, V. V. (1963). Contributions to the theory of the dependence of the fine structure of X-ray absorption spectra II. Case of high temperatures. Bull. Acad. Sci. USSR Phys. Ser. 27, 392–397.Google Scholar
First citation Schwarzenbach, D., Abrahams, S. C., Flack, H. D., Prince, E. & Wilson, A. J. C. (1995). Statistical descriptors in crystallography. II. Report of a Working Group on Expression of Uncertainty in Measurement. Acta Cryst. A51, 565–569.Google Scholar
First citation Schweppe, J., Deslattes, R. D., Mooney, T. & Powell, C. J. (1994). Accurate measurement of Mg and Al Kα1,2 X-ray energy profiles. J. Electron Spectrosc. Relat. Phenom. 67, 463–478.Google Scholar
First citation Schweppe, J. E. (1995). Personal communication.Google Scholar
First citation Schwinger, J. (1949). On the classical radiation of accelerated electrons. Phys. Rev. 75, 1912–1925.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 Scofield, J. H. (1986). Personal communication to E. B. Saloman & J. H. Hubbell. Calculated photoeffect values 0.1 to 1.0 keV. [Presented in Saloman & Hubbell (1986)[link] and Saloman et al. (1988)[link].]Google Scholar
First citation Scott, V. D. & Love, G. (1983). Quantitative electron microprobe analysis. Cambridge University Press.Google Scholar
First citation Sears, V. F. (1983). Optimum sample thickness for total cross section measurements. Nucl. Instrum. Methods, 213, 561–562.Google Scholar
First citation Seka, W., Soures, J. M., Lewis, O., Bunkenburg, J., Brown, D., Jacobs, S., Mourou, X. & Zimmermann, J. (1980). High-power phosphate-glass laser system: design and performance characteristics. Appl. Opt. 19, 409–419.Google Scholar
First citation Seka, W., Soures, J. M., Lund, L. & Craxton, R. S. (1981). GDL: a high-power 0.35 µm laser irradiation facility. IEEE J. Quantum Electron. QE-17, 1689–1693.Google Scholar
First citation Sevillano, E., Meuth, H. & Rehr, J. J. (1978). Extended X-ray absorption fine structure Debye-Waller factors. I. Monatomic crystals. Phys. Rev. B, 20, 4908–4911.Google Scholar
First citation Shulman, R. G., Weisenberger, P., Teo, B. K., Kincaid, B. M. & Brown, G. S. (1978). Fluorescence X-ray absorption studies of rubendoxin and its compounds. J. Mol. Biol. 124, 305–315.Google Scholar
First citation Siddons, P. & Hart, M. (1983). Simultaneous measurements of the real and imaginary parts of the X-ray anomalous dispersion using X-ray interferometers. EXAFS and near edge structure, edited by A. Branconi, L. Incoccia & S. Stipcich, pp. 373–375. Berlin: Springer.Google Scholar
First citation Siegbahn, M. (1925). The spectroscopy of X-rays. Oxford University Press.Google Scholar
First citation Siemens (1996a). Parallel beam optics for measurements of samples with irregularly shaped surfaces. Laboratory Report X-ray Analysis, DXRD, 13. Karlsruhe: Siemens.Google Scholar
First citation Siemens (1996b). Goebel mirrors for X-ray reflectometry investigations. Laboratory Report X-ray Analysis, DXRD, 14. Karlsruhe: Siemens.Google Scholar
First citation Siemens (1996c) Grazing incidence diffraction with Goebel mirrors. Laboratory Report X-ray Analysis, DXRD, 15. Karlsruhe: Siemens.Google Scholar
First citation Sinfelt, J. H., Via, G. H. & Lytle, F. W. (1980). Structures of biometallic clusters. Extended X-ray absorption fine structure (EXAFS) studies of Ru–Cu clusters. J. Chem. Phys. 72, 4832–4843.Google Scholar
First citation Smith, D. Y. (1987). Anomalous X-ray scattering: relativistic effects in X-ray dispersion analysis. Phys. Rev. A, 35, 3381–3387.Google Scholar
First citation Snigirev, A. (1994). Bragg–Fresnel optics: new fields of applications. ESRF Newsletter, 22, 20–21.Google Scholar
First citation Soff, G. & Mohr, P. J. (1988). Vacuum polarization in a strong external field. Phys. Rev. A, 38, 5066–5075.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 Stearns, D. G. (1984). Broadening of extended X-ray absorption fine structure ensuing from the finite lifetime of the K hole. Philos. Mag. 49, 541–558.Google Scholar
First citation Stephens, P. W., Eng, P. J. & Tse, T. (1992). Construction and performance of a bent crystal X-ray monochromator. Rev. Sci. Instrum. 64, 374–378.Google Scholar
First citation Stephenson, S. T. (1957). The continuous X-ray spectrum. Handbuch der Physik. XXX: X-rays, edited by S. Flügge, pp. 337–370. Berlin: Springer.Google Scholar
First citation Stern, E. A. (1980). Editor. Laboratory EXAFS facilities 1980. AIP Conf. Proc. No. 64.Google Scholar
First citation Stern, E. A., Bunker, B. & Heald, A. (1981). Understanding the causes of non-transferability of EXAFS amplitude. EXAFS spectroscopy: techniques and applications, edited by B. K. Teo & D. C. Joy, pp. 59–79. New York: Plenum.Google Scholar
First citation Stern, E. A., Sayers, D. E. & Lytle, F. W. (1975). Extended X-ray absorption fine structure technique. III. Determination of physical parameters. Phys. Rev. B, 11, 4836–4846.Google Scholar
First citation Stibius-Jensen, M. (1979). Some remarks on the anomalous scattering factors for X-rays. Phys. Lett. A, 74, 41–44.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 Stohr, J., Denley, D. & Perfettii, P. (1978). Surface extended X-ray absorption fine structure in the soft X-ray region: study of an oxidized Al surface. Phys. Rev. B, 18, 4132–4135.Google Scholar
First citation Storm, E. & Israel, H. I. (1970). Photon cross sections from 0.001 to 100 MeV for elements 1 through 100. Nucl. Data Tables, A7, 565–681.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 Suller, V. P. (1992). Review of the status of synchrotron radiation storage rings. Third European Particle Accelerator Conference. Editions Frontières, 1, 77–81.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 Sussini, J. & Labergerie, D. (1995). Bi-morph piezo-electric mirror: a novel active mirror. Synchrotron Rad. News, 8, 21–26.Google Scholar
First citation Takama, T., Iwasaki, N. & Sato, S. (1980). Measurement of X-ray Pendellösung intensity beats in diffracted white radiation from silicon wafers. Acta Cryst. A36, 1025–1030.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 Takama, T. & Sato, S. (1984). Determination of the atomic scattering factors of germanium by means of the Pendellösung beat measurement using white radiation. Jpn. J. Appl. Phys. 20, 1183–1190.Google Scholar
First citation Taylor, A. (1949). A 5 kW crystallographic X-ray tube with a rotating anode. J. Sci. Instrum. 26, 225–229.Google Scholar
First citation Taylor, A. (1956). Improved demountable crystallographic rotating anode X-ray tube. Rev. Sci. Instrum. 27, 757–759.Google Scholar
First citation Taylor, B. N. & Kuyatt, C. E. (1994). Guidelines for evaluating and expressing the uncertainty of NIST measurement results. NIST Technical Note No. 1297. Gaithersburg, MD: National Institute of Standards and Technology.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. (1982). X-ray dichroism and polarized anomalous scattering of the uranyl ion. Acta Cryst. A38, 62–67.Google Scholar
First citation Templeton, D. H. & Templeton, L. K. (1985a). X-ray dichroism and anomalous scattering of potassium tetrachloroplatinate(II). Acta Cryst. A41, 356–371.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 Templeton, D. H. & Templeton, L. K. (1986). X-ray birefringence and forbidden reflections in sodium bromate. Acta Cryst. A42, 478–481.Google Scholar
First citation Templeton, D. H., Templeton, L. K., Philips, J. C. & Hodgson, K. O. (1980). Anomalous scattering of X-rays by cesium and cobalt measured with synchrotron radiation. Acta Cryst. A36, 436–442.Google Scholar
First citation Templeton, L. K. & Templeton, D. H. (1978). Cesium hydrogen tartrate and anomalous dispersion of cesium. Acta Cryst. A34, 368–373.Google Scholar
First citation Teo, B. K. (1981). EXAFS spectroscopy: techniques and applications, edited by B. K. Teo and D. C. Joy, Chap. 3, pp. 13–59. New York: Plenum.Google Scholar
First citation Teo, B. K. & Joy, D. C. (1981). Editors. EXAFS spectroscopy: techniques and applications. New York: Plenum.Google Scholar
First citation Teo, B. K. & Lee, P. A. (1979). Ab initio calculations of amplitude and phase functions for extended X-ray absorption fine structure spectroscopy. J. Am. Chem. Soc. 101, 2815–2832.Google Scholar
First citation Teo, B. K., Lee, P. A., Simons, A. L., Eisenberger, P. & Kincaid, B. M. (1977). EXAFS. Approximation, parameterization and chemical transferability of amplitude function. J. Am. Chem. Soc. 99, 3854–3856.Google Scholar
First citation Theisen, R. & Vollath, D. (1967). Tables of X-ray mass attenuation coefficients. Düsseldorf: Verlag Stahleisen.Google Scholar
First citation Thomas, L. H. (1927). The calculation of atomic fields. Proc. Cambridge Philos. Soc. 23, 542–548.Google Scholar
First citation Thompson, D. J. & Poole, M. W. (1979). Editors. European Synchrotron Radiation Facility. Suppl. II: The machine. Strasbourg: ESF.Google Scholar
First citation Thomsen, J. S. & Burr, A. F. (1968). Biography of the x-unit – the X-ray wavelength scale. Am. J. Phys. 36, 803–810.Google Scholar
First citation Tohji, K., Udagawa, Y., Kawasaki, T. & Masuda, K. (1983). Laboratory EXAFS spectrometer with a bent-crystal, a solid-state detector, and a fast detection system. Rev. Sci. Instrum. 54, 1482–1487.Google Scholar
First citation Uehling, E. A. (1935). Polarization effects in the positron theory. Phys. Rev. 48, 55–63.Google Scholar
First citation Veigele, W. J. (1973). Photon cross sections from 0.1 keV to 1 MeV for elements Z = 1 to Z = 94. At. Data, 5, 51–111.Google Scholar
First citation Victoreen, J. A. (1949). The calculation of X-ray mass absorption coefficients. J. Appl. Phys. 20, 1141–1147.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
First citation Waller, I. (1928). Über eine verallgemeinerte Streuungsformel. Z. Phys. 51, 213–231.Google Scholar
First citation Wang, M. S. (1986). Relativistic dispersion relation for X-ray anomalous scattering factor. Phys. Rev. A, 34, 636–637.Google Scholar
First citation Wang, M. S. & Pratt, R. H. (1983). Importance of bound–bound transitions in the dispersion relation for calculation of soft-X-ray forward Rayleigh scattering from light elements. Phys. Rev. A, 28, 3115–3116.Google Scholar
First citation Warren, B. E. (1968). X-ray diffraction. New York: Addison-Wesley.Google Scholar
First citation Wichmann, E. H. & Kroll, N. M. (1956). Vacuum polarization in a strong Coulomb field. Phys. Rev. 101, 843–859.Google Scholar
First citation Wilson, R. R. (1941). A vacuum-tight sliding seal. Rev. Sci. Instrum. 12, 91–93.Google Scholar
First citation Winick, H. (1980). Properties of synchrotron radiation. In Synchrotron radiation research, edited by H. Winick & S. Doniach. New York: Plenum.Google Scholar
First citation Winick, H. & Bienenstock, A. (1978). Synchrotron radiation research. Ann. Rev. Nucl. Sci. 28, 33–113.Google Scholar
First citation Winick, H. & Doniach, S. (1980). Editors. Synchrotron radiation research. New York: Plenum.Google Scholar
First citation Woodruff, D. P. (1986). Fine structure in ionization cross sections and applications to surface science. Rep. Prog. Phys. 49, 683–723.Google Scholar
First citation Yaakobi, B., Boehli, T., Bourke, P., Conturie, Y., Craxton, R. S., Delettrez, J., Forsyth, J. M., Frankel, R. D., Goldman, L. M., McCrory, L. R., Richardson, M. C., Seka, W., Shvarts, D. & Soures, J. M. (1981). Characteristics of target interaction with high-power UV laser radiation. Opt. Commun. 39, 175–179.Google Scholar
First citation Yaakobi, B., Bourke, P., Conturie, Y., Delettrez, J., Forsyth, J. M., Frankel, R. D., Goldman, L. M., McCrory, L. R., Seka, W., Soures, J. M., Burek, A. J. & Deslattes, R. D. (1981). High X-ray conversion efficiency with target irradiation by a frequency-tripled Nd:glass laser. Opt. Commun. 38, 196–200.Google Scholar
First citation Yamaguchi, T., Mitsunaga, T., Yoshida, N., Wakita, H., Fujiwara, M., Matsushita, T., Ikeda, S. & Nomura, M. (1993). XAFS study with an in situ electrochemical cell on manganese Schiff base complexes as a model of a photosystem. Jpn. J. Appl. Phys. 32, Suppl. 32-2, 533–535.Google Scholar
First citation Yao, T. (1992). Lanthanum hexaboride filament in an X-ray generator of a laboratory EXAFS facility. Rev. Sci. Instrum. 63, 2103–2104.Google Scholar
First citation Yeh, J. J. & Lindau, I. (1985). Atomic subshell photoionization cross sections and asymmetry parameters: 1 ≤ Z ≤ 103. At. Data Nucl. Data Tables, 32, 1–156.Google Scholar
First citation Yoshimatsu, M. & Kozaki, S. (1977). High-brilliance X-ray sources. X-ray optics, edited by H.-J. Queisser, Chap. 2. Berlin: Springer.Google Scholar
First citation Young, R. A. (1963). Balanced filters for X-ray diffractometry. Z. Kristallogr. 118, 233–247.Google Scholar
First citation Young, R. A. (1993). The Rietveld method. Oxford University Press.Google Scholar
First citation Zachariasen, W. H. (1945). Theory of X-ray diffraction in crystals. New York: Dover.Google Scholar








































to end of page
to top of page