research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

High-accuracy measurement of mass attenuation coefficients and the imaginary component of the atomic form factor of zinc from 8.51 keV to 11.59 keV, and X-ray absorption fine structure with investigation of zinc theory and nanostructure

crossmark logo

aSchool of Physics, University of Melbourne, Australia, bANSTO, Australian Synchrotron, Melbourne, Australia, and cLa Trobe University, Australia
*Correspondence e-mail: chantler@unimelb.edu.au

Edited by R. W. Strange, University of Essex, United Kingdom (Received 19 October 2020; accepted 8 June 2021; online 19 July 2021)

High-accuracy X-ray mass attenuation coefficients were measured from the first X-ray Extended Range Technique (XERT)-like experiment at the Australian Synchrotron. Experimentally measured mass attenuation coefficients deviate by ∼50% from the theoretical values near the zinc absorption edge, suggesting that improvements in theoretical tabulations of mass attenuation coefficients are required to bring them into better agreement with experiment. Using these values the imaginary component of the atomic form factor of zinc was determined for all the measured photon energies. The zinc K-edge jump ratio and jump factor are determined and results raise significant questions regarding the definitions of quantities used and best practice for background subtraction prior to X-ray absorption fine-structure (XAFS) analysis. The XAFS analysis shows excellent agreement between the measured and tabulated values and yields bond lengths and nanostructure of zinc with uncertainties of from 0.1% to 0.3% or 0.003 Å to 0.008 Å. Significant variation from the reported crystal structure was observed, suggesting local dynamic motion of the standard crystal lattice. XAFS is sensitive to dynamic correlated motion and in principle is capable of observing local dynamic motion beyond the reach of conventional crystallography. These results for the zinc absorption coefficient, XAFS and structure are the most accurate structural refinements of zinc at room temperature.

1. Introduction

X-ray attenuation coefficients and atomic form factors explain how X-rays interact with matter and atoms (Rontgen, 1896[Rontgen, W. (1896). Nature, 53, 274-276.]). X-ray absorption fine-structure (XAFS) spectra display the energy-dependent fine structure in the mass attenuation coefficients, sensitive to many structural details of materials (Rehr & Albers, 2000[Rehr, J. J. & Albers, R. C. (2000). Rev. Mod. Phys. 72, 621-654.]). Thus, X-ray mass attenuation coefficients, particularly across absorption edge regions, are widely used in many techniques and fields including biological (Coulthard & Sham, 1996[Coulthard, I. & Sham, T. K. (1996). Phys. Rev. Lett. 77, 4824-4827.]; Streltsov et al., 2018[Streltsov, V. A., Ekanayake, R. S., Drew, S. C., Chantler, C. T. & Best, S. P. (2018). Inorg. Chem. 57, 11422-11435.]), molecular, condensed matter and solid state fields (Palmberg & Rhodin, 1968[Palmberg, P. W. & Rhodin, T. N. (1968). J. Appl. Phys. 39, 2425-2432.]; Bohic et al., 2001[Bohic, S., Simionovici, A., Snigirev, A., Ortega, R. G., Devès, D., Heymann, D. & Schroer, C. G. (2001). Appl. Phys. Lett. 78, 3544-3546.]) to study structural parameters including bond lengths, thermal parameters, oxidation number, coordination number and electron inelastic mean free paths of materials and molecules (Ignatov et al., 2001[Ignatov, A. Y., Ali, N. & Khalid, S. (2001). Phys. Rev. B, 64, 014413.]; Han et al., 2002[Han, S., Stern, E., Haskel, D. & Moodenbaugh, A. R. (2002). Phys. Rev. B, 66, 094101.]; Islam et al., 2015[Islam, M. T., Chantler, C. T., Cheah, M. H., Tantau, L. J., Tran, C. Q. & Best, S. P. (2015). J. Synchrotron Rad. 22, 1475-1491.], 2016[Islam, M. T., Best, S. P., Bourke, J. D., Tantau, L. J., Tran, C. Q., Wang, F. & Chantler, C. T. (2016). J. Phys. Chem. C, 120, 9399-9418.]). Attenuation coefficients and atomic form factors are important in the study of the density of electron states (Joly et al., 1999[Joly, Y., Cabaret, D., Renevier, H. & Natoli, C. R. (1999). Phys. Rev. Lett. 82, 2398-2401.]; Hossain et al., 2005[Hossain, F. M., Riley, D. P. & Murch, G. E. (2005). Phys. Rev. B, 72, 235101.]), roughness of sample surfaces (Glover et al., 2009[Glover, J. L., Chantler, C. T. & de Jonge, M. D. (2009). Phys. Lett. A, 373, 1177-1180.]), dynamics of atomic processes, elastic scattering and other scattering of X-ray interactions with matter (Hopersky et al., 2004[Hopersky, A., Petrov, I., Nadolinsky, A., Yavna, V. & Koneev, R. (2004). J. Phys. B At. Mol. Opt. Phys. 37, 3313-3319.]) and self-absorption of materials (Trevorah et al., 2019[Trevorah, R. M., Chantler, C. T. & Schalken, M. J. (2019). IUCrJ, 6, 586-602.]). Mass attenuation across the XAFS can be used to confirm or determine near-neighbour coordination lengths of materials (Sayers et al., 1972[Sayers, D. E., Lytle, F. W. & Stern, E. A. (1972). J. Non-Cryst. Solids, 8-10, 401-407.]; Glover et al., 2010[Glover, J. L., Chantler, C. T., Barnea, Z., Rae, N. A. & Tran, C. Q. (2010). J. Phys. B At. Mol. Opt. Phys. 43, 085001.]).

Many improvements and developments in analytic techniques and experiment have been introduced in XAFS analysis including the XERT protocol for high-accuracy data with uncertainties below 0.1% and removal of systematic errors (Chantler, 2009[Chantler, C. T. (2009). Eur. Phys. J. Spec. Top. 169, 147-153.]). The XAFS spectrum consists of oscillations at energies just after the absorption edge due to the interference of the photoelectron wave with the neighbouring electrons (Joly, 2001[Joly, Y. (2001). Phys. Rev. B, 63, 125120-125130.]). The oscillations can be investigated using FEFF (Rehr et al., 1991[Rehr, J. J., Mustre de Leon, J., Zabinsky, S. I. & Albers, R. C. (1991). J. Am. Chem. Soc. 113, 5135-5140.]; Rehr & Albers, 2000[Rehr, J. J. & Albers, R. C. (2000). Rev. Mod. Phys. 72, 621-654.]) and eFEFFIT (Smale et al., 2006[Smale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559-1563.]; Schalken & Chantler, 2018[Schalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920-934.]), from modified IFEFFIT (Newville, 2001[Newville, M. (2001). J. Synchrotron Rad. 8, 96-100.]) analysis propagating robust uncertainties of XAFS data over the entire energy spectrum. The implementation of the XERT-like experiment, the first in Australia, is presented by Ekanayake et al. (2021[Ekanayake, R. S. K., Chantler, C. T., Sier, D., Schalken, M. J., Illig, A. J., de Jonge, M. D., Johannesen, B., Kappen, P. & Tran, C. Q. (2021). J. Synchrotron Rad. 28, 1476-1491.]), hereafter called Part 1.

This paper presents X-ray mass attenuation coefficients and derivations of the imaginary component of the atomic form factor of zinc to a high accuracy. Experimental details and systematic corrections are provided in Part 1. The photoelectric mass absorption coefficient is also calculated from the mass attenuation coefficient after subtracting the scattering contribution. These measurements can be used as standards in XAFS and are also useful in crystallography and medical fields. These high-accuracy data permit direct access to critical measurements of electron inelastic mean free paths (IMFPs) and other key parameters. Discrepancies between the tabulated and the high-accuracy experimental values are investigated. X-ray mass attenuation coefficients are calculated using relativistic quantum mechanics under several assumptions. XCOM (Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]) and FFAST (Chantler, 1995[Chantler, C. T. (1995). J. Phys. Chem. Ref. Data, 24, 71-643.], 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]) are the two theoretical tabulations recommended by the National Institute of Standards and Technology (NIST). XCOM tabulated unrenormalized values with Hartree–Slater non-relativistic wavefunctions (Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]) and renormalized values with relativistic Hartree–Fock potential (Scofield, 1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]) are compared with experiment. The use of renormalization has been an ongoing discussion in the field.

The edge jump ratio r is the ratio of the mass photoabsorption coefficient just above the absorption edge to the mass photoabsorption coefficient just below the absorption edge (McMaster et al., 1969[McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.]; Veigele, 1973[Veigele, W. J. (1973). At. Data Nucl. Data Tables, 5, 51-111.]; Broll, 1986[Broll, N. (1986). X-ray Spectrom. 15, 271-285.]; Bennal & Badiger, 2007[Bennal, A. S. & Badiger, N. M. (2007). J. Phys. B At. Mol. Opt. Phys. 40, 2189-2199.]; Kaya et al., 2011[Kaya, N., Apaydin, G. & Tirasoglu, E. (2011). Radiat. Phys. Chem. 80, 677-681.]). Investigation of the experimental edge jump J, or JK for the K-edge, and the edge jump ratio r, or rK here, is important for better understanding of absorption spectra since it checks the predictions of theory which are often used from tabulations for many applications. The edge jump is generally important for radiology, shielding, X-ray filters, reference materials, X-ray fluorescence (Budak et al., 2003[Budak, G., Karabulut, A. & Ertuǧrul, M. (2003). Radiat. Meas. 37, 103-107.]) and in particular applications for individual elements or materials including, for example, characterization of dopants in superconductors (Mallikarjuna et al., 2002[Mallikarjuna, M. L., Appaji Gowda, S. B., Gowda, R. & Umesh, T. K. (2002). Radiat. Phys. Chem. 65, 217-223.]). In many cases oxides or other compounds have been used which complicated any analysis. The zinc K-edge jump ratio and jump factor raise significant questions about how to define the edge and the background subtraction prior to XAFS analysis.

XAFS analysis was performed on the above calculated measurements and the bond lengths of zinc were determined. These high-accuracy values can be used as a baseline to discriminate different IMFP theories for zinc data.

2. The mass attenuation coefficients and imaginary component of the atomic form factor

The mass attenuation coefficients were obtained by considering the mean of the results from measurements using different foils and apertures. Evaluated values of mass attenuation coefficients with the uncertainties along the interested energy region are plotted in Fig. 1[link].

[Figure 1]
Figure 1
Calculated mass attenuation coefficient of zinc (a) over the energy range 8.51 keV to 11.59 keV; (b) covering the edge and XAFS region; (c) in the central XAFS region; and (d) absolute and relative percentage uncertainties. The zinc K-absorption edge is observed at 9.66 keV and the associated XAFS lies between 9.66 keV and 10.10 keV.

The imaginary component of the form factor can be derived from the photoelectric mass absorption coefficients [[{{\mu}/{\rho}}]_{\rm{pe}},]

[f_{2} = {{EuA[{{\mu}/{\rho}}]_{\rm{pe}}} \over {2hcR_{\rm{e}}}}, \eqno(1)]

where E is the energy, u is the atomic mass unit, A is the atomic mass, Re is the classic electron radius, h is the plank constant and c is the speed of light. The photoelectric absorption is the dominant contribution for the total mass attenuation and it can be derived by subtracting the attenuation contributions due to Coherent (Rayleigh) and incoherent (Compton) scatterings. These scattering contributions [[{{\mu}/{\rho}}]_{\rm{R+C}}] were obtained from the FFAST calculations.

Additional uncertainties were estimated to be the half of the discrepancy between the XCOM and FFAST tabulated [[{{\mu}/{\rho}}]_{\rm{R+C}}] values. The scattering uncertainties are always less than 0.036% of the photoelectric absorption coefficient. The uncertainty of f2 is

[\sigma_{f_{2}} = {{EuA}\over{2hcR_{\rm{e}}}} \left(\,\sigma_{[\mu/\rho]}^{2}+\Delta_{\rm{R+C}}^{2}\,\right)^{1/2}. \eqno(2)]

This work demonstrates the attenuation measurements of zinc over an extended energy range including K-edge XAFS spectrum from XERT and solid-state properties of the sample on an absolute scale. The energy-dependent mass attenuation coefficients from 34.77 to 327.76 (cm2 g−1) were calculated over the energy range from 8.51 keV to 11.5914 keV. The absolute experimental uncertainties range from 0.023% to 0.036%, limited by various experimental systematic errors such as energy drifts during the monochromator tuning, monochromator glitches, and ion chamber current fluctuations.

Quantifying uncertainties is crucial in identifying any data affected by any systematics. Often badly affected points are deleted by researchers; yet a quantitative uncertainty including variance can validate them with an appropriate weighting from the inconsistency. The imaginary component of the atomic form factor was determined to be from 0.492 (e atom−1) to 4.898 (e atom−1). Uncertainty contributions and the dominant systematic corrections in the current work are summarized in Table 1[link]. Table S1 of the supporting information provides: the measured mass attenuation coefficients [[{{\mu}/{\rho}}]_{\rm{pe}}] or [[{{\mu}/{\rho}}]_{\rm{tot}}] as a function of energy with energy E and estimated energy uncertainty σE; the total (or absolute) estimated percentage uncertainties [\sigma_{[{{\mu}/{\rho}}]_{\rm{tot}}}\%] and estimated relative percentage uncertainties [\sigma_{[{{\mu}/{\rho}}]_{\rm{rel}}}\%]; the mass photoabsorption coefficient [[{{\mu}/{\rho}}]_{\rm{pe}}\,({{{\rm{cm}}^{2}}\,{\rm{g}}^{-1}})] with percentage uncertainties [\sigma_{[{{\mu}/{\rho}}]_{\rm{pe}}}\%], particularly after subtraction of the estimated scattering coefficients (Rayleigh and Compton), are provided in later columns, together with the estimated imaginary component of the atomic form factor f2 (e atom−1).

Table 1
Uncertainty and variation of mass attenuation coefficients during processing at several systematic correction stages; and the magnitude of specific systematic corrections

Contributions to measurements are labelled [[{{\mu}/{\rho}}]_{\rm{rel}}] if they particularly contribute to the relative structure of adjacent points, the edge shape or the XAFS; or [[{{\mu}/{\rho}}]_{\rm{abs}}] if they primarily scale all values with a slowly varying function. Hence there are two final uncertainties – relating to the absolute value of the mass attenuation coefficient [[{{\mu}/{\rho}}]] and relating to the point-wise and local structure, for example, for XAFS analysis which can be denoted [[{{\mu}/{\rho}}]_{\rm{rel}}].

Quantity Magnitude of correction [[{{\mu}/{\rho}}]] Uncertainty and variance [\sigma_{[{{\mu}/{\rho}}]}] Comments
[[{{\mu}/{\rho}}]_{\rm{rel}}]   <0.319% Variancea
  22–536% (10 µm foil) <1.853% (10 µm) Blank normalizationb
  Up to 97% (50 µm foil) <0.133% (100 µm)  
  Up to 57% ±15% (100 µm) 0.00039–1.46% Dark currentc
  Up to 1.31% (10 µm)    
(total)   <0.042% Total variance after correctionsd
 
  1–5 × 10−3%   Harmonic correctione
 
(50 µm) <14.2% <10.5% Fluorescence correctionf
(25 µm) <0.101% <0.003%  
(10 µm) <0.0123% <0.0003%  
(total) <0.139% <0.028%  
 
(10 µm) <2.52% <0.02% Roughnessg
(25 µm) <0.815% <0.015%  
(total) <1.56% <0.0037%  
 
(50 µm) <9.89% <0.239% Bandwidthh
(25 µm) <4.91% <0.119%  
(10 µm) <1.703% <0.041%  
(total) <7.24% <0.0037%  
 
[[{{\mu}/{\rho}}]_{\rm{abs}}]   <0.037% Nominal thicknessesi
  3.55–7.60% 0.000018–0.024237% Average ICDj
  0.374–7.606% <0.024% Full-foil mapk
E (keV) −1 to +3 eV, ±1–3 eV <0.0038% Energyl
 
Quantity Magnitude range (cm2 g−1) Uncertainty range [\sigma_{[{{\mu}/{\rho}}]}] Comments
[[{{\mu}/{\rho}}]_{\rm{rel}}] 34.765–325.321 0.000677–0.027% After systematic correctionsm
[[{{\mu}/{\rho}}]_{\rm{abs}}] 34.765–327.760 0.023–0.0357% After normalizing to absolute thickness from full-foil mapn
aStandard errors from counting statistics (variance including precision before systematic corrections). bBlank correction and net uncertainty (Part 1, §3). Large for thin foils. cDark current correction and net uncertainty (±0.5 counts s−1) (Part 1, §3). dTotal dispersion of measurement precision after above corrections. eHarmonic coefficient and contribution is very small here (Part 1, §7). fSecondary photons from fluorescent scattering. Correction large for 50 µm sample and directly above absorption edges, 0 below Zn edge (Part 1, §6). Maximum uncertainty only applies to 50 µm. gEffect of roughness is greatest for 10 µm sample when attenuation is large (Part 1, §8). hBandwidth correction greatest for 50 µm sample along the edge where d[μ/ρ]/dE is greatest. iUse of nominal thickness and corresponding uncertainty (Part 1, §4,5). jUse of local integrated column density and corresponding uncertainty (Part 1, §5). kAbsolute accuracy of the full-foil mapping technique (Part 1, §4). lError in energy calibration data. Correction minimal at the absorption edge (Part 1, §9). mRelative measurements and uncertainties after correcting for systematics. nAbsolute measurements and uncertainties after normalizing to absolute thickness with full-foil map.

Supplementary documents are provided including the data of Table S1 in text format (see ZnTable2.dat in the supporting zip file). The format of [[{{\mu}/{\rho}}]_{\rm{tot}}\,({{{\rm{cm}}^{2}}\,{\rm{g}^{-1}}})] versus E with uncertainties (Zn_K_muoverrhovsE.dat) provides input for efeffit (Smale et al., 2006[Smale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559-1563.]; Schalken & Chantler, 2018[Schalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920-934.]), IFEFFIT (Newville, 2001[Newville, M. (2001). J. Synchrotron Rad. 8, 96-100.]), Athena (Ravel & Newville, 2005[Ravel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537-541.]) and particularly mu2chi (Schalken & Chantler, 2018[Schalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920-934.]). Corresponding files are represented as a template for both a cif standard (Zn_K.cif) and χ versus k with uncertainties (Zn_K_eFEFFIT.dat) data for an efeffit, IFEFFIT and Athena input standard.

3. Comparison with theory

Figure 2[link] compares our measured mass attenuation coefficients with FFAST tabulated values (Chantler, 1995[Chantler, C. T. (1995). J. Phys. Chem. Ref. Data, 24, 71-643.], 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]), XCOM tabulated values with Hartree–Slater non-relativistic wavefunctions including Scofield relativistic perturbations (Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]) (unrenomalized) and also renormalized values in a relativistic Hartree–Fock potential as defined by Scofield (1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]). There is an agreement in experimental results with FFAST values in the range of 8.5 keV to the edge energy and above 12 keV. The deviation of calculated values from the FFAST values begins at around the edge energy (9.66 keV) and decreases and converges to a 1% discrepancy above 12 keV (see Fig. 1 of Part 1).

[Figure 2]
Figure 2
The percentage discrepancy between the derived mass attenuation coefficient and the tabulated FFAST values (circles) (Chantler, 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]) over the energy range 8.5 keV to 11.59 keV, and the percentage discrepancies between XCOM tabulated values with Hartree–Slater calculations (crosses) (Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]), Scofield tabulated values renormalized with Hartree–Fock calculations (stars) (Scofield, 1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]) and Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) measurements (squares). The near-edge region has a large discrepancy from current measurements due to the solid-state XAFS structure and the triangle effect. Renormalized Scofield (1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]) values deviate dramatically everywhere, implying that the earlier belief that correct theory lies between the unrenormalized and the renormalized predictions is not valid for these data.

At lower energies the variation of our experimental measurements is smooth (Fig. 1[link]). The deviation from the XCOM values also begins at the edge with 5–10% uncertainty or error. Above the absorption edge, the experimental values deviate up and down significantly due to the oscillatory XAFS as expected. There is a dramatic discrepancy of XCOM theory with the experimental measurement at the white line, likely due to the inaccuracy of the edge energy predicted by XCOM. Whilst the experimental claimed accuracy is much better than this, the theoretical claimed uncertainty far above the (K) edge is about 1%, as observed. Perhaps surprisingly, given other high-accuracy results, XCOM has a 2–5% discrepancy from experimental results below the edge, favouring FFAST; though the two theoretical predictions are in very good agreement with one another above about 10.3 keV. Both disagree with both high-accuracy experiments.

Scofield (1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]) discussed and provided two theoretical predictions for atomic absorption coefficients – the unrenormalized values in XCOM and another renormalized set, obtained by renormalizing the cross sections by the amplitude of the potential at the nucleus, in a relativistic Hartree–Fock potential compared with XCOM's Hartree–Slater potential. The decision as to whether or not to perform this renormalization has been an ongoing discussion in the field. Some previous investigations have favoured the unrenormalized results (Saloman & Hubbell, 1987[Saloman, E. & Hubbell, J. (1987). Nucl. Instrum. Methods Phys. Res. A, 255, 38-42.]), whilst some have claimed that the difference in the estimates should be used as an uncertainty, and that the correct value should lie between the two approaches (Gerward, 1992[Gerward, L. (1992). Nucl. Instrum. Methods Phys. Res. B, 69, 407-412.]).

Those comparisons were made with significantly less accurate experimental data than that presented in this work, making a definitive resolution to this discussion impossible at that time. At all energies, XCOM [using Hartree–Slater (HS) non-relativistic wavefunctions] is more consistent with experimental data than the renormalized data referencing Hartree–Fock (HF) calculations. At lower energies, the discrepancy from the current experimental results is about 3% for XCOM data with HS calculations whereas data with HF calculations have a (much) higher (12%) discrepancy below 9.6 keV. This current evidence justifies the less-fully relativistic unrenormalized HS results [XCOM] over the HF renormalized data. Note that FFAST is a better relativistic and self-consistent model and in turn appears in better agreement with the experimental data.

The experimentally calculated data deviate from the theory by about 10–20% at the edge, 2.5% after the edge and 1–2% far above and below the edge. About 70% of the experimentally calculated data are in agreement (within 1–2σ) with FFAST. The XAFS oscillations near the edge are absent in theory because the theory was developed as a free-atom model and so deviations are therefore large near the edge energy. More importantly, the near-edge region shows a generally triangular discrepancy quite separate from the oscillatory solid state effects. This is a general observation for many materials such as silver (Tran et al., 2005[Tran, C. Q., Chantler, C. T., Barnea, Z., Jonge, M. D., Dhal, B. B., Chung, C. T., Paterson, D. & Wang, J. (2005). J. Phys. B At. Mol. Opt. Phys. 38, 89-107.]; Islam et al., 2014[Islam, M. T., Tantau, L. J., Rae, N. A., Barnea, Z., Tran, C. Q. & Chantler, C. T. (2014). J. Synchrotron Rad. 21, 413-423.]; Tantau et al., 2015[Tantau, L. J., Chantler, C. T., Bourke, J. D., Islam, M. T., Payne, A. T., Rae, N. A. & Tran, C. Q. (2015). J. Phys. Condens. Matter, 27, 266301.]), molybdenum (de Jonge et al., 2005[Jonge, M. D. de, Tran, C. Q., Chantler, C. T., Barnea, Z., Dhal, B. B., Cookson, D. J., Lee, W. & Mashayekhi, A. (2005). Phys. Rev. A, 71, 032702.]; Chantler & Bourke, 2010[Chantler, C. T. & Bourke, J. D. (2010). J. Phys. Chem. Lett. 1, 2422-2427.]), tin (de Jonge et al., 2007[Jonge, M. D. de, Tran, C. Q., Chantler, C. T., Barnea, Z., Dhal, B. B., Paterson, D., Kanter, E. P., Southworth, S. H., Young, L., Beno, M. A., Linton, J. A. & Jennings, G. (2007). Phys. Rev. A, 75, 032702.]), copper (Chantler et al., 2001[Chantler, C. T., Tran, C. Q., Barnea, Z., Paterson, D., Cookson, D. J. & Balaic, D. X. (2001). Phys. Rev. A, 64, 062506.]; Glover et al., 2008[Glover, J. L., Chantler, C. T., Barnea, Z., Rae, N. A., Tran, C. Q., Creagh, D. C., Paterson, D. & Dhal, B. B. (2008). Phys. Rev. A, 78, 052902.]) and gold (Glover et al., 2010[Glover, J. L., Chantler, C. T., Barnea, Z., Rae, N. A. & Tran, C. Q. (2010). J. Phys. B At. Mol. Opt. Phys. 43, 085001.]; Islam et al., 2010[Islam, M. T., Rae, N. A., Glover, J. L., Barnea, Z., de Jonge, M. D., Tran, C. Q., Wang, J. & Chantler, C. T. (2010). Phys. Rev. A, 81, 022903.]). Improvements in theoretical understanding are definitely required in the future especially for the baseline.

Data from Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) are compared with these results over the common energy range, as it has the highest previous claimed accuracy of 0.044% to 0.197%. Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) performed an X-ray absorption spectroscopy (XAS) experiment for zinc material on a different beamline using a similar technique. However, these detailed approaches in analysis were quite different. At lower energies our current results and Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) are in excellent agreement (1σ). Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) provide no XAFS and the spacing is unsuitable for any detailed nanostructure determination. However, it ought to be useful for an absolute measurement of the attenuation. These experimental results are in fair agreement with the earlier high-accuracy results to about 0.1% or 1–3σ of the earlier high-accuracy data set. The signature for the deviation can be investigated from the original paper, and may be due to harmonic, roughness or bandwidth in either analysis. Blank and dark-current normalization were performed in both experiments and the pattern of discrepancy does not match these possible sources of systematic error. The energy calibration could provide this signature noting that the absence of XAFS and near-edge structure prevents the confirmation of this systematic or the confirmation of its absence. Perhaps the most likely cause is a systematic error of the previous analysis of scattering which decreases from 3% to 6% just above the edge to 1–2% or 2–4% around 11.5 keV. The current work corrected for significant direct scattering but also for significant background scattering whereas the earlier work only investigated direct scattering (Ekanayake et al., 2021[Ekanayake, R. S. K., Chantler, C. T., Sier, D., Schalken, M. J., Illig, A. J., de Jonge, M. D., Johannesen, B., Kappen, P. & Tran, C. Q. (2021). J. Synchrotron Rad. 28, 1476-1491.]).

Other existing experimental data of Unonius & Suortti (1989[Unonius, L. & Suortti, P. (1989). J. Appl. Cryst. 22, 46-52.]) and Hopkins (1959[Hopkins, J. I. (1959). J. Appl. Phys. 30, 185-187.]) were compared in Fig. 1 of Part 1. Due to the lack of correction for numerous systematics, it is not surprising that discrepancies from current experiment and theory are 10–13%. These measurements are also sparse and provide no XAFS to confirm energy analysis. High-accuracy measurements with absolute uncertainties can therefore be used in understanding theoretical tabulations and limiting atomic physics modelling and nanostructural models. The derived mass attenuation coefficients are useful in solid-state and atomic fields in investigating many-body effects such as core-hole lifetime, interatomic interactions, threshold energy and thermal effects.

4. X-ray absorption edge jumps and edge jump ratios

The mass attenuation coefficient is related to the total attenuation cross section σtot (barns atom−1) by

[\sigma_{\rm{tot}} = \left[{{\mu}\over{\rho}}\right] \, {{A}\over{N_{\rm{A}}}}, \eqno(3)]

where NA is the Avogadro number and A is the atomic weight. The difference between the mass photoabsorption coefficient above and below the edge gives the K-shell mass photoabsorption coefficient and then the jump ratio is

[r_{K} = {{\tau_{K}+\tau_{L}+\tau_{M}+\ldots}\over{\tau_{L}+\tau_{M}+\ldots}} = {{ [{{\mu}/{\rho}}]_{\rm{pe,\,above\,edge}} }\over{ [{{\mu}/{\rho}}]_{\rm{pe,\,below\,edge}} }}, \eqno(4)]

where [\tau_{i}] = [[{{\mu}/{\rho}}]_{{\rm{pe}},i}] (cm2 g−1) is the photoelectric absorption coefficient for the shell i. The K-shell jump factor is then defined as (Latyshev, 1947[Latyshev, G. D. (1947). Rev. Mod. Phys. 19, 132-145.]; Schmickley & Pratt, 1967[Schmickley, R. D. & Pratt, R. H. (1967). Phys. Rev. 164, 104-116.])

[\eqalignno{ J_{K} & = {{r_{K}-1} \over {r_{K}}} = {{\tau_{K}}\over{\tau_{K}+\tau_{L}+\tau_{M}+\ldots_{\vphantom{\big{|}}}}} \cr& = {{[{{\mu}/{\rho}}]_{\rm{pe,\,above\,edge}} - [{{\mu}/{\rho}}]_{\rm{pe,\,below\,edge}}}\over{[{{\mu}/{\rho}}]_{\rm{pe,\,above\,edge}} }}. &(5)}]

Four main methods for calculating the edge jump ratio and jump factor are reported:

(1) A gamma-ray attenuation method, where gamma-rays are attenuated through a sample and the attenuation coefficient is determined (Puttaswamy et al., 1981[Puttaswamy, K. S., Gowda, R. & Sanjeevaiah, B. (1981). Can. J. Phys. 59, 853-858.]; Ouseph et al., 1982[Ouseph, P. J., Hoskins, K. D., Berman, J. I. & Bolander, A. J. (1982). Am. J. Phys. 50, 275-276.]; Mallikarjuna et al., 2002[Mallikarjuna, M. L., Appaji Gowda, S. B., Gowda, R. & Umesh, T. K. (2002). Radiat. Phys. Chem. 65, 217-223.]). The early key idea was to have a radioactive source with lines on either side of the absorption edge and use the relative attenuation to determine the jump ratio and jump factor for the edge. Uncertainties of order 5% are reported; however, in this approach, the energies of the two lines are often significantly (very) far apart, e.g. 7 keV or 12 keV or 20 keV away from the edge; and hence interpretation of the edge jump is seriously limited by the extrapolation required.

(2) A Compton peak attenuation method, where the broad Compton-scattered spectral photons are allowed to attenuate through an elemental solid to measure attenuation coefficients above and below the edge for the continuum spectrum (Ayala & Mainardi, 1996[Ayala, A. P. & Mainardi, R. T. (1996). Radiat. Phys. Chem. 47, 177-181.]; Budak & Polat, 2004[Budak, G. & Polat, R. (2004). J. Quant. Spectrosc. Radiat. Transfer, 88, 525-532.]; Polat et al., 2004[Polat, R., İçelli, O. & Budak, G. (2004). Anal. Chim. Acta, 505, 307-314.]). Measurement step sizes are typically 50 eV or 58 eV. Uncertainties are estimated as 1.5%, although discrepancies of 5% are observed.

In methods (1) and (2), no influence of scattering or fluorescence on the measurement is normally corrected for. This introduces a somewhat ill-defined uncertainty and systematic error. In general, this is similar to replacing [[{{\mu}/{\rho}}]_{\rm{pe}}] with [[{{\mu}/{\rho}}]], so that typically the numerator and denominator of equation (4)[link] are increased by a scattering offset; and the denominator of equation (5)[link] is increased by the same amount. In both methods (1) and (2), normally only one experimental sample thickness is measured. Although in an ideal situation this might yield a correct result, this manuscript and Part 1 prove that the thickness dependence of systematics is quite significant.

(3) An energy-dispersive X-ray fluorescence (EDXRF) method, where X-ray photons are allowed to pass through a sample and then the attenuation is measured (Ertugrul et al., 2002[Ertugrul, M., Karabulut, A. & Budak, G. (2002). Radiat. Phys. Chem. 64, 1-3.]; Budak et al., 2003[Budak, G., Karabulut, A. & Ertuǧrul, M. (2003). Radiat. Meas. 37, 103-107.]; Polat et al., 2005[Polat, R., Budak, G., Gürol, A., Karabulut, A. & Ertugřul, M. (2005). Radiat. Meas. 39, 409-415.]; Bennal & Badiger, 2007[Bennal, A. S. & Badiger, N. M. (2007). J. Phys. B At. Mol. Opt. Phys. 40, 2189-2199.]; Kaya et al., 2011[Kaya, N., Apaydin, G. & Tirasoglu, E. (2011). Radiat. Phys. Chem. 80, 677-681.]; Singh Sidhu et al., 2011[Singh Sidhu, B., Dhaliwal, A. S., Mann, K. S. & Kahlon, K. S. (2011). Radiat. Phys. Chem. 80, 28-32.]). This method derives the jump factor using the fairly complex expression

[J_{K} = {{ \sigma_{K\alpha} }\over{ \left(\sigma_{\rm{tot}}-\sigma_{\rm{ts}}\right)\,\omega_{K} }} \, \left(1+{{I_{K\beta}}\over{I_{K\alpha}}}\,\right). \eqno(6)]

Whilst this equation quotes the cross sections in barns atom−1, this is exactly equivalent to using the mass attenuation coefficients in cm2 g−1. The method measures the intensity in fluorescence of the two characteristic radiations IKα, IKβ; and separately measures the mass attenuation coefficients. The method requires: that the self-absorption of the two characteristic radiations (Kα, Kβ) are computed in fluorescence using theoretical mass attenuation coefficients (Trevorah et al., 2019[Trevorah, R. M., Chantler, C. T. & Schalken, M. J. (2019). IUCrJ, 6, 586-602.]); that the K-shell fluorescence yield ωK is taken from an available tabulation (e.g. Broll, 1986[Broll, N. (1986). X-ray Spectrom. 15, 271-285.]; Bambynek et al., 1972[Bambynek, W., Crasemann, B., Fink, R. W., Freund, H. U., Mark, H., Swift, C. D., Price, R. E. & Rao, P. V. (1972). Rev. Mod. Phys. 44, 716-813.]); that the efficiencies of the detector at the two characteristic energies are known; and that separately the scattering coefficients σts are taken from a suitable tabulation (e.g. Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]; Chantler, 1995[Chantler, C. T. (1995). J. Phys. Chem. Ref. Data, 24, 71-643.]). Typical uncertainties are estimated as 3–8%, yet the uncertainties in the contributing theoretical corrections are significant. Scattering may or may not be corrected for in the extracted numerator and denominator.

A variation of this method produces secondary characteristic Kα, Kβ spectra from a radioactive source incident on several targets. Spectra are then measured using an EDXRF spectrometer to measure the attenuation coefficient of the sample across the energy resolution of the detector with good statistics in the region of the characteristic radiation (Kaya et al., 2007[Kaya, N., Tıraşoğlu, E., Apaydın, G., Aylıkcı, V. & Cengiz, E. (2007). Nucl. Instrum. Methods Phys. Res. B, 262, 16-23.]). Without correction for scattering, this will measure ratios based on attenuation coefficients rather than photoelectric coefficients. Kaya et al. (2011[Kaya, N., Apaydin, G. & Tirasoglu, E. (2011). Radiat. Phys. Chem. 80, 677-681.]) comment that the absorption curves are often distorted above the edge leading to major challenges in defining a clear attenuation or absorption coefficient (XAFS); and similarly that scattering is quite significant at these energies and will generally lead to ratios of attenuation coefficient being measured instead of photoabsorption.

(4) A Bremsstrahlung transmission method, where a continuous bremsstrahlung radiation is allowed to pass through a sample to determine a sudden drop in intensity (Nayak & Badiger, 2006[Nayak, S. V. & Badiger, N. M. (2006). J. Phys. B At. Mol. Opt. Phys. 39, 2893-2900.]) across the edge, and then used for measuring the absorption jump ratio and factors. In principle, this is a direct attenuation measurement with potentially fine spacing (50–600 eV) around the edge, and by fitting the above-edge and below-edge region (neglecting XAFS) one can derive the relevant attenuation coefficients, from which the photoelectric absorption coefficients can be extracted.

We, however, measure the zinc K-edge jump, jump ratio and jump factor from the current data herein and hence use an alternative XERT-like method. Synchrotron X-rays are allowed to pass through the sample to measure attenuation coefficients, from which photoabsorption coefficients and edge-jump ratios can be derived (Chantler et al., 2001[Chantler, C. T., Tran, C. Q., Barnea, Z., Paterson, D., Cookson, D. J. & Balaic, D. X. (2001). Phys. Rev. A, 64, 062506.]; Rae et al., 2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]; Sier et al., 2020[Sier, D., Cousland, G. P., Trevorah, R. M., Ekanayake, R. S. K., Tran, C. Q., Hester, J. R. & Chantler, C. T. (2020). J. Synchrotron Rad. 27, 1262-1277.]). Following Fig. 3[link], we initially removed the pre-edge background. Then we investigated the structure of the oscillation, defined spline and removed the spline. Far above the edge, the theory and experiment presented here are in good agreement. A key issue is the challenge to extrapolate the curves from below and from above the edge to define or extract the edge jump and the relevant cross sections.

[Figure 3]
Figure 3
As with XAFS analysis, the slope of the pre-edge is extrapolated to the edge energy. The value of the spline through the above-edge data at the (above-)edge position is taken as the above-edge value. Therefore the XAS structure above the edge is fitted to a low-order piece-wise continuous polynomial.

The `triangle effect' implies that near the edge both XCOM and FFAST underestimate the average experimental attenuation coefficient. FFAST performs better than XCOM. However, both theoretical predictions continue to increase towards the edge from above the edge, roughly as (EEK)−3; however, the imputed spline from experimental data passing through the XAFS and XANES structure curves downwards towards the edge. The correct theoretical prediction, or equivalently the correct form for the spline near the edge, is currently not yet known.

The theoretical predictions are well defined and have low uncertainty per se. However, there is likely a discrepancy separate from the complexity of the XAFS structure in matching the edge jump ratios and edge jump factors.

On the experimental side above the edge for many materials including ideal metals and elemental solids, the selection of one particular spline and form can easily change the above-edge estimate of the attenuation coefficient by about 3 cm2 g−1 in this case or by about 1% in general.

Below the edge the issue of extrapolating to the edge whilst avoiding any pre-edge structure or indeed the beginning of the edge is a challenge even if very fine point spacing is used, so we estimate an uncertainty from this of about 0.3 cm2 g−1 in this case or perhaps 1% in general. Therefore both the experimental and the theoretical uncertainty of [[{{\mu}/{\rho}}]_{\rm{tot}}] (E < EK), `A', and [[{{\mu}/{\rho}}]_{\rm{tot}}] (E > EK), `B', in Table 2[link] can be estimated by the difference between, for example, two estimates from different splines or different extrapolation methods, or different theoretical predictions. In Table 2[link] we have propagated the uncertainties assuming that the cause of a systematic in `A' or `B' are uncorrelated. In expressions involving `A'/`A', for example, the correlated uncertainty cancels. The statistical propagation of uncertainty neglecting this potential systematic (i.e. uncorrelated uncertainty) is a full factor of ten smaller (±0.012 for rK and ±0.0015 for JK).

Table 2
Derivation of above- and below-edge ratios based upon [[{{\mu}/{\rho}}]_{\rm{tot}}] or [[{{\mu}/{\rho}}]] (see text)

Source [[{{\mu}/{\rho}}]_{\rm{tot}}] (E < EK) [[{{\mu}/{\rho}}]_{\rm{tot}}] (E > EK) Jump rK,tot JK,tot
  `B' (cm2 g−1) `A' (cm2 g−1) `A' − `B' (cm2 g−1) `A'/`B' (`A' − `B')/`A'
XCOMa 35.098 253.951 218.853 7.236 0.862
FFASTb 33.810 260.906 227.096 7.717 0.8704
This work 33.369 269.991 236.622 8.091 0.8764
Estimate minimum σ 0.3 3.0 3.0 0.116 0.0148
 
Zn Z = 30c       7.602 0.868d
Zneg       7.692 0.870
Znfg       7.547 0.867
Znbg       7.592 0.868
Znhg       7.103 0.859
Zng       7.633 0.869
aXCOM (Scofield, 1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]; Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]). bFFAST (Chantler, 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]). cMcMaster et al. (1969[McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.]). dBroll (1986[Broll, N. (1986). X-ray Spectrom. 15, 271-285.]). eStorm & Israel (1970[Storm, E. & Israel, I. H. (1970). At. Data Nucl. Data Tables, 7, 565-681.]). fGerward et al. (2001[Gerward, L., Guilbert, N., Bjørn Jensen, K. & Levring, H. (2001). Radiat. Phys. Chem. 60, 23-24.]). gSingh Sidhu et al. (2011[Singh Sidhu, B., Dhaliwal, A. S., Mann, K. S. & Kahlon, K. S. (2011). Radiat. Phys. Chem. 80, 28-32.]). hRae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]).

We then compare the data directly from XCOM and FFAST with that from this work. Discrepancies are real and yet there is a good consensus. Comparing with the literature for Zn edge jump ratios and edge jump factors in the literature reveals some good consistency. However, even for the stable theory with no XAFS structure, reference a should be identical to reference f (see footnote of Table 2[link]) and is at least an underestimate of the associated error in extrapolation for the theory. Above the edge, XCOM is anomalously low compared with experiment or FFAST; below the edge it is anomalously high; hence the edge jump ratio is anomalously low.

There is a particularly strong discrepancy for McMaster et al. (1969[McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.]) (reference c in Table 2[link]) between the adjacent elements which is not replicated in other theory or experiment, which could be due to a typographic or convergence error; or could be due to the issue of scattering. Data from McMaster et al. (1969[McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.]) are taken from a variety of sources, and therefore inherently inconsistent. Given the definitions and equations above, we recompute these ratios in Table 3[link] after subtracting off elastic and inelastic scattering (Rayleigh and Compton scattering contributions to attenuation) so that we can report ratios of photoabsorption as defined. In this paper, we have used Chantler (2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]) for scattering components, ideal for amorphous systems or metals. In oriented crystals the coherent cross-section follows Bragg–Laue scattering or thermal diffuse scattering; these coefficients are very different from Chantler (2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]) or Scofield, 1973[Scofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.]) and Berger & Hubbell (1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]). In any case, the edge jump (`A' − `B') is independent of this scattering contribution as seen in the tables. Correction for scattering increases the values of JK from the previous table by about 0.006; and similarly increases the values of rK by about 0.39, much larger than claimed respective uncertainties. With the exception of the anomalous value for McMaster et al. (1969[McMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.]), none of the experimental or other theoretical values appear to follow the formulae stated using [[{{\mu}/{\rho}}]_{\rm{pe}}]. Instead all of them appear to take ratios of attenuation coefficients [[{{\mu}/{\rho}}]_{\rm{tot}}]. In light of this, we define (or redefine) the attenuation edge jump ratio and factor to be rK,tot and JK,tot.

Table 3
Experimental and theoretical values of the K-shell absorption jump factor and jump ratio for zinc metal

  [[{{\mu}/{\rho}}]_{\rm{pe}}] (E < EK) [[{{\mu}/{\rho}}]_{\rm{pe}}] (E > EK) Jump rK JK
Source `B' (cm2 g−1) `A' (cm2 g−1) `A' − `B' (cm2 g−1) `A'/`B' (`A' − 'B')/`A'
XCOMa 33.413 252.267 218.853 7.5450 0.8675
FFASTb 32.068 259.164 227.096 8.082 0.8763
This workc 31.627 268.250 236.622 8.482 0.8821
Estimated σ 0.3 3 3 0.124 0.0149
aXCOM (Berger & Hubbell, 1987[Berger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87-3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.]) using [μ/ρ]R+C (E = EK) ≃ 1.684 cm−2 g−1. bFFAST (Chantler, 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]) using [μ/ρ]R+C (E = EK) ≃ 1.742 cm−2 g−1. cUsing [μ/ρ]R+C (E = EK) ≃ 1.742 cm−2 g−1 (Chantler, 2000[Chantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597-1056.]).

With these difficulties, one might question the utility of the edge jump ratios and edge jump factors. However, they remain extremely important if non-trivial to extract experimentally. The widespread use of XRF and EDXRF argues strongly for increased understanding of these ratios and how to use and interpret them; the propagation errors and uncertainties of XANES and XAFS analysis, especially at low k, depend upon the stability and uniqueness of the definition of the edge and spline or other background subtraction; and more theoretical and experimental work is required in this area in attenuation, absorption and fluorescence measurements. Immediately, we can use this analysis to justify the data presented herein; and to emphasize or demonstrate the error of Singh Sidhu et al. (2011[Singh Sidhu, B., Dhaliwal, A. S., Mann, K. S. & Kahlon, K. S. (2011). Radiat. Phys. Chem. 80, 28-32.]) observed in Table 2[link] [references Singh Sidhu et al. (2011[Singh Sidhu, B., Dhaliwal, A. S., Mann, K. S. & Kahlon, K. S. (2011). Radiat. Phys. Chem. 80, 28-32.]) and Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.])].

Previous work to obtain the edge jumps of the zinc K-edge using this method (Sier et al., 2020[Sier, D., Cousland, G. P., Trevorah, R. M., Ekanayake, R. S. K., Tran, C. Q., Hester, J. R. & Chantler, C. T. (2020). J. Synchrotron Rad. 27, 1262-1277.]) on zinc selenide found the jump ratio to be rK,tot = 3.380. However, as this work was conducted on a compound, this value must be scaled by the relative atomic composition of zinc within zinc selenide (0.453) which converts the value to rK,tot = 7.461 as the attenuation ratio; and to rK = 3.406 as the photoabsorption ratio in the compound, and rK = 7.519 as the photoabsorption ratio for the Zn. Note that ZnSe does not have Rayleigh scattering but rather thermal diffuse scattering, so the elastic scattering coefficients are necessarily different from the tabulations. Similarly, in that work rK,tot = 2.153 was reported for the Se K-edge jump ratio with rK = 2.161 for the compound, and rK,tot = 3.936 and rK = 3.951 for the selenium edge.

Rae et al. (2010[Rae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.]) measured the attenuation coefficient of pure zinc metal. The data are far too sparse to enable effective spline fitting of the XAFS above the edge; or to extract a good below-edge background. Taking a ratio of the photoabsorption data points closest to the edge results in an rK of 7.426. The sparse nature of this data means that the post-edge point is significantly lower in attenuation than the true pre-edge, thus leading to a lower rK.

Overall for Zn metal we are able to report absorption ratios rK = 8.482 (12) (124) where the first uncertainty is statistical and the second is systematic; and JK = 0.882 (15) (149); and we are able to report rK,tot = 8.091 (12) (116); and JK,tot = 0.8764 (15) (148) following past evaluations using the attenuation coefficient ratios. These measurements are and must be higher than the theoretical predictions of either FFAST or XCOM, though FFAST is almost within uncertainty of the experiment, which is promising for future work. This confirmation allows us to proceed with some confidence to investigating the nanostructure from the XAFS signal.

5. X-ray absorption fine structure

The XAFS of zinc above the zinc K-edge (Fig. 1[link]) probes the oscillations of the photoabsorption coefficients above the edge due to the photoelectron interference. The oscillatory part of the spectrum was isolated using the fine-structure function

[\chi(E) = {{\mu(E)-\mu_{0}(E)} \over {\mu(E_{0})}}, \eqno(7)]

where μ0(E) is a smooth atom-like background, E0 is the edge energy – ideally the onset of the open channel to the continuum but often (as here) estimated by the maximum of the first derivative across the edge. χ is then expressed as a function of k, the effective photoelectron wavenumber,

[k = \left[{{{2m_{\rm{e}}} \over {\hbar^{2}}} \left(E-E_{0}\right)} \right]^{1/2}, \eqno(8)]

where me is the mass of an electron, E is the photon energy and E0 is the edge energy.

The background function μ0(E) is obtained from an approximated spline using the eFEFFIT package (Smale et al., 2006[Smale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559-1563.]; Schalken & Chantler, 2018[Schalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920-934.]). The spline approach has been widely used to isolate the oscillatory part in the XAFS community for several decades (Newville et al., 1993[Newville, M., Līviņ\us, P., Yacoby, Y., Rehr, J. J. & Stern, E. A. (1993). Jpn. J. Appl. Phys. 32, 125-127.]; Smale et al., 2006[Smale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559-1563.]; Islam et al., 2016[Islam, M. T., Best, S. P., Bourke, J. D., Tantau, L. J., Tran, C. Q., Wang, F. & Chantler, C. T. (2016). J. Phys. Chem. C, 120, 9399-9418.]). The uncertainty of the oscillatory measurements was quantified using

[\sigma_{\chi(k)} = {{\sigma_{\mu(E)}} \over {\mu(E_{0})}}, \eqno(9)]

where σχ(k) is the uncertainty of χ(k) and σμ(E) is the experimental uncertainties of the X-ray absorption measurements. Figure 4[link] illustrates the oscillatory part of the XAFS spectrum with fully propagated uncertainties.

[Figure 4]
Figure 4
The fine-structure function of the zinc K-edge produced by the Mu2chi non-interpolation background subtraction software in the eFEFFIT package.

The XAFS equation describes the oscillatory part χ(k) as a sum over the multiple scattering paths,

[\eqalignno{ \chi(k) = {}& \sum\limits_{j} N_{j}\,S^{\,2}_{0}\,F_{j}(k) \, {{\sin\left[2kr_{j}+\phi_{j}(k)\right]}\over{kr^{2}_{j}}} \cr& \times \exp\left(-2\sigma^{2}_{j}k^{2}\right) \exp\left[\,p-2r_{j}\,/\,\lambda(k)\right], &(10)}]

where j denotes the jth path, Nj is the degeneracy, S 20 is the many-body reduction factor, Fj(k) is the backscattering amplitude function, rj is half the path length, ϕj(k) is the total phase shift function, σj is the Debye–Waller factor and λ(k) is the photoelectron mean free path function. A dominant standard approach for XAFS analysis uses equation (10)[link], perhaps particularly via the IFEFFIT computational package (Newville, 2001[Newville, M. (2001). J. Synchrotron Rad. 8, 96-100.]) together with FEFF (Rehr & Albers, 2000[Rehr, J. J. & Albers, R. C. (2000). Rev. Mod. Phys. 72, 621-654.]; Rehr et al., 1991[Rehr, J. J., Mustre de Leon, J., Zabinsky, S. I. & Albers, R. C. (1991). J. Am. Chem. Soc. 113, 5135-5140.]), to investigate the local nanostructure of materials. The lattice constants of zinc have been determined by single-crystal X-ray diffraction, to be a = 266.36 (1) pm and c = 494.57 (2) pm with c/a = 1.857 at a temperature of 300 K (Hull & Davey, 1921[Hull, A. W. & Davey, W. P. (1921). Phys. Rev. 17, 549-570.]; Nuss et al., 2010[Nuss, J., Wedig, U., Kirfel, A. & Jansen, M. (2010). Z. Anorg. Allg. Chem. 636, 309-313.]). Zinc is located at the special Wyckoff position 2c ( 1/3,2/3,1/4) in space group P63/mmc(194). The theoretical prediction from FEFF was used in eFEFFIT (Smale et al., 2006[Smale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559-1563.]; Schalken & Chantler, 2018[Schalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920-934.]), modified from the IFEFFIT package to propagate uncertainties. Data were fitted across several k-ranges to investigate robustness. Most effective was the range of k from 4.5 Å −1 to 17 Å −1 using absolute uncertainties of input data. Below this range, the theory underpinning the fitting function begins to break down. Above this range the uncertainties prove that the data are now dominated by noise.

After extensive analysis, the two nearest-neighbour bond lengths, also with the highest ranks from FEFF, were given two separate fitting parameters, whilst all other bond lengths were scaled by the usual α (Ravel, 2005[Ravel, B. (2005). J. Alloys Compd. 401, 118-126.]). Independent [\sigma_{i}^{2}] were used for the two nearest-neighbour bond lengths (Ravel & Kelly, 2007[Ravel, B. & Kelly, S. D. (2007). AIP Conf. Proc. 882, 150-152.]; Kelly et al., 2002[Kelly, S. D., Kemner, K. M., Fein, J. B., Fowle, D. A., Boyanov, M. I., Bunker, B. A. & Yee, N. (2002). Geochim. Cosmochim. Acta, 66, 3855-3871.]; Ravel, 2009[Ravel, B. (2009). J. Phys. Conf. Ser. 190, 012026.]). Linear paths were quantified in terms of the [\sigma_{i}^{2}] values of the single-scattering paths that contain the end atoms (Hudson et al., 1995[Hudson, E. A., Rehr, J. J. & Bucher, J. J. (1995). Phys. Rev. B, 52, 13815-13826.]). Particular paths for double forward-scattering followed recommendations from Ravel (2016[Ravel, B. (2016). Artemis manual, https://bruceravel.github.io/demeter/documents/Artemis/forward.html.]) about increased thermal parameters expected ([2\sigma_{i}^{2}] and [1.5\sigma_{i}^{2}] depending upon path combination). FEFF6 and FEFF8 gave similar results, parameters, uncertainties and [\chi^{2}_{r}]. Path information and ordering does change between the versions, so parameter files need to be checked carefully in going from one to the other. Results presented here are for FEFF6.

This analysis expanded the scattering to 45 paths, i.e. out to a maximum R ≃ 6.7 Å. A model with a single scale α and a single thermal parameter σ2 fails with [\chi^{2}_{r}] = 290, and a model with an α and σ2 assigned to the nearest-neighbour bond length and another α and σ2 to all other paths also fails with [\chi^{2}_{r}] = 16.8. Increasing the number of σ2 parameters to three is necessary and significant and a model with a single scale α yielded [\chi^{2}_{r}] = 18.1, and a model with an α assigned to the nearest-neighbour bond length and another to all other paths yielded a relatively poor [\chi^{2}_{r}] = 14.4. A developed model with three α and three σ2 values for the relevant paths led to [\chi^{2}_{r}] = 9.5. Finally, incorporating a [\sigma^{2}_{4}] value for two single-scattering paths (`8' and `9') led to the final [\chi^{2}_{r}] = 6.66. Other investigations of further independent parameters including a separated α showed insignificant improvement of the fits. However, note that the space is highly correlated in this area.

The incorporation of the path lengths and σ2 parameters led to uncertainties in these parameters reducing from 0.5% to between 0.10% to 0.29% for the bond lengths or to ±0.003 Å to ±0.008 Å; and led to σ2 thermal parameters with uncertainties around 3–5%. Perhaps more significantly, S0 2 converged from the simpler fits with fewer parameters of 1.7 ± 0.3 to 0.904 ± 0.037 with four independent [\sigma_{i}^{2}] parameters.

This demonstrates the need for four independent σ2 thermal parameters, and the need for three independent path length scalings. Whilst the parameters are moderately robust in these simpler analyses, the fits are poor until the two shortest bond lengths are given independent path lengths and thermal broadening.

Zinc could have a single scaling (expansion) parameter relative to a crystallographic determination, if the metal is ideal at zero temperature; or of course if the crystallographic structure, the difference between mean lattice sites, were fully consistent with the dynamical bond lengths and path lengths. In this experiment, however, dynamical motion at room temperature might be expected to lead to significant variation of expansion of the path lengths relative to a corresponding crystallographic determination. Fitted parameters are summarized in Table 4[link] and results are illustrated in Fig. 5[link]. Fits give good physical agreement and parameters are consistent within 1σ.

Table 4
FEFF6 and eFEFFIT parameters

Parameter Value Error
S 20 0.904 0.037
[E^{\,0}_{\rm{corr}}] 4.70 eV 0.29 eV
Zn—Zna 2.660 Å 0.003 Å
Zn—Znb 2.832 Å 0.008 Å
[\alpha_{1}]a 0.9999 0.0010
[\alpha_{2}]b 0.9737 0.0027
[\alpha_{3}]c 1.0061 0.0013
[\sigma^{2}_{1}]a 0.0101 Å2 0.0003 Å2
[\sigma^{2}_{2}]b 0.0198 Å2 0.0008 Å2
[\sigma^{2}_{3}]c 0.0221 Å2 0.0007 Å2
[\sigma^{2}_{4}]d 0.0169 Å2 0.0008 Å2
[\chi^{2}_{r}] 6.66  
aFor nearest-neighbour (shortest) scattering path. bFor the second nearest atom. cMost other scattering paths (see supporting information for a full listing). dFor single-scattering paths `8' and `9'.
[Figure 5]
Figure 5
Experimental data (black) with absolute uncertainties and the fitted model (red) across k = 4.5 Å−1 to 17 Å−1 for the zinc K-edge. Data were fitted over a Hanning window.

An interesting result of this analysis is the different expansions obtained for the nearest and second-nearest scattering paths. The reported nearest-neighbour and the second-nearest-neighbour Zn—Zn bond lengths are 2.6636 Å and 2.9120 Å (Nuss et al., 2010[Nuss, J., Wedig, U., Kirfel, A. & Jansen, M. (2010). Z. Anorg. Allg. Chem. 636, 309-313.]) (Fig. 6[link]). This work determines bond lengths 0.11% (±0.10%) and 2.7% (±0.3%) smaller, respectively. The scaling parameter obtained for the other bond lengths are 0.62% ± 0.13% larger than the scaling of the nearest neighbour bond length.

[Figure 6]
Figure 6
Crystal structure of zinc at 300 K with a form of hexagonal close packing, in which each atom has six nearest neighbours at 2.66 Å in its own plane and others at a greater distance of 4.95 Å.

These 5–7 standard error variations can be due to dynamic motion, or to the path-length including motion perpendicular to the bond length. XAFS is sensitive to atomic vibrations perpendicular to the bond direction, and to (positively or negatively) correlated motion of the atoms of the corresponding bond to increase or decrease the bond length compared with the crystallographic lattice site separation. Hence a bond length 0.004 Å to 0.08 Å smaller than X-ray diffraction (XRD) results at room temperature would be no surprise especially if the motion of both atoms was positively correlated.

The shift of next-neighbour bond lengths compared with XRD may alternatively be due to the sensitivity with temperature to the c/a axis. According to crystallographic determinations, a 5 K error in temperature would have only a shift of c/a of order 0.0003 Å so is inadequate to explain the lattice variation observations. Similarly, rolled foils have some static disorder from twinning or stacking faults, compared for example with an ideal crystal structure. Changing ABABAB to ABABCAB compromises the assigned Wykoff positions in the earlier calculations. This static disorder would then confuse the separation of static disorder from thermal disorder. This question can explicitly be clarified with a temperature series. To first order this will increase the disorder, i.e. the σ on the bond lengths, rather than the bond lengths themselves. In particular, this will increase the σ or mean isotropic thermal parameters for farther atomic pairs. To second order this would increase the bond lengths or lattice separations especially with a high density of phase boundaries, stacking faults or microstructure such as twinning. However, the largest observed effect with respect to the nearest-neighbour bond length is a decrease rather than an increase, so is the incorrect sign. The bond lengths fitted for the farther Zn moieties is an increase so may be augmented by such details. In fact, we know that thermal parameters increase with path length and distance, and this is naturally due to the increasing disorder, static and dynamic, as seen in most XAS studies. The uncertainty of the larger bond lengths is about 0.13% suggesting significant variation of bond lengths, and supported by the increased thermal parameters.

In this case the evidence is inconclusive, but quite suggestive of local dynamical motion and local order combined with distant increasing static disorder. Hence this suggests dynamic motion in the crystal lattice inaccessible to other techniques.

6. Conclusion

X-ray mass attenuation coefficients and the imaginary component of the atomic form factor of zinc were measured over the energy range from 8.51 keV to 11.5914 keV including the zinc K-edge and XAFS with 496 data points at different energies. The accuracy of measured mass attenuation coefficients is about 0.023% to 0.036% and some 50% better than previous claims. These measurements are the most accurate zinc K-edge data set.

Comparison of the experimental data with the tabulated FFAST and XCOM theory shows very good agreement. The edge jump and jump ratio analysis shows some discrepancies due to the challenges of defining both the below-edge and above-edge absorption coefficient, so that most of the preceding literature uses ratios and jump factors for the attenuation coefficient thereby including scattering – we separate these ideas by defining (or redefining) the attenuation ratio and factor to be rK,tot and JK,tot. These measures are also challenging because of the experimental pre-edge and edge-width below the edge; and of the typical use of splines to attempt to deal with the above-edge XAFS. Nonetheless, this is a valuable confirmation of the scaling and absolute value of the coefficients and can be used for any experiment as a test of the quality of the data and data processing.

The high-accuracy measurements were improved for structural investigations. Robust investigation of zinc nano­structure was carried out using eFEFFIT with propagated uncertainties from XERT. X-ray absorption fine-structure analysis shows an excellent agreement between the measured and tabulated values and yields bond lengths and nano­structure of zinc with uncertainties of 0.11% to 0.27% or 0.003 Å to 0.009 Å, respectively. Thermal σ2 parameters were fitted with an accuracy of from approximately 5%, and S0 2 is determined to be 0.904 ± 0.037. Observed significant variation from reported crystal structures suggests local dynamic motion compared with a standard crystal lattice and increasing static disorder as expected in a non-ideally crystalline environment such as a metal foil. XAFS is sensitive to dynamic correlated motion so is in principle able to observe local dynamic motion beyond the reach of conventional crystallography.

These high-accuracy studies also allow one to investigate solid state effects including inelastic mean free paths, and inelastic and elastic scattering cross-sections.

Supporting information


Acknowledgements

We sincerely acknowledge and are very grateful for the collaboration with Zwi Barnea who has driven the development of XERT with us and also has driven the investigation of Zn. Part of this research was undertaken on the XAS beamline at the Australian Synchrotron, part of ANSTO. We thank beamline scientists in the Australian Synchrotron for their efforts and dedication to build up part of this methodology. We thank all of the synchrotron team including Jeremy Wykes, Chris Glover and Susan Cumberland. We thank our attending team members Stephen P. Best and Victor A. Streltsov for their contributions here and in earlier experiments. We thank Hamish A. Melia for his support for this work and we thank Ryan J. Trevorah and Geoffrey P. Cousland for helpful reading of the manuscript.

References

First citationAyala, A. P. & Mainardi, R. T. (1996). Radiat. Phys. Chem. 47, 177–181.  CrossRef CAS Web of Science Google Scholar
First citationBambynek, W., Crasemann, B., Fink, R. W., Freund, H. U., Mark, H., Swift, C. D., Price, R. E. & Rao, P. V. (1972). Rev. Mod. Phys. 44, 716–813.  CrossRef CAS Web of Science Google Scholar
First citationBennal, A. S. & Badiger, N. M. (2007). J. Phys. B At. Mol. Opt. Phys. 40, 2189–2199.  Web of Science CrossRef CAS Google Scholar
First citationBerger, M. J. & Hubbell, J. H. (1987). XCOM: Photon Cross Sections on A National Bureau of Standards. Technical Report NBSIR-87–3597. Center for Radiation, National Bureau of Standards, Washington, DC, USA.  Google Scholar
First citationBohic, S., Simionovici, A., Snigirev, A., Ortega, R. G., Devès, D., Heymann, D. & Schroer, C. G. (2001). Appl. Phys. Lett. 78, 3544–3546.  Web of Science CrossRef CAS Google Scholar
First citationBroll, N. (1986). X-ray Spectrom. 15, 271–285.  CrossRef CAS Web of Science Google Scholar
First citationBudak, G., Karabulut, A. & Ertuǧrul, M. (2003). Radiat. Meas. 37, 103–107.  Web of Science CrossRef CAS Google Scholar
First citationBudak, G. & Polat, R. (2004). J. Quant. Spectrosc. Radiat. Transfer, 88, 525–532.  CrossRef CAS Google Scholar
First citationChantler, C. T. (1995). J. Phys. Chem. Ref. Data, 24, 71–643.  CrossRef CAS Web of Science Google Scholar
First citationChantler, C. T. (2000). J. Phys. Chem. Ref. Data, 29, 597–1056.  Web of Science CrossRef CAS Google Scholar
First citationChantler, C. T. (2009). Eur. Phys. J. Spec. Top. 169, 147–153.  Web of Science CrossRef Google Scholar
First citationChantler, C. T. & Bourke, J. D. (2010). J. Phys. Chem. Lett. 1, 2422–2427.  Web of Science CrossRef CAS Google Scholar
First citationChantler, C. T., Tran, C. Q., Barnea, Z., Paterson, D., Cookson, D. J. & Balaic, D. X. (2001). Phys. Rev. A, 64, 062506.  Web of Science CrossRef Google Scholar
First citationCoulthard, I. & Sham, T. K. (1996). Phys. Rev. Lett. 77, 4824–4827.  CrossRef PubMed CAS Web of Science Google Scholar
First citationEkanayake, R. S. K., Chantler, C. T., Sier, D., Schalken, M. J., Illig, A. J., de Jonge, M. D., Johannesen, B., Kappen, P. & Tran, C. Q. (2021). J. Synchrotron Rad. 28, 1476–1491.  CrossRef CAS IUCr Journals Google Scholar
First citationErtugrul, M., Karabulut, A. & Budak, G. (2002). Radiat. Phys. Chem. 64, 1–3.  Web of Science CrossRef CAS Google Scholar
First citationGerward, L. (1992). Nucl. Instrum. Methods Phys. Res. B, 69, 407–412.  CrossRef Web of Science Google Scholar
First citationGerward, L., Guilbert, N., Bjørn Jensen, K. & Levring, H. (2001). Radiat. Phys. Chem. 60, 23–24.  Web of Science CrossRef CAS Google Scholar
First citationGlover, J. L., Chantler, C. T., Barnea, Z., Rae, N. A. & Tran, C. Q. (2010). J. Phys. B At. Mol. Opt. Phys. 43, 085001.  Web of Science CrossRef Google Scholar
First citationGlover, J. L., Chantler, C. T., Barnea, Z., Rae, N. A., Tran, C. Q., Creagh, D. C., Paterson, D. & Dhal, B. B. (2008). Phys. Rev. A, 78, 052902.  Web of Science CrossRef Google Scholar
First citationGlover, J. L., Chantler, C. T. & de Jonge, M. D. (2009). Phys. Lett. A, 373, 1177–1180.  Web of Science CrossRef CAS Google Scholar
First citationHan, S., Stern, E., Haskel, D. & Moodenbaugh, A. R. (2002). Phys. Rev. B, 66, 094101.  Web of Science CrossRef Google Scholar
First citationHopersky, A., Petrov, I., Nadolinsky, A., Yavna, V. & Koneev, R. (2004). J. Phys. B At. Mol. Opt. Phys. 37, 3313–3319.  Web of Science CrossRef CAS Google Scholar
First citationHopkins, J. I. (1959). J. Appl. Phys. 30, 185–187.  CrossRef Web of Science Google Scholar
First citationHossain, F. M., Riley, D. P. & Murch, G. E. (2005). Phys. Rev. B, 72, 235101.  Web of Science CrossRef Google Scholar
First citationHudson, E. A., Rehr, J. J. & Bucher, J. J. (1995). Phys. Rev. B, 52, 13815–13826.  CrossRef CAS Web of Science Google Scholar
First citationHull, A. W. & Davey, W. P. (1921). Phys. Rev. 17, 549–570.  CrossRef CAS Google Scholar
First citationIgnatov, A. Y., Ali, N. & Khalid, S. (2001). Phys. Rev. B, 64, 014413.  Web of Science CrossRef Google Scholar
First citationIslam, M. T., Best, S. P., Bourke, J. D., Tantau, L. J., Tran, C. Q., Wang, F. & Chantler, C. T. (2016). J. Phys. Chem. C, 120, 9399–9418.  Web of Science CrossRef CAS Google Scholar
First citationIslam, M. T., Chantler, C. T., Cheah, M. H., Tantau, L. J., Tran, C. Q. & Best, S. P. (2015). J. Synchrotron Rad. 22, 1475–1491.  Web of Science CrossRef IUCr Journals Google Scholar
First citationIslam, M. T., Rae, N. A., Glover, J. L., Barnea, Z., de Jonge, M. D., Tran, C. Q., Wang, J. & Chantler, C. T. (2010). Phys. Rev. A, 81, 022903.  Web of Science CrossRef Google Scholar
First citationIslam, M. T., Tantau, L. J., Rae, N. A., Barnea, Z., Tran, C. Q. & Chantler, C. T. (2014). J. Synchrotron Rad. 21, 413–423.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationJoly, Y. (2001). Phys. Rev. B, 63, 125120–125130.  Web of Science CrossRef Google Scholar
First citationJoly, Y., Cabaret, D., Renevier, H. & Natoli, C. R. (1999). Phys. Rev. Lett. 82, 2398–2401.  Web of Science CrossRef CAS Google Scholar
First citationJonge, M. D. de, Tran, C. Q., Chantler, C. T., Barnea, Z., Dhal, B. B., Cookson, D. J., Lee, W. & Mashayekhi, A. (2005). Phys. Rev. A, 71, 032702.  Google Scholar
First citationJonge, M. D. de, Tran, C. Q., Chantler, C. T., Barnea, Z., Dhal, B. B., Paterson, D., Kanter, E. P., Southworth, S. H., Young, L., Beno, M. A., Linton, J. A. & Jennings, G. (2007). Phys. Rev. A, 75, 032702.  Google Scholar
First citationKaya, N., Apaydin, G. & Tirasoglu, E. (2011). Radiat. Phys. Chem. 80, 677–681.  Web of Science CrossRef CAS Google Scholar
First citationKaya, N., Tıraşoğlu, E., Apaydın, G., Aylıkcı, V. & Cengiz, E. (2007). Nucl. Instrum. Methods Phys. Res. B, 262, 16–23.  Web of Science CrossRef CAS Google Scholar
First citationKelly, S. D., Kemner, K. M., Fein, J. B., Fowle, D. A., Boyanov, M. I., Bunker, B. A. & Yee, N. (2002). Geochim. Cosmochim. Acta, 66, 3855–3871.  Web of Science CrossRef CAS Google Scholar
First citationLatyshev, G. D. (1947). Rev. Mod. Phys. 19, 132–145.  CrossRef CAS Web of Science Google Scholar
First citationMallikarjuna, M. L., Appaji Gowda, S. B., Gowda, R. & Umesh, T. K. (2002). Radiat. Phys. Chem. 65, 217–223.  Web of Science CrossRef CAS Google Scholar
First citationMcMaster, W. H., Del Grande, N. K., Mallett, J. H. & Hubbell, J. H. (1969). Compilation of X-ray Cross Sections. Technical Report UCRL-50174. Lawrence Livermore Laboratory, USA.  Google Scholar
First citationNayak, S. V. & Badiger, N. M. (2006). J. Phys. B At. Mol. Opt. Phys. 39, 2893–2900.  Web of Science CrossRef CAS Google Scholar
First citationNewville, M. (2001). J. Synchrotron Rad. 8, 96–100.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNewville, M., Līviņ\us, P., Yacoby, Y., Rehr, J. J. & Stern, E. A. (1993). Jpn. J. Appl. Phys. 32, 125–127.  Google Scholar
First citationNuss, J., Wedig, U., Kirfel, A. & Jansen, M. (2010). Z. Anorg. Allg. Chem. 636, 309–313.  Web of Science CrossRef ICSD CAS Google Scholar
First citationOuseph, P. J., Hoskins, K. D., Berman, J. I. & Bolander, A. J. (1982). Am. J. Phys. 50, 275–276.  CrossRef Web of Science Google Scholar
First citationPalmberg, P. W. & Rhodin, T. N. (1968). J. Appl. Phys. 39, 2425–2432.  CrossRef CAS Web of Science Google Scholar
First citationPolat, R., Budak, G., Gürol, A., Karabulut, A. & Ertugřul, M. (2005). Radiat. Meas. 39, 409–415.  Web of Science CrossRef CAS Google Scholar
First citationPolat, R., İçelli, O. & Budak, G. (2004). Anal. Chim. Acta, 505, 307–314.  Web of Science CrossRef CAS Google Scholar
First citationPuttaswamy, K. S., Gowda, R. & Sanjeevaiah, B. (1981). Can. J. Phys. 59, 853–858.  CrossRef CAS Web of Science Google Scholar
First citationRae, N. A., Chantler, C. T., Barnea, Z., de Jonge, M. D., Tran, C. Q. & Hester, J. R. (2010). Phys. Rev. A, 81, 022904.  Web of Science CrossRef Google Scholar
First citationRavel, B. (2005). J. Alloys Compd. 401, 118–126.  Web of Science CrossRef CAS Google Scholar
First citationRavel, B. (2009). J. Phys. Conf. Ser. 190, 012026.  CrossRef Google Scholar
First citationRavel, B. (2016). Artemis manual, https://bruceravel.github.io/demeter/documents/Artemis/forward.htmlGoogle Scholar
First citationRavel, B. & Kelly, S. D. (2007). AIP Conf. Proc. 882, 150–152.  CrossRef CAS Google Scholar
First citationRavel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537–541.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationRehr, J. J. & Albers, R. C. (2000). Rev. Mod. Phys. 72, 621–654.  Web of Science CrossRef CAS Google Scholar
First citationRehr, J. J., Mustre de Leon, J., Zabinsky, S. I. & Albers, R. C. (1991). J. Am. Chem. Soc. 113, 5135–5140.  CrossRef CAS Web of Science Google Scholar
First citationRontgen, W. (1896). Nature, 53, 274–276.  Google Scholar
First citationSaloman, E. & Hubbell, J. (1987). Nucl. Instrum. Methods Phys. Res. A, 255, 38–42.  CrossRef Web of Science Google Scholar
First citationSayers, D. E., Lytle, F. W. & Stern, E. A. (1972). J. Non-Cryst. Solids, 8–10, 401–407.  CrossRef CAS Google Scholar
First citationSchalken, M. J. & Chantler, C. T. (2018). J. Synchrotron Rad. 25, 920–934.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSchmickley, R. D. & Pratt, R. H. (1967). Phys. Rev. 164, 104–116.  CrossRef CAS Web of Science Google Scholar
First citationScofield, J. H. (1973). Technical Report UCRL-51326. Lawrence Livermore Laboratory, USA.  Google Scholar
First citationSier, D., Cousland, G. P., Trevorah, R. M., Ekanayake, R. S. K., Tran, C. Q., Hester, J. R. & Chantler, C. T. (2020). J. Synchrotron Rad. 27, 1262–1277.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSingh Sidhu, B., Dhaliwal, A. S., Mann, K. S. & Kahlon, K. S. (2011). Radiat. Phys. Chem. 80, 28–32.  Web of Science CrossRef CAS Google Scholar
First citationSmale, L. F., Chantler, C. T., de Jonge, M. D., Barnea, Z. & Tran, C. Q. (2006). Radiat. Phys. Chem. 75, 1559–1563.  Web of Science CrossRef CAS Google Scholar
First citationStorm, E. & Israel, I. H. (1970). At. Data Nucl. Data Tables, 7, 565–681.  CrossRef CAS Google Scholar
First citationStreltsov, V. A., Ekanayake, R. S., Drew, S. C., Chantler, C. T. & Best, S. P. (2018). Inorg. Chem. 57, 11422–11435.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTantau, L. J., Chantler, C. T., Bourke, J. D., Islam, M. T., Payne, A. T., Rae, N. A. & Tran, C. Q. (2015). J. Phys. Condens. Matter, 27, 266301.  Web of Science CrossRef PubMed Google Scholar
First citationTran, C. Q., Chantler, C. T., Barnea, Z., Jonge, M. D., Dhal, B. B., Chung, C. T., Paterson, D. & Wang, J. (2005). J. Phys. B At. Mol. Opt. Phys. 38, 89–107.  Web of Science CrossRef CAS Google Scholar
First citationTrevorah, R. M., Chantler, C. T. & Schalken, M. J. (2019). IUCrJ, 6, 586–602.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationUnonius, L. & Suortti, P. (1989). J. Appl. Cryst. 22, 46–52.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationVeigele, W. J. (1973). At. Data Nucl. Data Tables, 5, 51–111.  CrossRef Google Scholar

This article is published by the International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds