Next Article in Journal
Maximum Ionization in Restricted and Unrestricted Hartree-Fock Theory
Next Article in Special Issue
Data Needs for Modeling Low-Temperature Non-Equilibrium Plasmas: The LXCat Project, History, Perspectives and a Tutorial
Previous Article in Journal
Loading a Paul Trap: Densities, Capacities, and Scaling in the Saturation Regime
Previous Article in Special Issue
A Decade with VAMDC: Results and Ambitions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The XSTAR Atomic Database

1
Department of Physics, Western Michigan University, Kalamazoo, MI 49008, USA
2
Physics Center, Venezuelan Institute for Scientific Research (IVIC), Caracas 1020, Venezuela
3
Physique Atomique et Astrophysique, Université de Mons—UMONS, B-7000 Mons, Belgium
4
Cahill Center for Astronomy and Astrophysics, California Institute of Technology, Pasadena, CA 91125, USA
5
Dr. Karl Remeis-Observatory and Erlangen Centre for Astroparticle Physics, Sternwartstr. 7, 96049 Bamberg, Germany
6
Max Planck Institute for Extraterrestrial Physics, Gießenbachstraße 1, 85748 Garching bei München, Germany
7
NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
8
IPNAS, Université de Liège, Sart Tilman, B-4000 Liège, Belgium
9
ADNET Systems Inc., Bethesda, MD 20817, USA
*
Author to whom correspondence should be addressed.
Submission received: 3 December 2020 / Revised: 21 January 2021 / Accepted: 27 January 2021 / Published: 5 February 2021
(This article belongs to the Special Issue Development and Perspectives of Atomic and Molecular Databases)

Abstract

:
We describe the atomic database of the xstar spectral modeling code, summarizing the systematic upgrades carried out in the past twenty years to enable the modeling of K-lines from chemical elements with atomic number Z 30 and recent extensions to handle high-density plasmas. Such plasma environments are found, for instance, in the inner region of accretion disks round compact objects (neutron stars and black holes), which emit rich information about the system’s physical properties. Our intention is to offer a reliable modeling tool to take advantage of the outstanding spectral capabilities of the new generation of X-ray space telescopes (e.g., xrism and athena) to be launched in the coming years. Data curatorial aspects are discussed and an updated list of reference sources is compiled to improve the database provenance metadata. Two xstar spin-offs—the ISMabs absorption model and the uaDB database—are also described.

1. Introduction

The xstar (https://heasarc.gsfc.nasa.gov/docs/software/xstar/xstar.html) code computes the physical conditions and emission spectra of a photoionized gas and has been widely used in astrophysics, most notably in X-ray astronomy, for the past 20 years [1,2]. It assumes an ionizing radiation source surrounded by a spherical gas shell that absorbs and transfers parts of this radiation to finally emit a spectrum. From an input comprising the incident continuum, shell thickness, elemental abundances, and gas density, the code computes the ionization balance, level populations, and temperature to generate the gas opacity and the emitted line and continuum fluxes. The computational model determines simultaneously the gas state and radiation field at every user-defined spatial zone of the cloud. To ensure convergence in local thermodynamic equilibrium (LTE) conditions, it implements a detailed and consistent treatment of the radiative (bound–bound, bound–free, and free–free) and collisional (electron-impact excitation and ionization, electron–ion recombination, charge transfer and three-body recombination) processes, a procedure that in turn depends on the accuracy of the underlying database of atomic rate coefficients. Therefore, xstar is an ideal platform to study data curation schemes for implementing and maintaining application-based atomic databases.
Since 2001, we have been systematically computing for the xstar database the atomic data required to model the K-lines and edges of both cosmically abundant and trace elements [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46]. These X-ray spectral features are prime candidates to devise diagnostics to determine the plasma conditions and chemical abundances in a wide variety of astronomical entities: interstellar, intergalactic, and intra-cluster media; active galactic nuclei (AGN); X-ray binaries; supernova remnants; stars; the Sun, and some solar planets and satellites. This long-term collaboration has reached an important milestone, and the corresponding upgrading of the xstar database with the new datasets has more than doubled its volume, bringing to the fore a series of data curational problems discussed in this report. Our final goal is to upgrade xstar to take advantage of the relatively high spectral resolution (5 eV in the soft X-ray band ∼0.3–12 keV) and sensitivity of the microcalorimeter-based spectrometers in the joint NASA/JAXA X-ray Imaging and Spectroscopy Mission (xrism, https://heasarc.gsfc.nasa.gov/docs/xrism) to be launched in 2022 and in the European Space Agency flagship Advanced Telescope for High-ENergy Astrophysics (athena, https://sci.esa.int/web/athena). As an example of this accuracy level, Figure 1 shows the xstar fit in the 6.50–6.60 keV region of the spectrum of the Perseus Cluster taken by the hitomi (https://www.nasa.gov/hitomi) space telescope before its demise in 2016 [47].
We have recently introduced high-density ( n e > 10 18  cm 3 ) effects in the xstar database, which come into play in the reflection spectra of the inner region of the accretion disks around compact objects (e.g., black-hole candidates and neutron stars): continuum lowering [40,41,42,43,44,45]; dielectronic recombination suppression [48,49]; collisional ionization; three-body recombination; and stimulated emission. Since the Fe abundance is a key measure of line reprocessing in reflection models [50], we are now in a good position to test if the neglect of such high-density effects is responsible for the anomalously high abundances recurrently derived  [51]. Regarding xstar database curation, the inclusion of high-density effects presents serious execution difficulties that are now managed by pre-processing the database at each density point. Since this pre-processing time is generally long, a grid of pre-computed databases has been deployed that, considering the data volumes involved, favors a cloud computing environment for the xstar modeling of high-density plasmas.
Finally, the development of the xstar database has given rise to two spin-offs: the ISMabs (https://heasarc.gsfc.nasa.gov/xanadu/xspec/models/ismabs.html) absorption model [35] and the uaDB (https://heasarc.gsfc.nasa.gov/uadb/index.php) online database. The former is the result of benchmarks of the theoretical photoionization curves with observed K absorption in the interstellar medium (ISM), which has led to the implementation of a reliable model of ISM K absorption and to the matching of the theoretical K-line wavelengths with the astronomically observed and laboratory measured values, the latter two not always agreeing (e.g., O i [23,52]). The uaDB database allows the downloading of the xstar and other atomic rates without having to interfere with the code.

2. XSTAR Atomic Database

The xstar database comprises records of multilevel model ions (usually referred to as ion targets) and data structures to derive the rates for the microprocesses that determine the ionization and excitation of the plasma [2]. The rate types ( r t ) are the following:
01 .
Ground state ionization
03 .
Bound–bound collision
04 .
Bound–bound radiative
05 .
Bound–free collision (level)
06 .
Total recombination
07 .
Bound–free radiative (level)
08 .
Total recombination forcing normalization
09 .
Two-photon decay
11 .
Element data
12 .
Ion data
13 .
Level data
14 .
Bound–bound radiative, superlevel to spectroscopic level
15 .
Collisional ionization total rate
40 .
Bound–bound collisional, superlevel to spectroscopic level
41 .
Non-radiative Auger transition
42 .
Inner-shell photoabsorption followed by autoionization.
For each ionic species ( Z ; N ) in the ranges 1 Z 30 and 1 N Z , where Z and N are, respectively, the atomic and electron numbers, the data for the target models and processes are contained in ASCII files labeled d i o n t 0 d t . The ion index i o n is defined as
i o n = Z ( Z 1 ) / 2 + ( Z N ) + 1 ,
and d t is the data type (see Appendix A for a complete list). Each d i o n t 0 d t file contains integer, floating-point, and character variables. For each release, the database is transcribed to a binary FITS-format-derived [53,54] structure (atdb.fits) consisting of three one-dimensional arrays (floating point, integer, and character) and a pointer array, which are uploaded in main memory at runtime. This scheme ensures portability, fast data uploading when the code is invoked, and single-data accesses in main memory during processing.
The xstar data curation strategy essentially follows two principles:
  • Since the datasets are compiled from diverse sources, the original data structures and units are explicitly kept in the database records to be transcribed and unified at runtime. For instance, level energies are specified in eV but the photoionization cross-sections are usually tabulated in Rydbergs. Levels in some atomic models are listed in intermediate coupling (fine structure) while in others L S terms are adopted. These format differences are accounted for by defining data types, each of which adheres to its own set of rules for data tabulation.
  • As new data and processes are included in the database, new data types may be defined. Thus, photoionization cross-sections can be specified in more than one data type—valence shell and K shell, say—if the tabulation for each energy region comes from different sources. They are then concatenated and adjusted at runtime.
We show in Figure 2 the target model for Mg-like Zn xix complying with data type d t = 06 and contained in file d 454 t 006 . Each row is headed by the integer tuple ( d t , r t , 0 , N flt , N int , N chr ) , where d t is the data type, r t the rate type, N flt the number of floating-point parameters, N int number of integer parameters, and N chr the length of the character string. As shown in the rate-type list above, r t = 13 corresponds to level data. The first and second floating-point columns tabulate level energies (eV) and statistical weights g = 2 J + 1 (the other two floating-point columns have been edited out). The second integer tuple ( n , 2 S + 1 , L , Z , i , i o n ) lists: the n principal quantum number of the valence electron; the 2 S + 1 spin multiplicity; the total orbital angular momentum L; element atomic number Z; the i level index, and i o n index. The character string gives the level assignment (configuration and spectroscopic term), the end of the record being indicated with the % character. Records may span one arbitrary length line or multiple lines in the ASCII file.
This ion representation presents remarkable features. By looking at the level-i index and assignment, the Zn xix model contains data attributes for 31 levels from: the valence shell ( 1 i 10 ), open L shell ( 11 i 24 ), and open K shell ( 25 i 28 ); two superlevels (i = 29–30); and a continuum level ( i = 31 ). The latter is associated to the ionization potential of the ion. Therefore, the model comprises spectroscopic bound levels, quasi-bound L- and K-vacancy resonances, and artificial levels. The tabulation must identify the ground level with index 1 and the continuum level with the highest index. Apart from this constraint, the levels are not necessarily listed in energy order facilitating model upgrading, which simply involves adding new levels at the end of the existing tabulation. The level-assignment string has been carefully devised and used throughout to facilitate searches based on filled shells, shell holes, L S terms, and J total angular momenta.
Following [1], the superlevel accounts in H- and He-like ions for the recombination cascade from unaccounted high-lying levels to the spectroscopic levels of the model atom. Rates have been determined from population kinetic calculations. For other isoelectronic sequences, the superlevel is assumed to decay radiatively and collisionally to the ground state. Rates populating and depopulating the superlevel are then calculated by fitting to the total recombination rate, thus explicitly accounting for the contributions from excitation, ionization, and recombination to the level populations. As a result, every ionic process is accompanied by its detailed-balance inverse, ensuring the progressive convergence to the LTE level populations. The superlevel_[K] is defined in Section 3.1.
All other data types follow similar data structures (floating point, integer, and character) as those displayed in Figure 2. For instance, we show in Figure 3 an excerpt of file d 454 t 050 containing radiative data for the bound–bound transitions (upper level k 10 ) in Zn xix. As before, the header integer tuple prescribes the data and rate types d t = 50 and r t = 4 ; three floating-point items; four integer items; and no character strings. The floating-point tuple ( λ , g f ( i , k ) , A ( k , i ) ) lists the transition wavelength (Å), oscillator strength ( g f -value), and transition probability (A-value, s 1 ). The second integer tuple ( i , k , Z , i o n ) lists the lower and upper levels of the transition, atomic number, and ion index. By comparing the i and k transition indices with the i level indices in Figure 2, it may be seen that the tabulation includes both allowed and forbidden transitions, e.g., the resonance transition ( i , k ) = ( 1 , 5 ) and the forbidden transition ( i , k ) = ( 1 , 4 ) . Importantly, every excited level in the database must include a decay mechanism, radiative and/or collisional; otherwise, its population may grow out of bounds when modeling. For instance, the 3 s 1.3 p 1.3 P _ 0 metastable level ( i = 2 ) does not display a radiative transition ( j = 0 0 radiative transitions are strictly forbidden), and consequently, an alternative decay process must be provided in a separate data type; e.g., collisional de-excitation.
Data type d t = 86 , listing the radiative and Auger widths of K-vacancy levels, is also an interesting case as it involves transitions between different ion-charge states (Figure 4). The floating-point tuple ( E ( k N ) , A a ( k N ) , A a ( k N , i N 1 ) , A r ( k N ) ) lists: the energy E ( k N ) (in eV relative to the ionization threshold) and total Auger width A a ( k N ) (s 1 ) of the k N level of the N-electron ion ( i o n N = 454 ); the partial Auger width A a ( k N , i N 1 ) (s 1 ) leaving the ( N 1 )-electron ion ( i o n N 1 = 455 ) in the i N 1 level; and the total radiative width A r ( k N ) (s 1 ) of the k N level of the N-electron ion. The second integer tuple then specifies ( i N 1 , k N , Z , i o n N 1 , i o n N ) . In this example, the ( N 1 ) -electron ion is assumed to end up in its ground state ( i N 1 = 1 ) after the Auger transition; however, this is not always the case as the Auger transition may involve an excited end level of the ( N 1 ) -electron ion. Such partial Auger widths are not determined in some structure codes (e.g., hfr [55]) and in the R-matrix scattering method [56,57], the latter including radiative and Auger damping through an optical model potential [58].
Table A1 in Appendix B tabulates the number of levels for each ionic species ( Z ; N ) . It may be seen that most neutral and, in many cases, the singly ionized systems are not well represented, indicating that xstar is mainly destined to model photoionized plasmas for which such species have negligible fractions. The number of levels in the ion models is never above 1000 to enable practical runtimes when modeling (see Section 6.1), and the most relevant elements in the astronomical soft X-ray band—N, O, Ne, Fe, and Ni—are especially well treated.
As fully described in the original paper on the xstar database [2], atomic datasets have been collected from a myriad of individual sources and data compilations such as the NIST Atomic Spectra Database (https://www.nist.gov/pml/atomic-spectra-database) [59], TOPbase (http://cdsweb.u-strasbg.fr/topbase/topbase.html) [60], chianti (https://www.chiantidatabase.org/) [61], ADAS (https://www.adas.ac.uk/), Iron Project (http://cdsweb.u-strasbg.fr/topbase/testop/TheIP.html) [62], and PP95 (https://ned.ipac.caltech.edu/level5/Pradhan/Pradhan_contents.html) [63]. Major upgrades have been mainly concerned with the radiative recombination (RR) and dielectronic recombination (DR) data from the AMDPP tabulations (http://amdpp.phys.strath.ac.uk/tamoc/DATA/) [64,65] and from the systematic computations of atomic data for K-line diagnostics summarized in the next Section 3. A complete data provenance list is given in Appendix C.

3. Computation of Atomic Data for K-Line Diagnostics

Since xstar is widely used for spectral modeling in X-ray astrophysics, we have dedicated a systematic effort to compute the atomic data to characterize K-lines in ionic species with Z 30 —namely, valence and K-vacancy level energies, transition wavelengths, A-values, radiative and Auger widths, and photoionization cross-sections [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]. Due to the lack of spectroscopic data to validate most ionic models, a multi-code methodology was implemented right from the outset to compute the data and estimate their accuracy. Structure data were computed with the multiconfiguration codes hfr (Pauli Hamiltonian) [55], autostructure (Breit–Pauli Hamiltonian) [66,67], and grasp92 (Dirac–Coulomb–Breit Hamiltonian) [68], and the photoionization cross-sections with the Breit–Pauli R-matrix suite of codes [56,57] and autostructure. Due to the large target models for species in isoelectronic sequences with electron number 19 N 26 , the latter code was used to compute their photoionization cross-sections in the distorted-wave approximation. Initial priority was given to ions from the oxygen, nitrogen, and iron isonuclear sequences and then from even-Z elements, but those from odd-Z and trace elements have been recently completed.
In the calculations of accurate atomic data for K-line diagnostics, two key effects must be considered in detail: radiative and Auger damping and orbital relaxation. In Section 3.1 and Section 3.2, we briefly outline these two processes and the methods to include them in structure and scattering calculations.

3.1. Radiative and Auger Damping

Let us consider an ion, say, with electron number N 10 . When a photon promotes a K-shell electron to an excited Rydberg state,
h ν + 1 s 2 2 s λ 2 p μ 1 s 2 s λ 2 p μ n p ,
it subsequently decays via the radiative and Auger manifold
1 s 2 s λ 2 p μ n p K n 1 s 2 2 s λ 2 p μ + h ν n
K α 1 s 2 2 s λ 2 p μ 1 n p + h ν α
KL n 1 s 2 2 s λ 2 p μ 1 + e 1 s 2 2 s λ 1 2 p μ + e
KLL 1 s 2 2 s λ 2 p μ 2 n p + e 1 s 2 2 s λ 1 2 p μ 1 n p + e 1 s 2 2 p μ 2 n p + e
dominated by K α radiative decay (Equation (4)) and KLL Auger spectator-electron ionization (Equation (6)). Such channeling causes a damping effect (see Figure 5) since the 1 s 2 s λ 2 p μ n p resonances have broad and symmetric widths almost independent of n, which must be taken into account when devising plasma diagnostics. In the R-matrix method, it is computationally intractable to treat this channel array from first principles as it would involve target states with np orbitals in an ever increasing close-coupling expansion. Therefore, for the higher Rydberg resonances, damping is neatly accounted for via an additional imaginary potential V opt = i ( Γ r + Γ a ) / 2 involving the total radiative width Γ r —as developed in the Hickman–Robicheaux formalism [69,70] (which is essentially equivalent to the Davies–Seaton formalism [71,72])—and adapted to include the total Auger width Γ a  [58].
Due to their diagnostic importance, K α lines in the xstar database are treated as bound–bound transitions and assigned wavelengths and A-values, but the upper K-vacancy states also display Auger widths that lead to autoionization (see Figure 2, Figure 3 and Figure 4). However, the K resonances also appear in the photoabsorption cross-sections, as shown in Figure 5, smearing the edge due to their peculiarly symmetric and constant line profiles. As a result, the K α resonances must be trimmed out from the partial photoionization cross-sections that are used to derive the rates so as not to count them twice. However, it is more complicated to carry out this trimming procedure in the L-edge region due to the piling up of a large number of resonances converging to closely spaced thresholds; thus some degree of double counting is practically unavoidable, making the L-edge less useful to devise plasma diagnostics.
As shown in Figure 6, damping leads to a significant difference between photoabsorption and photoionization, and in the xstar code, both are needed: the former to determine opacities and the latter to derive the photoionization rates. However, the database essentially stores partial photoionization cross-sections, i.e., cross-sections leaving the photoionized ( N 1 ) -electron system in specific states, and consequently, the difference between the total photoabsorption and photoionization cross-sections is assigned to superlevel_[K] (see level i = 30 in Figure 2).
Due to the large number of target levels in ions with electron number N > 18 , their photoionization cross-sections were computed using the simpler distorted-wave method implemented in autostructure [29]. The level numbers for such ionic systems tabulated in Table A1 (Appendix B) underwent considerable trimming to comply with the modeling specifications of xstar and, hence, do not represent the extent of the actual collisional targets. In autostructure, photoabsorption is conveniently treated as the sum of two separate processes: photoionization and photoexcitation. However, the complicated decay routes of K-vacancy states delineated in Equations (3)–(6) must be explicitly included in the photoexcitation process, and were thus computed in L S coupling. Since xstar requires fine-structure partial photoionization cross-sections to determine rates, we decided to obviate the phototexcitation process in ions with Z 26 and 21 N 26 , thus underestimating their opacities, although the radiative and Auger decays of the L α , β and K α , β lines from these systems will be respectively taken into account in data types 50 and 86 (see Figure 3 and Figure 4).

3.2. Orbital Relaxation

As shown in [55,73], the photoionization of an inner-shell electron leads to a relaxation of the outer electrons of the ionized state due to a reduced screening of the atomic nuclear potential by the reduced inner electron cloud; i.e., an increased effective charge Z eff . For example, in the inner-shell photoionization of O i 1 s 2 2 s 2 2 p 4 to O ii 1 s 2 s 2 2 p 4 , the effective charge experienced by a 2p electron in the initial and final states is a complicated average screening including the other equivalent 2p electrons. Nevertheless, by using hydrogenic considerations, it is found that the electron orbital energy ( E n E R Z eff 2 / n 2 ) is more tightly bound for the relaxed 2p electron, and the electron density is concentrated closer to the nucleus ( r n a 0 n 2 / Z eff ).
To illustrate this phenomenon, we show in Figure 7 the outer 2p orbital in both neutral oxygen and its K-vacancy ionized daughter obtained from single-configuration Hartree–Fock calculations [74,75]. Clearly, the relaxed 2p orbital of the ionized O ii inner-shell state differs significantly from the neutral O i 2p orbital, a difference that needs to be accounted for to compute more accurate atomic transition properties.
In our hfr and autostructure structure calculations, orbital relaxation was taken into account by adopting different non-orthogonal orbital bases for each configuration. In most photoionization calculations, however, an orthonormal orbital basis is used because of the relative simplicity in computing transition and energy matrix elements. Relaxation then requires a substantial multi-configurational wave-function description including additional orbitals to account for the significant differences between initial- and final-state orbitals. Thus, additional care must be taken in computing inner-shell photoionization cross-sections when compared to outer-shell photoionization, as discussed more fully in [23]. Due to these difficulties, most large-scale computations of photoabsorption and photoionization cross-sections do not include orbital relaxation effects unless otherwise indicated.

4. Photoabsorption Cross-Section Benchmarks

Table A2 in Appendix C lists the reference sources for the computations of the photoabsorption cross-sections. Measurement of K absorption in the laboratory and ISM enables useful benchmarks of the theoretical cross-sections to improve spectral accuracy. In the following sections, we give a summary of these procedures.

4.1. Carbon Sequence

Calculations of the atomic data to model the K-shell photoabsorption by carbon ions were performed by [21], and a benchmark was obtained using Chandra X-ray high-resolution spectra [37]. The C ii K α “triplet” as well as the C iii K α and K β lines were identified in the X-ray spectra of four super-soft sources. C i was not modeled due to the presence of instrumental features at ∼43.6 Å.

4.2. Oxygen Sequence

K-shell photoabsorption cross-sections for O i–O vii were computed by [11], and comparisons of the theoretical resonance positions with astronomical observations and laboratory measurements were carried out by [32,33]. It was found that, due to a suspect experimental energy calibration scale, the X-ray astronomical observations provided a more reliable reference to perform the benchmark of the atomic data. For O i, the K-edge cross-section was re-examined in detail to yield a substantially improved version for the xstar atomic database [23].
In this exhaustive study, the K α ( 1 s 2 p ) resonance in O i was examined with R-matrix and Multi-Configuration Hartree–Fock (mchf) calculations, existing laboratory spectroscopic data, and several X-ray observational assessments (see Table 1). While practical R-matrix calculations for multi-electron systems are better tailored for computing scattering and atomic transition rates, they are less reliable in predicting the energy differences between atomic states as compared to limited bound-state structure calculations. (R-matrix calculations necessarily include all quasi-bound and continuum final states, limiting a practical improvement of the basis description.) The R-matrix K α energy was found to be ∼0.3 eV higher than the final assessed observational value of 527.37 eV [23]. More complex mchf calculations, which were focused only on the initial 1 s 2 2 s 2 2 p 4 state and final 1 s 2 s 2 2 p 5 state, converged downward to a value of 527.49 eV, in good agreement with the observations. On the other hand, the experiments performed at the Advanced Light Source (ALS) were reporting a value of 526.79 eV [76], a discrepancy of almost 0.6  eV. Considering the reliability of the observational calibration and the convergence trend of the mchf calculations, the X-ray observation was unconventionally chosen instead of the laboratory measurement as the final resonance position in the atomic absorption model.
A further consideration in assessing the reliability of the ALS measurements was that the absolute energy calibration relied on earlier molecular measurements with uncertainty estimates that seemed questionable. Since the O i study by [23], new laboratory measurements have been performed using an electron beam ion trap (EBIT) to determine the K α resonance position at 527.26 ± 0.04  eV [52], in good agreement with the observations and the converged mchf result. This new measurement suggests a recalibration of the ALS reference spectrum.

4.3. Neon

For neon, the xstar atomic database includes the photoabsorption cross-sections by [13]. A benchmark for Ne ii and Ne iii was performed with Chandra and XMM-Newton observations of the bright low-mass X-ray binaries (LMXBs) Cygnus X–2 and XTE J1817-330 [35]. Column densities were obtained for the cold (Ne i) and warm (Ne ii, Ne iii) components of the ISM.

4.4. Magnesium

New photabsorption cross-sections for Mg species were computed by [25]. X-ray XMM-Newton spectra of the LMXB GS 1826-238 were used to perform a benchmark of the data. The resulting ionization fractions indicate that Mg is predominantly ionized rather than in neutral form, as also shown by UV observations.

4.5. Silicon

High-resolution Chandra spectra in the Si K-edge region (6–7 Å) from 16 LMXBs were analyzed by [39]. The Si model included the Si i photoabsorption cross-sections computed by Gorczyca et al. (2020, in preparation), while those for the ionized species were taken from [18]. The absorption features identified in the spectra agreed with the theoretical atomic data, even though the individual resonances of the K α “triplet” could not be resolved. Although a model without the dust component was used, good data fits were obtained, highlighting the need for accurate modeling of the gaseous component before attempting to address the solid component [39].

5. High-Density Effects on Atomic Parameters

xstar needs to address spectra from high-density sources such as the X-ray reflection from the inner region of the accretion disks around compact objects (electron temperature T e 10 5 10 7  K and density n e 10 18 10 22  cm 3 ) [51,83]. This requires taking into account a series of atomic processes usually neglected in plasma photoionization models, which have been fully discussed by [46] and recently implemented in the code. We discuss in Section 5.1 and Section 5.2 two of these processes—continuum lowering and DR suppression—since they led to significant modifications of the database design and implementation.

5.1. Continuum Lowering

The density effects on the atomic thresholds (ionization potential and K edge), K α wavelengths and A-values, and Auger rates of oxygen and iron ions have been studied by [40,41,42,43,44,45]. The atom is assumed to be embedded in a weakly coupled plasma represented in the Multi-Configuration Dirac–Fock (mcdf) method with a time-independent Debye–Hückel screened Dirac–Coulomb Hamiltonian:
H D C D H = i c α i · p i + β i c 2 Z r i e μ r i + i > j 1 r i j e μ r i j ,
where r i j = | r i r j | and the plasma screening parameter μ (inverse of the Debye shielding length λ D ) is given in atomic units ( λ D in Bohr radius a 0 ; n e in a 0 3 ; and k T e in Hartree) by
μ = 1 λ D = 4 π n e k T e .
As shown in Figure 8, the density effects mainly cause a threshold lowering that increases with μ , which for the ionization potential can be approximated in eV by the universal formula [45] in terms of the effective charge Z eff = Z N + 1
Δ E 0 = ( 26.30 ± 0.08 ) μ Z eff .
In xstar continuum lowering leads to density-dependent targets with reduced numbers of levels, which must be derived at each density point causing a severe overhead at runtime. This bottleneck can be mitigated by pre-computing a μ  grid of atomic databases but at the price of increasing the database volume significantly (∼10 GB), thus hampering the standard package downloading procedure. Consequently, deployment strategies are being considered in terms of cloud virtual machines and containers.

5.2. DR Suppression

Rigorous collisional–radiative plasma modeling shows that DR is suppressed at high densities [84]. As mentioned in Section 2, RR and DR rates are treated separately in the xstar database using the AMDPP tabulation [64,65], which allows an estimate of the DR-rate density attenuation using the analytic formulae by [48,49]. The highly approximate value of the DR suppression factors has nevertheless been stressed by [49], as well as the need for detailed collisional–radiative calculations. The intermediate density range ( 10 log n e 18  cm 3 ) is expected to be the most susceptible since, at the higher densities ( log n e > 18  cm 3 ), DR is practically negligible.
It is worth mentioning that in the unified treatment of electron–ion recombination [85,86], where the total recombination rate is obtained directly from a Maxwellian average of the photoionization cross-section through the Milne relation, it would be more difficult to evaluate the DR density effects. It would imply a rigorous approach to introduce plasma interactions in the Hamiltonian, such as those described for continuum lowering (see Section 5.1), followed by large-scale computations of photoionization cross-sections and fitting procedures to obtain a set of interpolation formulae.
To characterize the plasma effects caused by DR suppression, we have compared the heating and cooling rates in the ionization-parameter–temperature plane:
3 log ( ξ ) 3 erg cm s 1 and 3.5 log ( T ) 7.5 K
for xstar models with and without metal DR. We assume a power-law spectrum, density n e = 10 10  cm 3 , and luminosity L = 10 32  erg s 1 (see Figure 9). The larger temperature differences occur in the interval 1 log ( ξ ) 3 , where T is lower by as much as 40% when DR is excluded, and are caused by increases in the Fe average charge at log ( ξ ) 1.3 and log ( ξ ) 2.5 . As shown in Figure 9 (right panel), the DR/noDR heating-rate ratio reaches a maximum at ( log ( ξ ) , log ( T ) ) ( 1 , 6.3 ) , where the dominant species with and without DR are Fe xiii and Fe xvii, respectively. On the other hand, the cooling-rate ratio reaches a maximum at ( log ( ξ ) , log ( T ) ) ( 1 , 5.4 ) , caused by a higher oxygen average charge when DR is excluded.
Concerning the Fe anomalous overabundance derived from the reflection spectra of accretion disks [51], we have computed the X-ray spectra reflected from an optically thick atmosphere using the xillver code [87,88]. The atmosphere is assumed to have plane-parallel geometry and a constant gas density of log ( n e ) = 18  cm 3 . The ionization parameter is set to ξ = 4 π F x / n e = 32  erg cm s 1 implying an illuminating flux of F x = 3 × 10 18  erg cm 2  s 1 . The ionization balance is calculated self-consistently using the routines and atomic data from xstar, setting all elemental abundances to their solar values. It is found that metal DR suppression leads to significant changes in the ionization state of the gas resulting in more ionized species. As shown in Figure 10, there is a strong enhancement of the Fe xvii L lines and Fe xviii K lines.
It is worth mentioning that continuum lowering is not taken into account in this xillver test calculation since it is expected to have a noticeable impact only at densities n e > 10 19  cm 3 . The calculation was performed at n e = 10 18  cm 3 to show that DR effects become conspicuous at lower densities. Further characterization of high-density effects are reported in [46].
Density effects on DR and the high-n levels through continuum lowering are taken into account in the xstar database via the rates into one or two superlevels for each ion. The recombination rate onto the superlevel is tabulated on a density–temperature grid and stored in data types 70 and 99 (see Appendix A). For the former, rates are tabulated on a log scale, while for the latter, they are tabulated directly. The total recombination rate onto the ion is the sum of the rate onto the superlevels plus the rates onto the other levels that are treated explicitly; these are called the “spectroscopic levels”. The recombination rates onto the superlevels are chosen such that, when they are used to calculate a recombination rate and all the rates are summed up, the total recombination rate for the superlevel(s) plus the spectroscopic levels is equal to the total rate taken from one of the AMDPP compilations [64] where available or from [89] otherwise. We apply density-dependent suppression factors to the recombination rates onto the superlevel to take into account density effects on radiative and dielectronic recombination. They are derived from a general formula that depends on the isoelectronic sequence, charge, density, and temperature [48,49].
The superlevels decay directly to the ground level without the emission of any observable cascade radiation for ions with three or more electrons. The exception is the decay of members of the H and He isoelectronic sequences, for which we use explicitly calculated cascade matrix calculations to treat the decay of the superlevels to the spectroscopic levels [90,91]. The superlevels are chosen to have energies close to the continuum. We include both radiative and collisional transitions to the ground level (and other levels in the case of H- and He-like ions).

6. Database Curation

The recent inclusion in the xstar database of atomic parameters for odd-Z elements, Z = ( 9 , 11 , 15 , 17 , 19 ) , and trace elements with Z = ( 21 25 , 27 , 29 , 30 )  [26,27,28,29] practically doubled the database volume when compared with previous versions that used hydrogenic scaling for these elements, bringing to the fore a series of curatorial problems. The database upgrade was thus lengthy and demanding in detail. In Section 6.1, Section 6.2, Section 6.3 and Section 6.4, we give a brief account of some the issues encountered and steps taken to solve them.

6.1. Atomic Models

As shown in Table 9 of [26], the target models for ions with atomic numbers Z = ( 21 25 , 27 , 29 , 30 ) and electron number 19 N Z 2 include levels from the valence configurations μ = 3 d x 4 s y and the L- and K-vacancy configurations 3 p 1 μ + 1 , 2 p 1 μ + 1 , and 1 s 1 μ + 1 . The total number of target levels in each of these species is larger than 500 and in some cases (e.g.,  N = 23 ) close to 2000. It was found that to model astrophysical plasmas in xstar with such large atomic representations was impractical and with little tangible gains. We thus excluded the levels belonging to μ configurations containing 4s orbitals—specifically, y = 0 in μ for N > 19 . This decision led to a reduction in the number of target levels to a few hundreds and implied extensive level renumbering for these sequences.

6.2. Metastable Levels

As discussed in Section 2, the xstar atomic models display occasional low-lying metastable levels (e.g., the 3 s 1.3 p 1.3 P _ 0 level in Mg-like ions) with no radiative decay transitions. Since metastable lavels in low valence-shell configurations are astrophysically important, alternative decay mechanisms such as collisional de-excitation must be specified for such levels. For instance, we computed a large number of distorted-wave collision strengths with autostructure for transitions within the valence-shell configurations of trace-element ions (see record with ID = mb01 and code = 1150 in Table A2). However, for ions with electron number N > 19 and ground configuration 3 d x , atomic models with a few hundred levels must be considered (see Section 6.1), which leads to the appearance of unexpected metastable levels in the 3p 1 configuration at fairly high energies. Most of these levels were detected by running the code to look for levels with anomalously large populations, which were then removed by hand followed by level reordering.

6.3. Database Pointers

As previously mentioned, the raw xstar database consists of a list of ASCII files labeled d i o n t 0 d t , where i o n is the ion identifier and d t the data type. When the package is released, these files are read into three one-dimensional arrays (integer, real, and character variables) and a set of pointers to be finally structured into the FITS file atdb.fits. At runtime, these arrays are read into main memory to be accessed according to the user model. This data structure arrangement is efficient when uploading from disk and during processing, but can be cumbersome in database updating when new record types are created. Proficient nursing and testing are required before the new version is released.

6.4. Metadata

The xstar database compiles atomic data computed to order and from a myriad of other data sources. The data provenance of the first release was detailed in [2], but since then, the database has undergone several major upgrades that have not been formally registered within its file structure. One of the aims of the present paper is to create an up-to-date data provenance blueprint to be included in the database using the FITS metadata facilities.
In Table A2 of Appendix C, we have tabulated an up-to-date list of data source references including the original four- and five-letter mnemonics used for identification in [2] and a new four-integer code to be inserted in every database record. For this purpose, we will adopt the third integer (currently displaying a zero) of the record header ( d t , r t , 0 , N flt , N int , N chr ) described in Section 2 to denote this data provenance code.

6.5. Relational Integrity

As described in Section 2, the data in each d i o n t 0 d t file are self-contained to simplify and speed up the calculation of a wide variety of rates at runtime, but as a result, this scheme obviates the relational integrity of the database. For instance, the wavelengths displayed in file d i o n t 050 (see Figure 3) are not automatically recalculated if level energies are updated in file d i o n t 006 (Figure 2). The level attributes specified in the latter file are also repeated in several other data types, e.g., the K-vacancy level energy in file d i o n t 086 (Figure 4). The extent of this complication is further illustrated in the list of data types in Appendix A. Data type 14 lists the ionization potential for every ionic species, but this datum is also contained in the floating-point tuple of every level in data type 06. Some of the level integer attributes of the latter data type are included in data types dealing with the effective ion charge (57), photoionization cross-sections (49, 53, 85, and 88), superlevel recombination rates (70 and 99), and satellite-level autoionization rates (72).
This database model can therefore make updating an involved and error-prone process, which can be mitigated by adopting an SQL-based relational database engine and devising a utility to perform the transcription to the atdb.fits file. Despite the wide variety of data types (see Appendix A) and record lengths, this is possible as shown by the development of the uaDB database (see Section 8). However, this scheme would hamper the generation of new data types, which is central to the xstar data curation strategy (see Section 2), and also lead to a more voluminous atdb.fits file. This proposition is being considered in the xstar development roadmap.

6.6. Atomic Data Accuracy

As astronomical observations become ever more accurate (see, for example, Figure 1), the reliability of the atomic data used in spectral modeling is under recurrent scrutiny, and hence, data assessment activities are key. In this respect, we have seen three schemes: by a reputed institution such as NIST [92]; by appointed review panels (e.g., VAMDC [93]), and by an open community of both data producers and users (e.g., AtomPy [94]). Our research group comprises both computational atomic physicists and astrophysical modelers, so we favor the latter. However, it has not been easy to determine accuracy ratings in our datasets for K-line diagnostics (Section 3) due to the lack of laboratory measurements and despite adopting a multi-code computational strategy. As described in Section 4, we have pursued benchmarks of the photoabsorption cross-sections using astronomical observations of ISM K absorption, but they led to a polemic between laboratory spectroscopists and observational astronomers that was fortunately settled.
We have mentioned in Section 2 that our main concern regarding the xstar database upgrade has been the complete datasets for K-shell processes in ions with Z 30 , at the expense of neglecting those for the valence-shell that in general imply data collection rather than new computations. It may be seen in Appendix C that most of the current datasets for valence-shell transitions go back to the original version [2] then demanding an extensive overhaul. This is one of our main priorities in the database curation.

7. ISMabs

Although the main components of the ISM are neutral ionic species, the also perceived existence of charged ions prompted us to develop a new ISM X-ray absorption model referred to as ISMabs [35]. It includes atomic data for singly and doubly ionized species of cosmically abundant elements, namely H, He, C, N, O, Ne, Mg, Si, S, Ar, Ca, and Fe, in addition to the neutral systems. This model allows ion column densities to be determined directly, and the inclusion of ionized species has led to improved spectral fits as compared to those only considering neutral systems. The completeness and accuracy of the relevant atomic data are crucial to avoid misidentification and misinterpretation of the absorption features detected in analyses of ISM X-ray absorption spectra relying on fits of Gaussian profiles. We have performed with this model a detailed study of X-ray absorption in the local ISM by analyzing spectra from 24 Galactic sources. We have estimated the fractions of neutral and singly and doubly ionized species of O, Ne, and Fe, confirming the dominance of the cold component and a pervasive low degree of ionization [36].

8. Universal Atomic Database (uaDB)

The collection of atomic data in xstar is comprised of data in many different formats from many different sources. To assist in disseminating these data to the larger community and to account for the addition of new data, the Universal Atomic Database (uaDB) was developed. The uaDB is a mySQL database hosted at NASA/GSFC, accessed via its website (https://heasarc.gsfc.nasa.gov/uadb/index.php), and data can be queried based on a number of properties such as ion and data type. The web page also contains tools to help assess what data are available across ions or data types. One such tool displays the type of data available on a grid of element vs. ion (see Figure 11) while another displays all atomic lines present in a given wavelength range.
The main features of uaDB are:
  • Each dataset contains a reference to the source paper(s);
  • Data are stored in the original format and units as published;
  • Multiple entries can exist for the same quantity (e.g., theoretical vs. measured energy levels);
  • The coupling scheme of atomic states is built into the database structure;
  • Data collections exist that define a complete model; no duplicate data exist within a collection.
All data in uaDB have a reference to the data source. The reference consists of a link to the journal article where the data were obtained and to its entry in NASA/ADS. The user is never more than a couple mouse clicks away from reading the source article. Data are entered into uaDB in the same format and with the same units as published. This means that, for example, some energy levels are stored in eV and others in cm 1 . The uaDB understands these units and can internally convert them for data querying in an energy range. The uaDB can also convert between quantities that differ by a statistical weight (e.g., f-values and g f -values) since it understands the level structure. We do not modify published data in order to be as true as possible to the original source and to help identify transcription errors. However, this can make it more difficult to compare data from different sources. To help compare data, tools are developed that can perform the necessary conversions.
When compiling a set of atomic data for modeling purposes, it is important to have coverage; that is, for each state included in the model, a full set of data connecting that state with other states is necessary. If there are data allowing the population of a state to increase but no data to decrease it, then the model as a whole will be flawed. Furthermore, if the atomic model encompasses a large set of ions, it is difficult to obtain atomic data from the literature of equal quality. We often need to combine level-resolved data with term-resolved data to achieve coverage for an ion. To help with these difficulties, uaDB has different coupling schemes built into its structure and is aware of how different coupling schemes relate. Thus, when combining L S states with L S J , uaDB can detect if any states are being double-counted. There are also routines for converting data from one coupling scheme to another based on simple statistical weighting.
The uaDB currently contains the complete set of data included in xstar as well as the level data from chianti. To assist in distinguishing these data, uaDB defines collections. A collection is a dataset where there is no duplication of values for the same quantity. Currently, in uaDB the collection is just a label that can be used to constrain a query, but in the future collections could be created dynamically to construct a custom model. Collections are not part of the xstar development program, but have been widely used in, for instance, the PyNeb Python package to model emission lines in gaseous nebulae and, more recently, to assess atomic data accuracy [95,96].

9. Discussion and Conclusions

The present report describes in detail the updated version of the underlying atomic database of the xstar photoionization code to replace the original implementation of [2]. We include complete lists of the current rate types, data types (Appendix A), ionic models (Appendix B), and source references (Appendix C). We summarize the systematic calculations performed to characterize the metal K-lines for astronomical X-ray spectral modeling and recent extensions of the database, namely continuum lowering and dielectronic recombination suppression, to deal with high-density ( n e 10 22  cm 3 ) plasmas. We have tried to convey the intricacies in the development and maintenance of an application-based atomic database, which go way beyond the compilation of the data and lead to curatorial and integrity problems caused by the ever-increasing dataset volumes.
An important finding is that the atomic models used to compute the data are not necessarily suitable for plasma modeling, particularly for ionic species with electron number N > 19 . Level-number trimming was necessary to implement functional atomic representations for such systems, although their soundness in non-LTE modeling deserves further attention as the appearance of unexpected metastable levels becomes an issue.
A further point worth emphasizing is that the xstar atomic database and output files are structured under the FITS format specifications, which has been established as the standard for astronomical datasets associated to the new generation of telescopes. We believe that this compliance will reduce user stress in intense data processing, which is bound to take place at the data repositories rather than locally.
We have briefly recounted two projects that have emerged from the xstar database, namely the ISMabs absorption model and the uaDB database. The former has become an important initiative as it has led to benchmarks of the atomic data with observations and laboratory measurements; however, it has not been a smooth process as unexpected inaccuracies in the calibration of the laboratory wavelength scale were detected and finally resolved [23,52]. This outcome has given us some confidence in the identification of the spectral features and the accuracy of our atomic data. The uaDB database model will be helpful in future projects to disseminate the larger atomic data volumes required in non-LTE modeling.
A question that remains unanswered is whether our efforts to ensure accuracy and completeness for the xstar database are sufficient to exploit the high resolution and sensitivity expected from the microcalorimeter-based spectrometers aboard the new X-ray space telescopes. The encouraging fit of the showcase hitomi spectrum of Figure 1 required basic data manipulation, but at least it was conclusive. Therefore, we envision further database refinement in the near future, especially improved datasets for valence-shell processes, to provide a reliable modeling platform for high-resolution spectroscopy in the new era. In this respect more laboratory benchmarks would be welcome.

Author Contributions

Conceptualization, C.M., M.A.B. and T.R.K.; methodology, C.M., E.G., J.A.G., J.D., M.A.B., M.C.W., P.P., P.Q., T.R.K. and T.W.G.; software, C.M., E.G., J.A.G., M.A.B., M.C.W., P.P., T.R.K. and T.W.G.; validation, J.A.G., M.A.B., P.P. and T.R.K.; formal analysis, C.M., E.G., J.A.G., J.D., M.A.B., M.C.W., P.P., P.Q., T.R.K. and T.W.G.; investigation, C.M., E.G., J.A.G., J.D., M.A.B., M.C.W., P.P. and T.R.K.; validation, C.M., M.A.B., P.P. and T.R.K.; resources, M.A.B., P.Q., T.R.K.; data curation, C.M., M.A.B., P.P. and T.R.K.; writing—original draft preparation, C.M., E.G., M.C.W. and T.R.K.; writing—review and editing, C.M., E.G., J.A.G., J.D., M.A.B., M.C.W., P.P., P.Q., T.R.K. and T.W.G.; visualization, C.M., M.C.W. and T.R.K; supervision, T.R.K.; project administration, M.A.B., P.Q. T.R.K.; funding acquisition, J.A.G., P.P., P.Q. and T.R.K. All authors have read and agreed to the published version of the manuscript.

Funding

J.D. is a Research Fellow of the Belgian Fund for Research in Industry and Agriculture (FRIA). P.P. & P.Q. are, respectively, Research Associate and Research Director of the Belgian Fund for Scientific Research F.R.S.–FNRS. Financial support from these organizations as well as from the NASA Astrophysics Research and Analysis Program (grants 12-APRA12-0070; 80NSSC17K0345; and 80NSSC19K0586) and the Alexander von Humboldt Foundation are gratefully acknowledged.

Data Availability Statement

The xstar modeling code, the uaDB database, and the ISMabs model are available from the website of the NASA High Energy Astrophysics Science Archive Research Center (HEASARC, https://heasarc.gsfc.nasa.gov/).

Acknowledgments

We are grateful to Anil. K. Pradhan (Ohio State University) for access to and space-time allocation at the Ohio Supercomputer Center and also to the NASA Private Cloud at GSFC for access and virtual machine installation, especially to Michael D. Moore for invaluable support.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ADASAtomic Data and Analysis Structure
ADSAstrophysics Data System
ALSAdvanced Light Source
AMDPPAtomic and Molecular Diagnostic Processes in Plasmas
FITSFlexible Image Transport System
GSFCGoddard Space Flight Center
HZBHelmholtz-Zentrum Berlin
IPNASInstitut de Physique Nucléaire, Atomique et de Spectrométrie
LTELocal Thermodynamic Equilibrium
NASANational Aeronautics and Space Administration
NISTNational Institute of Standards and Technology
PP95Atomic Data for the Analysis of Emission Lines
SQLStructured Query Language
WSRCWisconsin Synchrotron Radiation Center

Appendix A. Data Types

This appendix lists the data types of the xstar database. Unless otherwise stated, the temperature T is given in 10 4 K units. The record attributes are: r 1 , r 2 , ⋯ for the real variables; i 1 , i 2 , ⋯ for the integer variables; and s 1 , s 2 , ⋯ for the character strings. The following notation is used: i o n is the ion identifier; Z the atomic number; N the electron number; E ( ) the ionization potential; E ( i ) the ith level energy; n the principal quantum number; ν the effective quantum number; S the total spin quantum number; L the total orbital angular momentum quantum number; J the total angular momentum quantum number; λ the radiative transition wavelength; g f ( i , k ) the weighted oscillator strength; and A ( k , i ) the transition rate.
01 .
Radiative recombination rate coefficient of N-electron recombined ion [89,97]: r 1 = A rad (cm 3 s 1 ); r 2 = η ; i 1 = i o n N .
02 .
H 0 charge exchange rate coefficient of N-electron recombined ion [98]: r 1 = a ( 10 9 cm 3 s 1 ); r 2 = b ; r 3 = c ; r 4 = d ; r 5 = T 1 (K); r 6 = T 2 (K); r 7 = Δ E / k ( 10 4 K); i 1 = i o n N ; s 1 = recombining ion identifier.
06 .
Data attributes of the ith level of N-electron ion: r 1 = E ( i ) (eV); r 2 = ( 2 J + 1 ) ; r 3 = ν (effective quantum number); r 4 = E ( ) (eV); i 1 = n ; i 2 = ( 2 S + 1 ) ; i 3 = L ; i 4 = Z ; i 5 = i ; i 6 = i o n N ; s 1 = level assignment (electron configuration and spectroscopic term).
07 .
Dielectronic recombination rate coefficient of N-electron recombined ion [89,97]: r 1 = A di (cm 3 s 1 K 3 / 2 ); r 2 = B di ; r 3 = T 0 ; r 4 = T 1 ; i 1 = i o n N .
14 .
Ionization potential of N-electron ion: r 1 = E ( ) (eV); i 1 = Z N + 1 ; i 2 = Z ; i 3 = i o n N ; s 1 = ion identifier.
22 .
Dielectronic recombination rate coefficient of the N-electron recombined ion [99]: r 1 = a ; r 2 = b ; r 3 = c ; r 4 = d ; r 5 = f ; i 1 = i o n N .
30 .
Total radiative recombination rate (hydrogenic) for N-electron recombined ion [100]: i 1 = Z ; i 2 = i o n N .
38 .
Total radiative recombination rate coefficient of N-electron recombined ion [http://amdpp.phys.strath.ac.uk/tamoc/DATA/RR/]: r 1 = A (cm 3 s 1 ) ; r 2 = B ; r 3 = T 0 (K); r 4 = T 1 (K); r 5 = C ; r 6 = T 2 (K); i 1 = Z ; i 2 = N 1 ; i 3 = M ; i 4 = W ; i 5 = i o n N .
39 .
Total dielectronic recombination rate coefficient of N-electron recombined ion [http://amdpp.phys.strath.ac.uk/tamoc/DATA/DR/]: r 1 rj max = ( C ( j ) , j = 1 , j max ) (cm 3 s 1 K 3 / 2 ); rj max + 1 rj 2 max = ( T ( j ) , j = 1 , j max ) (K); i 1 = Z ; i 2 = N 1 ; i 3 = M ; i 4 = W , i 5 = i o n N .
49 .
Partial photoionization cross-section of i N th level of the N-electron ion leaving the ( N 1 )-electron ion in the k N 1 th level: r 1 rj 2 max = ( E ( j ) , σ ( E ( j ) ) , j = 1 , j max ) (Energy in Ryd relative to E ( ) , cross-section in Mb); i 1 = n ; i 2 = L ; i 3 = 2 J ; i 4 = Z ; i 5 = k N 1 ; i 6 = i o n N 1 ; i 7 = i N ; i 8 = i o n N .
50 .
Line ( k i ) radiation rates of N-electron ion: r 1 = λ (Å); r 2 = g f ( i , k ) ; r 3 = A ( k , i ) (s 1 ); i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
51 .
Electron-impact effective collision strength for the k i transition of N-electron ion (CHIANTI fit [61,101]): r 1 = Δ E (Ryd); r 2 = C ; r 3 r 7 = ( Υ red ( j ) , j = 1 , 5 ) (reduced effective collision strength); i 1 = i t (transition type); i 2 = i (lower level); i 3 = k (upper level); i 4 = Z ; i 5 = i o n N .
53 .
TOPbase partial photoionization cross-section (resonance averaged) of i N th level of the N-electron ion leaving the ( N 1 )-electron ion in the k N 1 th level: r 1 rj 2 max = ( E ( j ) , σ ( E ( j ) ) , j = 1 , j max ) (Energy in Ryd relative to E ( ) , cross-section in Mb); i 1 = n ; i 2 = L ; i 3 = 2 J ; i 4 = Z ; i 5 = k N 1 ; i 6 = i o n N 1 ; i 7 = i N ; i 8 = i o n N .
54 .
Radiative transition probability A k i for the k i transition of N-electron ion computed by quantum defect theory (or hydrogenic): r 1 = 0.0 E+0; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
56 .
Electron-impact effective collision strengths for the k i transition of N-electron ion: r 1 rj max = ( log T e ( j ) , j = 1 , j max ) (K); rj ( max + 1 ) rj ( 2 max ) = ( Υ ( T e ( j ) ) , j = 1 , j max ) (effective collision strength); i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
57 .
Effective ion charge for ith level of N-electron ion used in collisional ionization rates: r 1 = Z eff ; i 1 = n ; i 2 = L ; i 3 = 2 J ; i 4 = i ; i 5 = i o n N
59 .
Partial photoionization cross-section of i N th level of the N-electron ion leaving the ( N 1 )-electron ion in the k N 1 th level [102]: r 1 = E ( t h ) (eV); r 2 = E ( 0 ) (eV); r 3 = σ ( 0 ) (Mb); r 4 = y ( a ) ; r 5 = P ; r 6 = y ( w ) ; i 1 = N ; i 2 = n (shell principal quantum number); i 3 = l (subshell orbital quantum number); i 4 = k N 1 ; i 5 = i o n N 1 ; i 6 = i N ; i 7 = i o n N ; s 1 = shell–ion identifier.
60 .
Analytic fits for effective collision strengths in H-like ions [103]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = n p h o t ; i 4 = i o n N ; s 1 = Transition.
62 .
Analytic fits for effective collision strengths in H-like ions [103]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = 1 ; i 4 = i o n N ; s 1 = Transition.
63 .
Collisional transition probability C i k for N-electron ion computed by quantum defect theory (or hydrogenic): i 1 = 1 i 2 = i (lower level); i 3 = k (upper level); i 4 = Z ; i 5 = i o n N .
66 .
Fits to fine-structure collision strengths for He-like ions [104]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
67 .
Analytic fits for effective collision strengths in He-like ions [105]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
68 .
Analytic fits for effective collision strengths in He-like ions [106]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
69 .
Fits to L S collision strengths for He-like ions [104]: r 1 rj max = coefficients; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
70 .
Coefficients for recombination and photoionization cross-sections of superlevels: r 1 rj nd = ( n e ( j ) , j = 1 , j n d ) (cm 3 ); rj nd + 1 rj nd + nt = ( T e ( j ) , j = 1 , j n t ) ; rj nd + nt + 1
rj nd + nt + nt nd = ( ( log α ( j , j ) , j = 1 , j n t ) , j = 1 , j n d ) ; rj nd + nt + nt nd + 1
rj nd + nt + nt nd + 2 nx = ( E ( j ) , σ ( j ) , j = 1 , j n x ) (Ryd, Mb); i 1 = n d ; i 2 = n t ; i 3 = n x ; i 4 = n ; i 5 = L ; i 6 = 2 S + 1 ; i 7 = Z ; i 8 = k N 1 ; i 9 = i o n N 1 ; i 10 = i N ; i 11 = i o n N .
71 .
Radiative transition rates from superlevels to spectroscopic levels: r 1 rj nd = ( n e ( j ) , j = 1 , j n d ) (cm 3 ); rj nd + 1 rj nd + nt = ( T e ( j ) , j = 1 , j n t ) ; rj nd + nt + 1 rj nd + nt + nt nd = ( ( A ( j , j ) , j = 1 , j n t ) , j = 1 , j n d ) (s 1 ); r nd + nt + nt nd + 1 = λ (Å); i 1 = n d ; i 2 = n t ; i 3 = i (lower level); i 4 = k (upper level); i 5 = Z ; i 6 = i o n N .
72 .
Autoionization rates for satellite levels: r 1 = A a ( k , i ) (s 1 ); r 2 = E ( k ) (eV above ionization limit); r 3 = ( 2 J + 1 ) ; i 1 = ( 2 S + 1 ) ; i 2 = L ; i 3 = k (level); i 4 = i (continuum level); i 5 = Z ; i 6 = i o n N ; s 1 = level configuration.
73 .
Fit to effective collision strengths for satellite levels of He-like ions [107]: r 1 r 7 = fit coefficients; i 1 = i (lower level); i 2 = j (upper level); i 3 = Z ; i 4 = i o n N .
74 .
Delta functions to add to photoionization cross-sections to match ADAS DR rates: r 1 = E ( ) (Ryd); r 2 rj m + 1 = ( E ( j ) , j = 1 , j m ) (Ryd); rj m + 2 rj 2 m + 1 = ( f ( j ) , j = 1 , j m ) (cm 2 ); i 1 = n ; i 2 = L ; i 3 = 2 S + 1 ; i 4 = Z ; i 5 = k N 1 ; i 6 = i o n N 1 ; i 7 = i N ; i 8 = i o n N .
75 .
Autoionization rates for Fe xxiv satellites [5]: r 1 = A a ( k , i ) (s 1 ); r 2 = E ( k ) (eV above ionization limit); i 1 = i o n N , i 2 = k N ; i 3 = i o n N 1 ; i 4 = i N 1 ; i 5 = i o n N .
76 .
Two-photon radiation rate for ( k i ) transition of N-electron ion: r 1 = A ( k , i ) (s 1 ); i 1 = i (lower level); i 2 = k (upper level); i 3 = N ; i 4 = i o n N ; s 1 = transition.
77 .
Collision transition rates from superlevels to spectroscopic levels: r 1 rj nd = ( n e ( j ) , j = 1 , j n d ) (cm 3 ); rj nd + 1 rj nd + nt = ( T e ( j ) , j = 1 , j n t ) ; rj nd + nt + 1 rj nd + nt + nt nd = ( ( C ( j , j ) , j = 1 , j n t ) , j = 1 , j n d ) (s 1 ); rj nd + nt + nt nd + 1 = λ (Å); i 1 = n d ; i 2 = n t ; i 3 = i (lower level); i 4 = k (upper level); i 5 = Z ; i 6 = i o n N .
81 .
Collision strengths for Fe xix [108]: r 1 = Υ ( k , i ) ; i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
82 .
Decay rates for Fe UTA [109]: r 1 = λ (Å); r 2 = E ( k ) (eV); r 3 = g f ( i , k ) ; r 4 = A r ( k , i ) (s 1 ); r 5 = A a ( k , i ) (s 1 ); i 1 = i (lower level); i 2 = k (upper level); i 3 = i o n N .
83 .
Level data for Fe UTA [109]: r 1 = E ( i ) (eV); r 2 = ( 2 J + 1 ) ; r 3 = 0.0 ; r 4 = 0.0 ; i 1 = 1 ; i 2 = i (level); i 3 = i o n N ; s 1 = level configuration assignment.
85 .
Photoionization cross-sections for Fe ions obtained by summation of resonances near the K edge [3]: r 1 = Z eff ; r 2 = E t h (Ryd); r 3 = f ; r 4 = γ ; r 5 = scaling factor; i 1 = n ; i 2 = L ; i 3 = 2 J ; i 4 = Z ; i 5 = k N 1 ; i 6 = i o n N 1 ; i 7 = i N ; i 8 = i o n N .
86 .
Auger and radiative widths of k N th K-vacancy level: r 1 = E ( k N ) (eV, relative to E ( ) ); r 2 = A a ( k N ) (s 1 ); r 3 = A a ( k N , i N 1 ) (s 1 ); r 4 = A r ( k N ) (s 1 ); i 1 = i N 1 ; i 2 = k N ; i 3 = Z ; i 4 = i o n N 1 ; i 5 = i o n N .
88 .
Photoionization cross-section damped excess of i N th level of the N-electron ion leaving the ( N 1 )-electron ion in superlevel_[K] k N 1 : r 1 rj 2 max = ( E ( j ) , σ ( E ( j ) ) , j = 1 , j max ) (Energy in Ryd relative to E ( ) , cross-section in Mb); i 1 = n ; i 2 = L ; i 3 = 2 J ; i 4 = Z ; i 5 = k N 1 ; i 6 = i N ; i 7 = i o n N .
91 .
APED line ( k i ) radiation rates [110]: r 1 = λ (Å); r 2 = 0.0 ; r 3 = A ( k , i ) (s 1 ); i 1 = i (lower level); i 2 = k (upper level); i 3 = Z ; i 4 = i o n N .
92 .
APED collision strengths [110]: r 1 = T e ( 1 ) (K); r 2 = T e ( j m a x ) (K); r 3 rj max + 2 = ( T e ( j ) , j = 1 , j m a x ) (K); rj max + 3 rj 2 max + 2 = ( Υ ( j ) , j = 1 , j m a x ) ; i 1 = i (lower level); i 2 = k (upper level); i 3 = 113 ; i 4 = Z ; i 5 = i o n N .
95 .
Collisional ionization rates for N-electron ion [111]: r 1 = E ( t h ) (eV); r 2 = T 0 (K); r 3 rj max + 2 = ( ρ ( j ) , j = 1 , j m a x ) (effective collision strength); i 1 = i (level); i 5 = i o n N .
98 .
Electron-impact effective collision strengths for the k i transition of the N-electron ion (CHIANTI fit [61,101]): r 1 = Δ E (Ryd); r 2 = C ; r 3 rj max + 2 = ( Υ red ( j ) , j = 1 , j m a x ) (reduced effective collision strength); i 1 = i ; i 2 = k ; i 3 = i t (transition type); i 4 = i o n N .
99 .
Coefficients for recombination and photoionization cross-sections of superlevels: r 1 rj nd = ( n e ( j ) , j = 1 , j n d ) (cm 3 ); rj nd + 1 rj nd + nt = ( T e ( j ) , j = 1 , j n t ) ; rj nd + nt + 1 rj nd + nt + nt nd = ( ( α ( j , j ) , j = 1 , j n t ) , j = 1 , j n d ) ; rj nd + nt + nt nd + 1
rj nd + nt + nt nd + 2 nx = ( E ( j ) , σ ( j ) , j = 1 , j n x ) (Ryd, Mb); i 1 = n d ; i 2 = n t ; i 3 = n x ; i 4 = n ; i 5 = L ; i 6 = 2 S + 1 ; i 7 = Z ; i 8 = k N 1 ; i 9 = i o n N 1 ; i 10 = i N ; i 11 = i o n N .

Appendix B. Ion Models

Table A1. Number of energy levels included in the xstar ion models.
Table A1. Number of energy levels included in the xstar ion models.
Z123456789101112131415
N
1333333333333333333333333333333
2 46323232565824152545258525852
3 33326434322432243224322
4 3322535343794379437943
5 3225816353163531635316353
6 8617953795379537953
7 823434343434633434
8 193191919571919
9 311877777
10 4346464650
11 32446042
12 345731
13 31855
14 33
15 3
Z161718192021222324252627282930
N
1333333333333333333333333333333
2585238524652525252525052435252
3432243224322222222225222532222
47943794379434343434319943794343
516353163531635353535353551531635353
67953535353535353535362353535353
73434343434343434343474434343434
81919191929191919191963919671919
97777777777340711777
104650725078505050505027650985050
11604220842208424242424285422164242
125331153311493131313131146311713131
13765520855208555555555555552355555
14827124971249717171717168712697171
15315521155211555555555578552195555
163314331143313131313170311453131
17 362106210101010106210631010
18 33354545454543154405454
19 35221221221221221170221143221221
20 33225825825825854258258258258
21 38444144144155441441441441
22 3353753733537537537537
23 316346370463463463463
24 316073295295295295
25 3197128128128128
26 33404040
27 3333
28 333
29 33
30 3

Appendix C. Atomic Data Provenance

Table A2. Data provenance including the chianti [61], PP95 [63], and AMDPP [64,65] compilations. rt = rate type. dt = data type.
Table A2. Data provenance including the chianti [61], PP95 [63], and AMDPP [64,65] compilations. rt = rate type. dt = data type.
IDCodertdt ( Z ; N ) References
ac811000356 ( 8 ; 1 )  [112]
ac921001356 ( 2 ; 1 )  [113]
ak911002356 ( 6 ; 1 )  [114]
ak91a1003356 ( 10 ; 1 )  [115]
ak921004356 ( 14 ; 1 )  [116]
ak92a1005356 ( 20 ; 1 )  [117]
ak931006356 ( 26 ; 1 )  [118]
ba001007450 ( 28 ; 2 )  [91]
bm981008356 ( 26 ; 15 )  [119], chianti ([120,121])
bn001009356 ( 26 ; 20 )  [122], chianti ([123])
bp981010356 ( 26 ; 25 )  [90]
ch001011351 ( 10 ; 7 ) , ( 20 ; 8 ) , ( 26 ; 16 ) , ( 28 ; 14 ) chianti
ch001012450 ( 6 , 7 , 8 ; 4 ) , ( 12 , 14 ; 5 ) , ( 20 ; 8 ) ,chianti
( 26 ; 6 , 12 , 14 18 ) , ( 28 ; 14 )
ch011013351 ( 6 , 7 ; 3 ) chianti ([124])
ch011014450 ( 6 , 7 ; 3 ) chianti ([125,126])
ch021015450 ( 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 , 28 ; 3 ) chianti ([126,127])
ch021016351 ( 8 , 10 , 12 , 14 , 16 , 18 , 20 , 28 ; 3 ) chianti ([127])
ch041017351 ( 6 ; 4 ) chianti ([128,129])
ch041018450 ( 10 , 12 , 16 ; 4 ) chianti ([130,131,132])
ch051019351 ( 7 ; 4 ) chianti ([133])
ch051020450 ( 14 , 18 , 26 ; 4 ) chianti, chianti ([131])
ch061021351 ( 8 ; 4 ) chianti ([132,134])
ch061022450 ( 20 , 28 ; 4 ) chianti ([130,132])
ch071023351 ( 10 , 12 , 14 , 16 , 18 , 20 , 26 , 28 ; 4 ) chianti ([131,132])
ch071024450 ( 6 ; 5 ) chianti ([135,136,137]), nist
ch081025351 ( 6 ; 5 ) chianti ([138])
ch081026450 ( 7 ; 5 ) chianti ([139,140])
ch091027450 ( 8 , 10 ; 5 ) chianti ([135])
ch091028351 ( 7 ; 5 ) chianti ([141,142])
ch101029450 ( 16 , 18 , 20 ; 5 ) chianti ([135,143])
ch101030351 ( 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 5 ) chianti ([144])
ch111031351 ( 28 ; 5 ) chianti ([145])
ch111032450 ( 26 ; 5 ) chianti ([145])
ch121033450 ( 28 ; 5 ) chianti ([135,143,145])
ch121034351 ( 7 ; 6 ) chianti ([141,142])
ch131035450 ( 7 ; 6 ) chianti, chianti ([146])
ch131036351 ( 8 ; 6 ) chianti ([147,148,149,150])
ch141037450 ( 8 ; 6 ) chianti ([149])
ch141038351 ( 10 ; 6 ) chianti ([150,151,152])
ch151039450 ( 10 , 12 , 14 ; 6 ) chianti ([152,153])
ch151040351 ( 12 ; 6 ) chianti ([154])
ch161041351 ( 14 ; 6 ) chianti ([152,153]
ch161042450 ( 16 ; 6 ) chianti ([155])
ch171043450 ( 18 ; 6 ) chianti ([156]
ch171044351 ( 16 ; 6 ) chianti ([155,157])
ch181045351 ( 18 ; 6 ) chianti ([156])
ch181046450 ( 20 ; 6 ) chianti ([158])
ch191047351 ( 20 ; 6 ) chianti ([159,160])
ch191048450 ( 12 , 14 , 16 , 18 , 20 ; 7 ) chianti ([161])
ch201049450 ( 26 ; 7 ) chianti ([162])
ch201050351 ( 26 ; 6 ) chianti ([163,164])
ch211051450 ( 14 , 16 , 18 ; 8 ) chianti ([165])
ch211052351 ( 8 ; 7 ) chianti ([166])
ch221053351 ( 12 , 14 , 16 , 18 , 20 ; 7 ) chianti ([161])
ch221054450 ( 26 ; 8 ) chianti ([167])
ch231055351 ( 26 ; 7 ) chianti ([161,162])
ch241056450 ( 12 , 14 , 16 , 20 ; 9 ) chianti ([168]), nist
ch241057351 ( 10 , 12 ; 8 ) chianti ([169])
ch251058351 ( 14 , 16 , 18 ; 8 ) chianti ([165])
ch251059450 ( 26 , 28 ; 9 ) chianti ([168,170,171]), nist
ch261060450 ( 14 ; 10 ) chianti ([172])
ch261061351 ( 26 ; 8 ) chianti ([167])
ch271062450 ( 16 ; 10 ) chianti ([173])
ch271063351 ( 10 ; 9 ) chianti ([174])
ch281064351 ( 12 ; 9 ) chianti ([174,175])
ch281065450 ( 18 , 20 , 28 ; 10 ) chianti ([173,176])
ch291066450 ( 26 ; 10 ) chianti ([177])
ch291067351 ( 14 ; 9 ) chianti ([174,178])
ch301068351 ( 16 ; 9 ) chianti ([174,179])
ch301069450 ( 12 , 14 , 16 , 18 , 20 , 26 , 28 ; 11 ) chianti ([125,180])
ch311070450 ( 14 ; 12 ) chianti, chianti ([181])
ch311071351 ( 20 , 26 ; 9 ) chianti ([174,182])
ch321072351 ( 14 ; 10 ) chianti ([172])
ch321073450 ( 16 , 18 , 20 , 28 ; 12 ) chianti ([183]), nist
ch331074450 ( 14 ; 13 ) chianti, chianti ([184])
ch331075351 ( 16 ; 10 ) chianti ([185])
ch341076450 ( 16 ; 13 ) chianti ([186,187])
ch341077351 ( 18 , 20 , 28 ; 10 ) chianti ([176])
ch351078351 ( 26 ; 10 ) chianti ([177])
ch351079450 ( 26 ; 13 ) chianti ([188])
ch361080450 ( 26 ; 19 ) chianti ([189,190]), nist
ch361081351 ( 12 , 14 , 16 , 18 , 20 , 28 ; 11 ) chianti ([180])
ch371082351 ( 14 ; 12 ) chianti ([191])
ch371083450 ( 26 ; 20 ) chianti ([192])
ch381084351 ( 16 , 18 , 20 , 28 ; 12 ) chianti ([183])
ch391085351 ( 14 ; 13 ) chianti ([193])
ch401086351 ( 16 ; 13 ) chianti ([186,187,194])
ch411087351 ( 26 ; 18 ) chianti ([195,196])
ch421088351 ( 26 ; 19 ) chianti ([189,197])
ch431089351 ( 26 ; 14 ) chianti ([196,198])
cj941090356 ( 1 ; 1 ) [103,199]
cp991091450 ( 26 ; 21 ) [200]
cp991092356 ( 26 ; 21 ) [200]
eg991093356 ( 26 ; 11 ) [180,201]
ei991094356 ( 26 ; 12 ) chianti, [202]
fe011095450 ( 26 ; 25 ) [203,204,205,206]
fe101096356 ( 26 ; 17 ) chianti, [207]
fe241097356 ( 26 ; 3 ) [127,208]
gr581098450 ( 26 ; 23 ) [209]
im1099363 ( 7 , 12 , 16 , 18 , 28 ; 1 ) , ( 7 ; 2 ) Impact parameter method
kh841100450 ( 2 ; 2 ) [210]
km871101356 ( 10 , 14 , 16 ; 2 ) [105,106,107,211]
kn891102369 ( 6 ; 2 ) [104]
kn89b1103369 ( 2 , 8 , 12 , 20 ; 2 ) [104,107,211]
kn89c1104369 ( 26 ; 2 ) [104,105,106,107,211]
mg051105450 ( 12 ; 8 ) PP95 ([212]), [213,214]
ne011106450 ( 10 ; 7 ) chianti, PP95 ([212,215])
ne021107450 ( 10 ; 8 ) chianti, PP95 ([212,214])
np961108450 ( 26 ; 24 ) [216]
ny011109450 ( 7 ; 7 ) [217,218,219]
ox011110450 ( 8 ; 6 ) PP95 ([212,220]), chianti ([149])
ox021111450 ( 8 ; 7 ) chianti, PP95 ([212,215])
ox031112450 ( 8 ; 8 ) [218,219], PP95 ([212,214]
pb951113356 ( 18 , 20 ; 17 ) [207]
pp011114450 ( 6 ; 6 ) PP95 ([136,212,213])
pp011115356 ( 6 ; 6 ) PP95 ([221,222,223])
pp021116356 ( 7 ; 7 ) PP95 ([224,225])
pp031117356 ( 8 ; 8 ) PP95 ([225,226])
pp041118356 ( 16 ; 14 ) PP95 ([214])
pp051119450 ( 18 , 20 ; 16 ) PP95 ([212,214])
pp051120356 ( 18 ; 14 ) , ( 20 ; 19 ) PP95 ([214])
pp061121356 ( 18 ; 15 ) PP95 ([227])
pp061122450 ( 10 ; 9 ) , ( 20 ; 17 ) PP95 ([212])
pp071123356 ( 18 , 20 ; 16 ) PP95 ([228])
pp071124450 ( 16 ; 14 ) PP95 ([212,229,230])
pp081125450 ( 18 ; 15 ) PP95 ([212,231])
pp091126450 ( 18 ; 17 ) PP95 ([214])
pp101127450 ( 20 ; 19 ) PP95 ([232])
pp111128450 ( 18 ; 14 ) PP95 ([187,212])
rb961129356 ( 16 ; 15 ) [233], chianti ([234])
sm001130351 ( 26 ; 13 ) [235]
ss011131450 ( 16 ; 15 ) chianti
ss961132356 ( 18 , 20 ; 13 ) [236]
tb931133450 ( 6 , 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 2 ) [60]
zh961134356 ( 26 ; 24 ) [237]
zp971135356 ( 26 ; 23 ) [238]
zs871136356 ( 18 , 28 ; 2 ) [106]
bm011137356 ( 26 ; 3 ) [5]
bm011138450 ( 26 ; 3 ) [5]
bm021139356 ( 26 ; 4 10 ) [8]
pm011140450 ( 26 ; 2 9 ) [6]
pm021141450 ( 26 ; 18 25 ) [7]
mk011142450 ( 26 ; 10 17 ) [9]
gm011143450 ( 8 ; 2 8 ) [11]
gk011144450 ( 7 ; 1 7 ) [17]
pm031145450 ( 16 ; 2 16 ) [14]
pq011146450 ( 10 , 12 , 14 , 16 , 18 , 20 ; 2 Z ) [15]
pq021147450 ( 28 ; 2 27 ) [16]
pq031148450 ( 14 ; 2 13 ) [22]
pq041149450 ( 9 , 11 , 15 , 17 , 19 ; 2 ( Z 1 ) ) [26]
( 21 25 , 27 , 29 , 30 ; 2 ( Z 1 ) )
mb011150356 ( 9 , 11 , 15 , 17 , 19 , 21 25 , 27 ; 3 ( Z 2 ) ) ,Unpublished
( 28 ; 19 26 ) , ( 29 , 30 ; 3 26 )
fj011151356 ( 26 ; 2 , 3 ) [110]
fj011152450 ( 26 ; 2 , 3 ) [110]
gu011153450 ( 26 ; 11 21 ) [109]
gu011154136 ( 26 ; 11 21 ) [109]
kf011155525 ( 3 5 , 21 , 29 , 30 ; 2 ) ,[98]
( 6 8 , 16 ; ( Z 3 ) Z ) ,
( 10 , 12 , 14 , 16 , 18 ; ( Z 2 ) Z ) ,
( 26 , 28 ; ( Z 2 ) Z ) ,
( 20 ; ( Z 1 ) Z )
bb0111565, 1595 ( 1 30 ; 1 Z ) [111]
tb011157753 ( 1 30 ; 1 ) , ( 21 25 , 27 30 ; 2 ) ,Hydrogenic
( 28 ; 3 10 , 21 26 , 28 )
tb021158753 ( 2 20 , 26 ; 2 ) [239]
tb031159753 ( 6 , 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 3 ) [240]
tb041160753 ( 6 , 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 4 ) [241]
tb051161753 ( 6 , 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 5 ) [242]
tb061162753 ( 6 , 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 6 ) [243]
tb071163753 ( 7 , 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 7 ) [60]
tb081164753 ( 8 , 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 8 ) [60]
tb091165753 ( 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 9 ) [60]
tb101166753 ( 10 , 12 , 14 , 16 , 18 , 20 , 26 ; 10 ) [244]
tb111167753 ( 14 , 18 , 20 , 26 ; 11 ) [60]
tb121168753 ( 26 ; 12 ) [245]
tb131169753 ( 26 ; 13 ) [246]
tb141170753 ( 14 , 26 ; 14 ) [60,247]
tb151171753 ( 16 ; 16 ) , ( 18 ; 18 ) , ( 26 ; 15 21 ) [60]
tb161172753 ( 26 ; 22 ) [248]
tb171173753 ( 26 ; 23 ) [249]
tb181174753 ( 26 ; 24 ) [250]
tb191175753 ( 26 ; 25 ) [251]
tb201176753 ( 26 ; 26 ) [252]
tb211177753 ( 28 ; 27 ) [218]
vy0111781, 759 ( 3 , 4 , 5 , 9 , 11 , 15 , 17 , 19 , 21 25 ; 1 Z ) ,[102]
( 27 , 29 , 30 ; 1 Z ) , ( 13 ; 1 , 2 , 12 , 13 ) ,
( 14 ; 14 ) , ( 16 ; 16 ) , ( 18 ; 18 ) ,
( 20 ; 18 , 19 , 20 ) , ( 26 ; 19 25 ) , ( 28 ; 21 28 )
kp011179753 ( 26 ; 3 26 ) [3,5,8,10]
gg011180753 ( 6 ; 3 6 ) [21,37]
gm011181753 ( 8 ; 3 8 ) [11]
gg021182749 ( 8 ; 6 8 ) [23,32,33]
gk011183753 ( 7 ; 3 7 ) [17]
wk011184749 ( 10 , 12 , 14 , 16 , 18 , 20 ; 3 10 ) [18]
gg031185749 ( 10 ; 8 10 ) [13,35,253]
wk021186749 ( 12 , 14 , 16 , 18 , 20 ; 11 17 ) [19]
gg041187749 ( 12 ; 3 12 ) [25]
wk031188749 ( 13 ; 3 11 ) [24]
wk041189749 ( 28 ; 3 20 ) [20]
tr011190749 ( 9 , 11 ; 3 ( Z 2 ) ) , ( 15 , 17 , 19 ; 3 11 ) [27]
( 21 25 , 27 , 29 , 30 ; 3 11 )
tr021191749 ( 15 , 17 , 19 ; 12 ( Z 2 ) ) ,[28]
( 21 25 , 27 , 29 , 30 ; 12 18 )
tr031192749 ( 21 25 , 27 ; 19 ( Z 2 ) ) ,[29]
( 28 ; 21 26 ) , ( 29 , 30 ; 19 26 )
ap01119381 ( 3 30 ; 2 ) , ( 15 30 ; 14 Z ) [89,97]
ap01119487 ( 3 5 ; 2 Z ) , ( 7 , 21 , 29 , 30 ; 2 ) ,[89,97]
( 13 25 , 27 30 ; 13 Z )
ad011195838 ( 3 12 ; 3 Z ) , ( 13 30 ; 3 13 ) AMDPP ([65])
ad011196839 ( 3 13 ; 3 Z ) , ( 14 30 ; 3 12 ) ,AMDPP ([64])
( 26 , 28 ; 13 20 ) , ( 26 ; 21 26 )
ns011197822 ( 13 ; 13 ) , ( 14 ; 13 , 14 ) [99]
gt011198830 ( 1 30 ; 1 ) [100]

References

  1. Kallman, T.; Bautista, M. Photoionization and High-Density Gas. Astrophys. J. Suppl. Ser. 2001, 133, 221–253. [Google Scholar] [CrossRef]
  2. Bautista, M.A.; Kallman, T.R. The XSTAR Atomic Database. Astrophys. J. Suppl. Ser. 2001, 134, 139–149. [Google Scholar] [CrossRef]
  3. Palmeri, P.; Mendoza, C.; Kallman, T.R.; Bautista, M.A. On the Structure of the Iron K Edge. Astrophys. J. 2002, 577, L119–L122. [Google Scholar] [CrossRef]
  4. Palmeri, P.; Mendoza, C.; Kallman, T.; Bautista, M. Photoionization of the Fe Ions: Structure of the K-Edge. In X-ray and Inner-Shell Processes; Bianconi, A., Marcelli, A., Saini, N.L., Eds.; American Institute of Physics: College Park, MD, USA, 2003; AIP Conf. Proc.; Volume 652, pp. 159–164. [Google Scholar] [CrossRef]
  5. Bautista, M.A.; Mendoza, C.; Kallman, T.R.; Palmeri, P. Atomic data for the K-vacancy states of Fe XXIV. Astron. Astrophys. 2003, 403, 339–355. [Google Scholar] [CrossRef] [Green Version]
  6. Palmeri, P.; Mendoza, C.; Kallman, T.R.; Bautista, M.A. A complete set of radiative and Auger rates for K-vacancy states in Fe XVIII-Fe XXV. Astron. Astrophys. 2003, 403, 1175–1184. [Google Scholar] [CrossRef]
  7. Palmeri, P.; Mendoza, C.; Kallman, T.R.; Bautista, M.A.; Meléndez, M. Modeling of iron K lines: Radiative and Auger decay data for Fe II-Fe IX. Astron. Astrophys. 2003, 410, 359–364. [Google Scholar] [CrossRef] [Green Version]
  8. Bautista, M.A.; Mendoza, C.; Kallman, T.R.; Palmeri, P. K-shell photoionization and electron impact excitation of Fe XVII-Fe XXIII. Astrophys. J. Suppl. Ser. 2004, 418, 1171–1178. [Google Scholar] [CrossRef] [Green Version]
  9. Mendoza, C.; Kallman, T.R.; Bautista, M.A.; Palmeri, P. Decay properties of K-vacancy states in Fe X-Fe XVII. Astron. Astrophys. 2004, 414, 377–388. [Google Scholar] [CrossRef]
  10. Kallman, T.R.; Palmeri, P.; Bautista, M.A.; Mendoza, C.; Krolik, J.H. Photoionization Modeling and the K Lines of Iron. Astrophys. J. Suppl. Ser. 2004, 155, 675–701. [Google Scholar] [CrossRef] [Green Version]
  11. García, J.; Mendoza, C.; Bautista, M.A.; Gorczyca, T.W.; Kallman, T.R.; Palmeri, P. K-Shell Photoabsorption of Oxygen Ions. Astrophys. J. Suppl. Ser. 2005, 158, 68–79. [Google Scholar] [CrossRef] [Green Version]
  12. Palmeri, P.; Kallman, T.R.; Mendoza, C.; Bautista, M.A.; Krolik, J.H. Photoionization Modeling: the K Lines and Edges of Iron. In X-ray Diagnostics of Astrophysical Plasmas: Theory, Experiment, and Observation; Smith, R., Ed.; American Institute of Physics: College Park, MD, USA, 2005; AIP Conf. Proc.; Volume 774, pp. 75–82. [Google Scholar] [CrossRef]
  13. Juett, A.M.; Schulz, N.S.; Chakrabarty, D.; Gorczyca, T.W. High-Resolution X-ray Spectroscopy of the Interstellar Medium. II. Neon and Iron Absorption Edges. Astrophys. J. 2006, 648, 1066–1078. [Google Scholar] [CrossRef] [Green Version]
  14. Palmeri, P.; Mendoza, C.; Bautista, M.A.; García, J.; Kallman, T.R. K lines in the sulphur isonuclear sequence. Rad. Phys. Chem. 2006, 75, 1465–1470. [Google Scholar] [CrossRef]
  15. Palmeri, P.; Quinet, P.; Mendoza, C.; Bautista, M.A.; García, J.; Kallman, T.R. Radiative and Auger Decay of K-Vacancy Levels in the Ne, Mg, Si, S, Ar, and Ca Isonuclear Sequences. Astrophys. J. Suppl. Ser. 2008, 177, 408–416. [Google Scholar] [CrossRef] [Green Version]
  16. Palmeri, P.; Quinet, P.; Mendoza, C.; Bautista, M.A.; García, J.; Witthoeft, M.C.; Kallman, T.R. Radiative and Auger Decay Data for Modeling Nickel K Lines. Astrophys. J. Suppl. Ser. 2008, 179, 542–552. [Google Scholar] [CrossRef] [Green Version]
  17. García, J.; Kallman, T.R.; Witthoeft, M.; Behar, E.; Mendoza, C.; Palmeri, P.; Quinet, P.; Bautista, M.A.; Klapisch, M. Nitrogen K-Shell Photoabsorption. Astrophys. J. Suppl. Ser. 2009, 185, 477–485. [Google Scholar] [CrossRef] [Green Version]
  18. Witthoeft, M.C.; Bautista, M.A.; Mendoza, C.; Kallman, T.R.; Palmeri, P.; Quinet, P. K-Shell Photoionization and Photoabsorption of Ne, Mg, Si, S, Ar, and Ca. Astrophys. J. Suppl. Ser. 2009, 182, 127–130. [Google Scholar] [CrossRef] [Green Version]
  19. Witthoeft, M.C.; García, J.; Kallman, T.R.; Bautista, M.A.; Mendoza, C.; Palmeri, P.; Quinet, P. K-shell Photoionization of Na-like to Cl-like Ions of Mg, Si, S, Ar, and Ca. Astrophys. J. Suppl. Ser. 2011, 192, 7. [Google Scholar] [CrossRef] [Green Version]
  20. Witthoeft, M.C.; Bautista, M.A.; García, J.; Kallman, T.R.; Mendoza, C.; Palmeri, P.; Quinet, P. K-shell Photoionization of Nickel Ions using R-matrix. Astrophys. J. Suppl. Ser. 2011, 196, 7. [Google Scholar] [CrossRef] [Green Version]
  21. Hasoglu, M.F.; Abdel-Naby, S.A.; Gorczyca, T.W.; Drake, J.J.; McLaughlin, B.M. K-shell Photoabsorption Studies of the Carbon Isonuclear Sequence. Astrophys. J. 2010, 724, 1296–1304. [Google Scholar] [CrossRef]
  22. Palmeri, P.; Quinet, P.; Mendoza, C.; Bautista, M.A.; García, J.; Witthoeft, M.C.; Kallman, T.R. Atomic decay data for modeling the Al K lines. Astron. Astrophys. 2011, 525, A59. [Google Scholar] [CrossRef] [Green Version]
  23. Gorczyca, T.W.; Bautista, M.A.; Hasoglu, M.F.; García, J.; Gatuzz, E.; Kaastra, J.S.; Kallman, T.R.; Manson, S.T.; Mendoza, C.; Raassen, A.J.J.; et al. A Comprehensive X-ray Absorption Model for Atomic Oxygen. Astrophys. J. 2013, 779, 78. [Google Scholar] [CrossRef] [Green Version]
  24. Witthoeft, M.C.; Bautista, M.A.; García, J.; Kallman, T.R.; Mendoza, C.; Palmeri, P.; Quinet, P. Photoionization and photoabsorption cross sections for the aluminum iso-nuclear sequence. At. Data Nucl. Data Tables 2013, 99, 53–61. [Google Scholar] [CrossRef]
  25. Hasoğlu, M.F.; Abdel-Naby, S.A.; Gatuzz, E.; García, J.; Kallman, T.R.; Mendoza, C.; Gorczyca, T.W. K-shell Photoabsorption of Magnesium Ions. Astrophys. J. Suppl. Ser. 2014, 214, 8. [Google Scholar] [CrossRef] [Green Version]
  26. Palmeri, P.; Quinet, P.; Mendoza, C.; Bautista, M.A.; García, J.; Witthoeft, M.C.; Kallman, T.R. Atomic decay data for modeling K lines of iron peak and light odd-Z elements. Astron. Astrophys. 2012, 543, A44. [Google Scholar] [CrossRef] [Green Version]
  27. Palmeri, P.; Quinet, P.; Mendoza, C.; Bautista, M.A.; Witthoeft, M.C.; Kallman, T.R. K-shell photoabsorption and photoionisation of trace elements. I. Isoelectronic sequences with electron number 3 ≤ N ≤ 11. Astron. Astrophys. 2016, 589, A137. [Google Scholar] [CrossRef] [Green Version]
  28. Mendoza, C.; Bautista, M.A.; Palmeri, P.; Quinet, P.; Witthoeft, M.C.; Kallman, T.R. K-shell photoabsorption and photoionization of trace elements. II. Isoelectronic sequences with electron number 12 ≤ N ≤ 18. Astron. Astrophys. 2017, 604, A63. [Google Scholar] [CrossRef] [Green Version]
  29. Mendoza, C.; Bautista, M.A.; Palmeri, P.; Quinet, P.; Witthoeft, M.C.; Kallman, T.R. K-shell photoabsorption and photoionization of trace elements. III. Isoelectronic sequences with electron number 19 ≤ N ≤ 26. Astron. Astrophys. 2018, 616, A62. [Google Scholar] [CrossRef]
  30. Kallman, T.R.; Bautista, M.A.; Goriely, S.; Mendoza, C.; Miller, J.M.; Palmeri, P.; Quinet, P.; Raymond, J. Spectrum Synthesis Modeling of the X-ray Spectrum of GRO J1655-40 Taken During the 2005 Outburst. Astrophys. J. 2009, 701, 865–884. [Google Scholar] [CrossRef]
  31. García, J.; Ramírez, J.M.; Kallman, T.R.; Witthoeft, M.; Bautista, M.A.; Mendoza, C.; Palmeri, P.; Quinet, P. Modeling the Oxygen K Absorption in the Interstellar Medium: An XMM-Newton View of Sco X-1. Astrophys. J. 2011, 731, L15. [Google Scholar] [CrossRef] [Green Version]
  32. Gatuzz, E.; García, J.; Mendoza, C.; Kallman, T.R.; Witthoeft, M.; Lohfink, A.; Bautista, M.A.; Palmeri, P.; Quinet, P. Photoionization Modeling of Oxygen K Absorption in the Interstellar Medium: The Chandra Grating Spectra of XTE J1817-330. Astrophys. J. 2013, 768, 60. [Google Scholar] [CrossRef] [Green Version]
  33. Gatuzz, E.; García, J.; Mendoza, C.; Kallman, T.R.; Witthoeft, M.; Lohfink, A.; Bautista, M.A.; Palmeri, P.; Quinet, P. Erratum: “Photoionization Modeling of Oxygen K Absorption in the Interstellar Medium, the Chandra Grating Spectra of XTE J1817-330” (2013, ApJ, 768, 60). Astrophys. J. 2013, 778, 83. [Google Scholar] [CrossRef] [Green Version]
  34. Gatuzz, E.; García, J.; Mendoza, C.; Kallman, T.R.; Bautista, M.A.; Gorczyca, T.W. Physical Properties of the Interstellar Medium Using High-resolution Chandra Spectra: O K-edge Absorption. Astrophys J. 2014, 790, 131. [Google Scholar] [CrossRef] [Green Version]
  35. Gatuzz, E.; García, J.; Kallman, T.R.; Mendoza, C.; Gorczyca, T.W. ISMabs: A Comprehensive X-ray Absorption Model for the Interstellar Medium. Astrophys. J. 2015, 800, 29. [Google Scholar] [CrossRef]
  36. Gatuzz, E.; García, J.A.; Kallman, T.R.; Mendoza, C. Oxygen, neon, and iron X-ray absorption in the local interstellar medium. Astron. Astrophys. 2016, 588, A111. [Google Scholar] [CrossRef] [Green Version]
  37. Gatuzz, E.; Ness, J.U.; Gorczyca, T.W.; Hasoglu, M.F.; Kallman, T.R.; García, J.A. Carbon X-ray absorption in the local ISM: Fingerprints in X-ray Novae spectra. Mon. Not. R. Astron. Soc. 2018, 479, 2457–2463. [Google Scholar] [CrossRef]
  38. Gatuzz, E.; García, J.A.; Kallman, T.R. On the discrepancy between the X-ray and UV absorption measurements of O VI in the local ISM. Mon. Not. R. Astron. Soc. 2019, 483, L75–L79. [Google Scholar] [CrossRef]
  39. Gatuzz, E.; Gorczyca, T.W.; Hasoglu, M.F.; Schulz, N.S.; Corrales, L.; Mendoza, C. Silicon ISM X-ray absorption: The gaseous component. Mon. Not. R. Astron. Soc. 2020, 498, L20–L24. [Google Scholar] [CrossRef]
  40. Deprince, J.; Palmeri, P.; Quinet, P.; Fritzsche, S.; Bautista, M.; Mendoza, C.; Kallman, T.R.; García, J.A. Plasma-Environment Effects on the Atomic Structure and K Lines of He- and Li-like Oxygen Ions. In Workshop on Astrophysical Opacities; Astronomical Society of the Pacific: San Francisco, CA, USA, 2018; AIP Conf. Ser.; Volume 515, p. 289. [Google Scholar]
  41. Deprince, J.; Bautista, M.A.; Fritzsche, S.; García, J.A.; Kallman, T.R.; Mendoza, C.; Palmeri, P.; Quinet, P. Plasma environment effects on K lines of astrophysical interest. I. Atomic structure, radiative rates, and Auger widths of oxygen ions. Astron. Astrophys. 2019, 624, A74. [Google Scholar] [CrossRef] [Green Version]
  42. Deprince, J.; Bautista, M.A.; Fritzsche, S.; García, J.A.; Kallman, T.; Mendoza, C.; Palmeri, P.; Quinet, P. Plasma environment effects on K lines of astrophysical interest. II. Ionization potentials, K thresholds, radiative rates, and Auger widths in Ne- through He-like iron ions (Fe XVII–Fe XXV). Astron. Astrophys. 2019, 626, A83. [Google Scholar] [CrossRef] [Green Version]
  43. Deprince, J.; Bautista, M.A.; Fritzsche, S.; García, J.A.; Kallman, T.R.; Mendoza, C.; Palmeri, P.; Quinet, P. Plasma-environment effects on K lines of astrophysical interest. III. IPs, K thresholds, radiative rates, and Auger widths in Fe IX–Fe XVI. Astron. Astrophys. 2020, 635, A70. [Google Scholar] [CrossRef]
  44. Deprince, J.; Bautista, M.; Fritzsche, S.; García, J.; Kallman, T.; Mendoza, C.; Palmeri, P.; Quinet, P. K-line X-ray fluorescence from highly charged iron ions under dense astrophysical plasma conditions. X-ray Spectrom. 2020, 49, 29–32. [Google Scholar] [CrossRef]
  45. Deprince, J.; Bautista, M.A.; Fritzsche, S.; García, J.A.; Kallman, T.R.; Mendoza, C.; Palmeri, P.; Quinet, P. Plasma environment effects on K lines of astrophysical interest. IV. IPs, K thresholds, radiative rates, and Auger widths in Fe II-Fe VIII. Astron. Astrophys. 2020, 643, A57. [Google Scholar] [CrossRef]
  46. Kallman, T.; Bautista, M.; Deprince, J.; Garcia, J.A.; Mendoza, C.; Ogorzalek, A.; Palmeri, P.; Quinet, P. Photoionization Models for High Density Gas. arXiv 2020, arXiv:2011.10603. [Google Scholar]
  47. Hitomi Collaboration; Aharonian, F.; Akamatsu, H.; Akimoto, F.; Allen, S.W.; Anabuki, N.; Angelini, L.; Arnaud, K.; Audard, M.; Awaki, H.; et al. The quiescent intracluster medium in the core of the Perseus cluster. Nature 2016, 535, 117–121. [Google Scholar] [CrossRef] [Green Version]
  48. Nikolić, D.; Gorczyca, T.W.; Korista, K.T.; Ferland, G.J.; Badnell, N.R. Suppression of Dielectronic Recombination due to Finite Density Effects. Astrophys. J. 2013, 768, 82. [Google Scholar] [CrossRef] [Green Version]
  49. Nikolić, D.; Gorczyca, T.W.; Korista, K.T.; Chatzikos, M.; Ferland, G.J.; Guzmán, F.; van Hoof, P.A.M.; Williams, R.J.R.; Badnell, N.R. Suppression of Dielectronic Recombination Due to Finite Density Effects. II. Analytical Refinement and Application to Density-dependent Ionization Balances and AGN Broad-line Emission. Astrophys. J. Suppl. 2018, 237, 41. [Google Scholar] [CrossRef] [Green Version]
  50. García, J.A.; Fabian, A.C.; Kallman, T.R.; Dauser, T.; Parker, M.L.; McClintock, J.E.; Steiner, J.F.; Wilms, J. The effects of high density on the X-ray spectrum reflected from accretion discs around black holes. Mon. Not. R. Astron. Soc. 2016, 462, 751–760. [Google Scholar] [CrossRef] [Green Version]
  51. García, J.A.; Kallman, T.R.; Bautista, M.; Mendoza, C.; Deprince, J.; Palmeri, P.; Quinet, P. The Problem of the High Iron Abundance in Accretion Disks around Black Holes. In Workshop on Astrophysical Opacities; Astronomical Society of the Pacific: San Francisco, CA, USA, 2018; AIP Conf. Ser.; Volume 515, p. 282. [Google Scholar]
  52. Leutenegger, M.A.; Kühn, S.; Micke, P.; Steinbrügge, R.; Stierhof, J.; Shah, C.; Hell, N.; Bissinger, M.; Hirsch, M.; Ballhausen, R.; et al. High-Precision Determination of Oxygen Kα Transition Energy Excludes Incongruent Motion of Interstellar Oxygen. Phys. Rev. Lett. 2020, 125, 243001. [Google Scholar] [CrossRef] [PubMed]
  53. Greisen, E.W. FITS: A Remarkable Achievement in Information Exchange. In Information Handling in Astronomy—Historical Vistas; Heck, A., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2003; Astrophysics and Space Science Library; Volume 285, p. 71. [Google Scholar] [CrossRef] [Green Version]
  54. Pence, W.D.; Chiappetti, L.; Page, C.G.; Shaw, R.A.; Stobie, E. Definition of the Flexible Image Transport System (FITS), version 3.0. Astron. Astrophys. 2010, 524, A42. [Google Scholar] [CrossRef] [Green Version]
  55. Cowan, R.D. The Theory of Atomic Structure and Spectra; University of California Press: Berkeley, CA, USA, 1981. [Google Scholar]
  56. Berrington, K.A.; Eissner, W.B.; Norrington, P.H. RMATRX1: Belfast atomic R-matrix codes. Comput. Phys. Commun. 1995, 92, 290–420. [Google Scholar] [CrossRef]
  57. Burke, P.G. R-Matrix Theory of Atomic Collisions; Springer: New York, NY, USA, 2011. [Google Scholar]
  58. Gorczyca, T.W.; Robicheaux, F. Auger decay of the photoexcited 2p-1nl Rydberg series in argon. Phys. Rev. A 1999, 60, 1216–1225. [Google Scholar] [CrossRef]
  59. Kramida, A.; Ralchenko, Y.; Reader, J.; NIST ASD Team. NIST Atomic Spectra Database; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2020. Available online: https://physics.nist.gov/asd (accessed on 15 November 2020).
  60. Cunto, W.; Mendoza, C.; Ochsenbein, F.; Zeippen, C.J. TOPbase at the CDS. Astron. Astrophys. 1993, 275, L5–L8. [Google Scholar]
  61. Dere, K.P.; Landi, E.; Mason, H.E.; Monsignori Fossi, B.C.; Young, P.R. CHIANTI—An atomic database for emission lines. Astron. Astrophys. Suppl. Ser. 1997, 125, 149–173. [Google Scholar] [CrossRef] [Green Version]
  62. Hummer, D.G.; Berrington, K.A.; Eissner, W.; Pradhan, A.K.; Saraph, H.E.; Tully, J.A. Atomic data from the IRON project. I. Goals and methods. Astron. Astrophys. 1993, 279, 298–309. [Google Scholar]
  63. Pradhan, A.K.; Peng, J. Atomic Data For The Analysis of Emission Lines. In The Analysis of Emission Lines; Williams, R.E.W., Livio, M., Eds.; Space Institute Symposium Series; Cambridge University Press: Cambridge, UK, 1995; Volume 8, pp. 8–34. [Google Scholar]
  64. Badnell, N.R.; O’Mullane, M.G.; Summers, H.P.; Altun, Z.; Bautista, M.A.; Colgan, J.; Gorczyca, T.W.; Mitnik, D.M.; Pindzola, M.S.; Zatsarinny, O. Dielectronic recombination data for dynamic finite-density plasmas. I. Goals and methodology. Astron. Astrophys. 2003, 406, 1151–1165. [Google Scholar] [CrossRef]
  65. Badnell, N.R. Radiative Recombination Data for Modeling Dynamic Finite-Density Plasmas. Astrophys. J. Suppl. Ser. 2006, 167, 334–342. [Google Scholar] [CrossRef] [Green Version]
  66. Eissner, W.; Jones, M.; Nussbaumer, H. Techniques for the calculation of atomic structures and radiative data including relativistic corrections. Comput. Phys. Commun. 1974, 8, 270–306. [Google Scholar] [CrossRef]
  67. Badnell, N.R. A Breit-Pauli distorted wave implementation for AUTOSTRUCTURE. Comput. Phys. Commun. 2011, 182, 1528–1535. [Google Scholar] [CrossRef]
  68. Parpia, F.A.; Fischer, C.F.; Grant, I.P. GRASP92: A package for large-scale relativistic atomic structure calculations. Comput. Phys. Commun. 1996, 94, 249–271. [Google Scholar] [CrossRef]
  69. Hickman, A.P. Complex potential model for dielectronic recombination. J. Phys. B At. Mol. Phys. 1984, 17, L101–L106. [Google Scholar] [CrossRef]
  70. Robicheaux, F.; Gorczyca, T.W.; Pindzola, M.S.; Badnell, N.R. Inclusion of radiation damping in the close-coupling equations for electron-atom scattering. Phys. Rev. A 1995, 52, 1319–1333. [Google Scholar] [CrossRef] [PubMed]
  71. Davies, P.C.W.; Seaton, M.J. Radiation damping in the optical continuum. J. Phys. B At. Mol. Phys. 1969, 2, 757–765. [Google Scholar] [CrossRef]
  72. Bell, R.H.; Seaton, M.J. Dielectronic recombination. I. General theory. J. Phys. B At. Mol. Phys. 1985, 18, 1589–1629. [Google Scholar] [CrossRef]
  73. Froese Fischer, C. The Hartree-Fock Method for Atom; Wiley: New York, NY, USA, 1977. [Google Scholar]
  74. Froese Fischer, C. A general multi-configuration Hartree-Fock program. Comput. Phys. Commun. 1991, 64, 431–454. [Google Scholar] [CrossRef]
  75. Froese Fischer, C.; Brage, T.; Jönsson, P. Computational Atomic Structure: An MCHF Approach; Institute of Physics Publ.: Bristol, UK, 1997. [Google Scholar]
  76. Stolte, W.C.; Lu, Y.; Samson, J.A.R.; Hemmers, O.; Hansen, D.L.; Whitfield, S.B.; Wang, H.; Glans, P.; Lindle, D.W. The K-shell Auger decay of atomic oxygen. J. Phys. B At. Mol. Phys. 1997, 30, 4489–4497. [Google Scholar] [CrossRef]
  77. Liao, J.Y.; Zhang, S.N.; Yao, Y. Wavelength Measurements of K Transitions of Oxygen, Neon, and Magnesium with X-ray Absorption Lines. Astrophys. J. 2013, 774, 116. [Google Scholar] [CrossRef] [Green Version]
  78. Juett, A.M.; Schulz, N.S.; Chakrabarty, D. High-Resolution X-ray Spectroscopy of the Interstellar Medium: Structure at the Oxygen Absorption Edge. Astrophys. J. 2004, 612, 308–318. [Google Scholar] [CrossRef] [Green Version]
  79. McLaughlin, B.M.; Ballance, C.P.; Bowen, K.P.; Gardenghi, D.J.; Stolte, W.C. High Precision K-shell Photoabsorption Cross Sections for Atomic Oxygen: Experiment and Theory. Astrophys. J. 2013, 771, L8. [Google Scholar] [CrossRef] [Green Version]
  80. McLaughlin, B.M.; Ballance, C.P.; Bowen, K.P.; Gardenghi, D.J.; Stolte, W.C. Erratum: “High Precision K-Shell Photoabsorption Cross Sections for Atomic Oxygen: Experiment and Theory” (2013, ApJL, 771, l8). Astrophys. J. 2013, 779, L31. [Google Scholar] [CrossRef] [Green Version]
  81. Menzel, A.; Benzaid, S.; Krause, M.O.; Caldwell, C.D.; Hergenhahn, U.; Bissen, M. Natural widths in open-shell atoms: The K absorption spectrum of atomic oxygen. Phys. Rev. A 1996, 54, R991–R994. [Google Scholar] [CrossRef]
  82. Caldwell, C.; Schaphorst, S.; Krause, M.; Jiménez-Mier, J. Photoexcited K Auger spectra of atomic and molecular oxygen. JESRP 1994, 67, 243–259. [Google Scholar] [CrossRef]
  83. Fabian, A.C.; Ross, R.R. X-ray Reflection. Space Sci. Rev. 2010, 157, 167–176. [Google Scholar] [CrossRef]
  84. Badnell, N.R.; Pindzola, M.S.; Dickson, W.J.; Summers, H.P.; Griffin, D.C.; Lang, J. Electric Field Effects on Dielectronic Recombination in a Collisional-Radiative Model. Astrophys. J. 1993, 407, L91. [Google Scholar] [CrossRef]
  85. Nahar, S.N.; Pradhan, A.K. Electron-ion recombination in the close-coupling approximation. Phys. Rev. Lett. 1992, 68, 1488–1491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Nahar, S.N.; Pradhan, A.K. Unified treatment of electron-ion recombination in the close-coupling approximation. Phys. Rev A 1994, 49, 1816–1835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. García, J.; Kallman, T.R. X-ray Reflected Spectra from Accretion Disk Models. I. Constant Density Atmospheres. Astrophys. J. 2010, 718, 695–706. [Google Scholar] [CrossRef] [Green Version]
  88. García, J.; Dauser, T.; Reynolds, C.S.; Kallman, T.R.; McClintock, J.E.; Wilms, J.; Eikmann, W. X-ray Reflected Spectra from Accretion Disk Models. III. A Complete Grid of Ionized Reflection Calculations. Astrophys. J. 2013, 768, 146. [Google Scholar] [CrossRef] [Green Version]
  89. Aldrovandi, S.M.V.; Pequignot, D. Radiative and Dielectronic Recombination Coefficients for Complex Ions. Astron. Astrophys. 1973, 25, 137. [Google Scholar]
  90. Bautista, M.A.; Pradhan, A.K. Ionization Structure and Spectra of Iron in Gaseous Nebulae. Astrophys. J. 1998, 492, 650–676. [Google Scholar] [CrossRef] [Green Version]
  91. Bautista, M.A. Theoretical Calculations of Atomic Data for Spectroscopy. In Atomic Data Needs for X-ray Astronomy; Bautista, M.A., Kallman, T.R., Pradhan, A.K., Eds.; NASA/CP-2000-209968; NASA Goddard Space Flight Center: Greenbelt, MD, USA, 2000; pp. 25–40. [Google Scholar]
  92. Ralchenko, Y.; Kramida, A. Development of NIST Atomic Databases and Online Tools. Atoms 2020, 8, 56. [Google Scholar] [CrossRef]
  93. Albert, D.; Antony, B.K.; Ba, Y.A.; Babikov, Y.L.; Bollard, P.; Boudon, V.; Delahaye, F.; Del Zanna, G.; Dimitrijević, M.S.; Drouin, B.J.; et al. A Decade with VAMDC: Results and Ambitions. Atoms 2020, 8, 76. [Google Scholar] [CrossRef]
  94. Mendoza, C.; Boswell, J.; Ajoku, D.; Bautista, M. AtomPy: An Open Atomic Data Curation Environment for Astrophysical Applications. Atoms 2014, 2, 123–156. [Google Scholar] [CrossRef] [Green Version]
  95. Juan de Dios, L.; Rodríguez, M. The impact of atomic data selection on nebular abundance determinations. Mon. Not. R. Astron. Soc. 2017, 469, 1036–1053. [Google Scholar] [CrossRef] [Green Version]
  96. Morisset, C.; Luridiana, V.; García-Rojas, J.; Gómez-Llanos, V.; Bautista, M. Atomic Data Assessment with PyNeb. Atoms 2020, 8, 66. [Google Scholar] [CrossRef]
  97. Aldrovandi, S.M.V.; Pequignot, D. Erratum; Radiative and Dielectronic Recombination Coefficients for Complex Ions. Astron. Astrophys. 1976, 47, 321. [Google Scholar]
  98. Kingdon, J.B.; Ferland, G.J. Rate Coefficients for Charge Transfer between Hydrogen and the First 30 Elements. Astrophys. J. Suppl. Ser. 1996, 106, 205. [Google Scholar] [CrossRef] [Green Version]
  99. Nussbaumer, H.; Storey, P.J. Dielectronic recombination at low temperatures. III—Recombination coefficients for Mg, Al, Si. Astron. Astrophys. Suppl. Ser. 1986, 64, 545–555. [Google Scholar]
  100. Gould, R.J.; Thakur, R.K. Atomic processes in a low-density hydrogen-helium plasma. Ann. Phys. 1970, 61, 351–386. [Google Scholar] [CrossRef]
  101. Burgess, A.; Tully, J.A. On the analysis of collision strengths and rate coefficients. Astron. Astrophys. 1992, 254, 436–453. [Google Scholar]
  102. Verner, D.A.; Yakovlev, D.G. Analytic FITS for partial photoionization cross sections. Astron. Astrophys. Suppl. Ser. 1995, 109, 125–133. [Google Scholar]
  103. Callaway, J. Effective Collision Strengths for Hydrogen and Hydrogen-Like Ions. At. Data Nucl. Data Tables 1994, 57, 9–20. [Google Scholar] [CrossRef]
  104. Kato, T.; Nakazaki, S. Recommended Data for Excitation Rate Coefficients of Helium Atoms and Helium-like Ions by Electron Impact. Atom. Data Nucl. Data Tables 1989, 42, 313. [Google Scholar] [CrossRef]
  105. Keenan, F.P.; McCann, S.M.; Kingston, A.E. Electron Impact Excitation Rates for Transitions from the Ground State to the n = 2 and 3 levels in Helium-Like Ions. Phys. Scr. 1987, 35, 432–436. [Google Scholar] [CrossRef]
  106. Zhang, H.; Sampson, D.H. Collision Rates for Excitation of Helium-like Ions with Inclusion of Resonance Effects. Astrophys. J. Suppl. Ser. 1987, 63, 487. [Google Scholar] [CrossRef]
  107. Sampson, D.H.; Goett, S.J.; Clark, R.E.H. Electron-Impact Collision Strengths for Inner-Shell Excitation of Doubly Excited Levels from Singly Excited Levels in He-like Ions. At. Data Nucl. Data Tables 1983, 28, 299. [Google Scholar] [CrossRef]
  108. Bhatia, A.K.; Fawcett, B.C.; Phillips, K.J.H.; Lemen, J.R.; Mason, H.E. A comparison of theoretical and solar-flare intensity ratios for the Fe XIX X-ray lines. Mon. Not. R. Astron. Soc. 1989, 240, 421–444. [Google Scholar] [CrossRef] [Green Version]
  109. Gu, M.F.; Holczer, T.; Behar, E.; Kahn, S.M. Inner-Shell Absorption Lines of Fe VI-Fe XVI: A Many-Body Perturbation Theory Approach. Astrophys. J. 2006, 641, 1227–1232. [Google Scholar] [CrossRef] [Green Version]
  110. Foster, A.R.; Ji, L.; Smith, R.K.; Brickhouse, N.S. Updated Atomic Data and Calculations for X-ray Spectroscopy. Astrophys. J. 2012, 756, 128. [Google Scholar] [CrossRef] [Green Version]
  111. Bryans, P.; Badnell, N.R.; Gorczyca, T.W.; Laming, J.M.; Mitthumsiri, W.; Savin, D.W. Collisional Ionization Equilibrium for Optically Thin Plasmas. I. Updated Recombination Rate Coefficients for Bare through Sodium-like Ions. Astrophys. J. Suppl. Ser. 2006, 167, 343–356. [Google Scholar] [CrossRef] [Green Version]
  112. Abu-Salbi, N.; Callaway, J. Electron-impact excitation of hydrogenic ions. Phys. Rev. A 1981, 24, 2372–2386. [Google Scholar] [CrossRef]
  113. Aggarwal, K.M.; Callaway, J.; Kingston, A.E.; Unnikrishnan, K. Excitation Rate Coefficients for Transitions among the N = 1, 2, and 3 Levels of He+. Astrophys. J. Suppl. Ser. 1992, 80, 473. [Google Scholar] [CrossRef]
  114. Aggarwal, K.M.; Kingston, A.E. Electron impact excitation of C VI: Collision strengths and rate coefficients. J. Phys. B At. Mol. Phys. 1991, 24, 4583–4602. [Google Scholar] [CrossRef]
  115. Aggarwal, K.M.; Kingston, A.E. Electron impact excitation of Ne X: Collision strengths and rate coefficients. Phys. Scr. 1991, 44, 517–527. [Google Scholar] [CrossRef]
  116. Aggarwal, K.M.; Kingston, A.E. Electron impact excitation of Si XIV: Collision strengths and rate coefficients. Phys. Scr. 1992, 46, 193–201. [Google Scholar] [CrossRef]
  117. Aggarwal, K.M.; Kingston, A.E. Electron impact excitation of Ca XX: Collision strengths and rate coefficients. J. Phys. B At. Mol. Phys. 1992, 25, 751–771. [Google Scholar] [CrossRef]
  118. Aggarwal, K.M.; Kingston, A.E. Electron Impact Excitation of Fe xxvi: Collision Strengths and Rate Coefficients. Astrophys J. Suppl. 1993, 85, 187. [Google Scholar] [CrossRef]
  119. Binello, A.M.; Mason, H.E.; Storey, P.J. Atomic data from the IRON Project. XXV. Electron impact excitation of fine-structure transitions in the ground configuration of Fe XII. Astron. Astrophys. Suppl. Ser. 1998, 127, 545–553. [Google Scholar] [CrossRef] [Green Version]
  120. Flower, D.R. Excitation of the Fe XII spectrum in the solar corona. Astron. Astrophys. 1977, 54, 163–166. [Google Scholar]
  121. Tayal, S.S.; Henry, R.J.W.; Pradhan, A.K. Electron Excitation of Fine-Structure Transitions within the Ground 3s23p3 Configuration in Fe XII. Astrophys. J. 1987, 319, 951. [Google Scholar] [CrossRef]
  122. Berrington, K.A.; Nakazaki, S.; Norrington, P.H. Atomic data from the IRON Project. XLI. Electron excitation rates among the 3d2 fine-structure levels of Ca-like Fe VII. Astron. Astrophys. Suppl. Ser. 2000, 142, 313–316. [Google Scholar] [CrossRef] [Green Version]
  123. Keenan, F.P.; Norrington, P.H. Relative emission line strengths for Fe VII in astrophysical plasmas. Astron. Astrophys. 1987, 181, 370–372. [Google Scholar]
  124. Gau, J.N.; Henry, R.J.W. Excitation of lithiumlike ions by electron impact. Phys. Rev. A 1977, 16, 986–990. [Google Scholar] [CrossRef]
  125. Wiese, W.L.; Smith, M.W.; Glennon, B.M. Atomic Transition Probabilities. Vol.: Hydrogen through Neon. A Critical Data Compilation; US Government Printing Office: Washington, DC, USA, 1966.
  126. Martin, I.; Karwowski, J.; Diercksen, G.H.F.; Barrientos, C. Transition probabilities in the lithium sequence. Astron. Astrophys. Suppl. Ser. 1993, 100, 595–605. [Google Scholar]
  127. Zhang, H.L.; Sampson, D.H.; Fontes, C.J. Relativistic Distorted-Wave Collision Strengths for the 85 Li-like Ions with 8 <= Z <= 92. At. Data Nucl. Data Tables 1990, 44, 31. [Google Scholar] [CrossRef]
  128. Berrington, K.A.; Burke, P.G.; Dufton, P.L.; Kingston, A.E. Electron-Impact-Excitation Collision Strengths for Be-like Ions. II. Intermediate-Energy Region and Collision Rates. At. Data Nucl. Data Tables 1985, 33, 195. [Google Scholar] [CrossRef]
  129. Berrington, K.A.; Burke, V.M.; Burke, P.G.; Scialla, S. Electron impact excitation of n = 3 states of C III: An application of a new R-matrix package. J. Phys. B At. Mol. Phys. 1989, 22, 665–676. [Google Scholar] [CrossRef]
  130. Muehlethaler, H.P.; Nussbaumer, H. Transition probabilities within 2s2-2s2p-2p2 in the Be I sequence, Be I—Ni XXV. Astron. Astrophys. 1976, 48, 109–114. [Google Scholar]
  131. Sampson, D.H.; Goett, S.J.; Clark, R.E.H. Collision Strengths and Line Strengths for Fine-Structure Transitions between the 2la2lb Configurations and the 2la3lb Configurations in Be-like Ions. At. Data Nucl. Data Tables 1984, 30, 125. [Google Scholar] [CrossRef]
  132. Zhang, H.L.; Sampson, D.H. Relativistic Distorted-Wave Collision Strengths and Oscillator Strengths for the 45 Δn =0 Transitions with n = 2 in the 85 Be-like Ions with 8 <= Z <= 92. At. Data Nucl. Data Tables 1992, 52, 143. [Google Scholar] [CrossRef]
  133. Ramsbottom, C.A.; Berrington, K.A.; Hibbert, A.; Bell, K.L. Electron impact excitation rates for transitions involving the n = 2 and n = 3 levels of beryllium-like N IV. Phys. Scr. 1994, 50, 246–253. [Google Scholar] [CrossRef]
  134. Kato, T.; Lang, J.; Berrington, K.A. Intensity Ratios of Emission Lines for O V Ions for Temperature and Density Diagnostics, and Recommended Excitation Rate Coefficients. At. Data Nucl. Data Tables 1990, 44, 133. [Google Scholar] [CrossRef]
  135. Dankwort, W.; Trefftz, E. Breit-Pauli approximation for highly ionized boron-like ions, up to Fe XXII. Astron. Astrophys. 1978, 65, 93–98. [Google Scholar]
  136. Nussbaumer, H.; Storey, P.J. C II two-electron transitions. Astron. Astrophys. 1981, 96, 91–95. [Google Scholar]
  137. Lennon, D.J.; Dufton, P.L.; Hibbert, A.; Kingston, A.E. C II emission lines formed in optically thin plasmas. Astrophys. J. 1985, 294, 200–206. [Google Scholar] [CrossRef]
  138. Blum, R.D.; Pradhan, A.K. Rate Coefficients for the Excitation of Infrared and Ultraviolet Lines in C II, N III, and O IV. Astrophys. J. Suppl. Ser. 1992, 80, 425. [Google Scholar] [CrossRef]
  139. Nussbaumer, H.; Storey, P.J. N III lines for solar diagnostics. Astron. Astrophys. 1979, 71, L5–L8. [Google Scholar]
  140. Stafford, R.P.; Hibbert, A.; Bell, K.L. Accurate transition probabilities for spectral lines of N III. Mon. Not. R. Astron. Soc. 1993, 260, L11–L14. [Google Scholar] [CrossRef] [Green Version]
  141. Stafford, R.P.; Bell, K.L.; Hibbert, A. Electron impact excitation of N III—Fine-structure collision strengths and Maxwellian-averaged rate coefficients. Mon. Not. R. Astron. Soc. 1994, 266, 715. [Google Scholar] [CrossRef] [Green Version]
  142. Stafford, R.P.; Bell, K.L.; Hibbert, A.; Wijesundera, W.P. Electron Impact Excitation of NII—Fine Structure Collision Strengths and Maxwellian-Averaged Rate Coefficients. Mon. Not. R. Astron. Soc. 1994, 268, 816. [Google Scholar] [CrossRef] [Green Version]
  143. Bhatia, A.K.; Feldman, U.; Seely, J.F. Atomic Data and Spectral Line Intensities for the Boron Isoelectronic Sequence (Ar XIV through Kr XXXII). At. Data Nucl. Data Tables 1986, 35, 319. [Google Scholar] [CrossRef]
  144. Zhang, H.L.; Graziani, M.; Pradhan, A.K. Atomic Data from the Iron Project: III. Rate Coefficients for Electron Impact Excitation of Boron-like Ions: NE VI, MG VIII, AL IX, SI X, S XII, AR XIV, CA XVI and Fe XXII. Astron. Astrophys. 1994, 283, 319. [Google Scholar]
  145. Sampson, D.H.; Weaver, G.M.; Goett, S.J.; Zhang, H.; Clark, R.E.H. Collision Strengths and Line Strengths for All Transitions among the Levels of the 1s22s22p, 1s22s2p2, and 1s22p3, Configurations of Boron-like Ions. At. Data Nucl. Data Tables 1986, 35, 223. [Google Scholar] [CrossRef]
  146. Bell, K.L.; Hibbert, A.; Stafford, R.P. Transition probabilities for some spectral lines of singly ionised nitrogen. Phys. Scr. 1995, 52, 240–251. [Google Scholar] [CrossRef]
  147. Aggarwal, K.M. Electron excitation rates among fine structure levels in O III. Astrophys. J. Suppl. Ser. 1983, 52, 387–398. [Google Scholar] [CrossRef]
  148. Aggarwal, K.M. Line intensity ratios for transitions in O III. Astron. Astrophys. 1985, 146, 149–158. [Google Scholar]
  149. Bhatia, A.K.; Kastner, S.O. Collision Strengths and Transition Rates for O III. At. Data Nucl. Data Tables 1993, 54, 133–164. [Google Scholar] [CrossRef]
  150. Lennon, D.J.; Burke, V.M. Atomic data from the IRON project. II. Effective collision strengths for infrared transitions in carbon-like ions. Astron. Astrophys. Suppl. Ser. 1994, 103, 273–277. [Google Scholar]
  151. Aggarwal, K.M. Electron impact excitation rates for fine-structure transitions in NE V and SI IX—An R-matrix approach. Astrophys. J. Suppl. Ser. 1984, 54, 1–15. [Google Scholar] [CrossRef]
  152. Bhatia, A.K.; Doschek, G.A. Atomic Data and Spectral Line Intensities for C-like Ne V. At. Data Nucl. Data Tables 1993, 55, 315–348. [Google Scholar] [CrossRef]
  153. Bhatia, A.K.; Doschek, G.A. Atomic Data and Spectral Line Intensities for C-Like Si IX. At. Data Nucl. Data Tables 1993, 55, 281–314. [Google Scholar] [CrossRef]
  154. Bhatia, A.K.; Doschek, G.A. Atomic Data and Spectral Line Intensities for C-like Mg VII. At. Data Nucl. Data Tables 1995, 60, 145. [Google Scholar] [CrossRef]
  155. Bhatia, A.K.; Kastner, S.O. Atomic parameters for carbon-like S XI. J. Quant. Spectros. Radiat. Transf. 1987, 37, 527–529. [Google Scholar] [CrossRef]
  156. Dere, K.P.; Mason, H.E.; Widing, K.G.; Bhatia, A.K. XUV electron density diagnostics for solar flares. Astrophys. J. Suppl. Ser. 1979, 40, 341–364. [Google Scholar] [CrossRef]
  157. Mason, H.E.; Bhatia, A.K. Theoretical intensity ratios for the UV lines of Mg VII, Si IX and S XI. Mon. Not. R. Astron. Soc. 1978, 184, 423–437. [Google Scholar] [CrossRef] [Green Version]
  158. Froese Fischer, C.; Saha, H.P. Multiconfiguration Hartree-Fock Results with Breit-Pauli Corrections for Transitions in the Carbon Sequence. Phys. Scr. 1985, 32, 181–194. [Google Scholar] [CrossRef]
  159. Aggarwal, K.M.; Berrington, K.A.; Keenan, F.P. Effective Collision Strengths among Fine-Structure Levels of CA XV. Astrophys. J. Suppl. Ser. 1990, 72, 815. [Google Scholar] [CrossRef]
  160. Aggarwal, K.M.; Berrington, K.A.; Keenan, F.P. Effective collision strengths among fine-structure levels of CA XV: Erratum. Astrophys. J. Suppl. Ser. 1991, 77, 441–449. [Google Scholar] [CrossRef] [Green Version]
  161. Bhatia, A.K.; Mason, H.E. Theoretical atomic structure and electron scattering data for ions in the nitrogen isoelectronic sequence—MG VI, SI VIII, S X, A XII and CA XIV. Mon. Not. R. Astron. Soc. 1980, 190, 925–929. [Google Scholar] [CrossRef] [Green Version]
  162. Bhatia, A.K.; Seely, J.F.; Feldman, U. Atomic Data and Spectral Line Intensities for the Nitrogen Isoelectronic Sequence (Ar XII through Kr XXX). At. Data Nucl. Data Tables 1989, 43, 99. [Google Scholar] [CrossRef]
  163. Mason, H.E.; Bhatia, A.K.; Doschek, G.A.; Feldman, U. Fe XXI as an electron density diagnostic in solar flares. Astron. Astrophys. 1979, 73, 74–81. [Google Scholar]
  164. Aggarwal, K.M. Collision Strengths and Excitation Rate Coefficients for Transitions in Fe XXI. Astrophys. J. Suppl. Ser. 1991, 77, 677. [Google Scholar] [CrossRef]
  165. Bhatia, A.K.; Feldman, U.; Doschek, G.A. New atomic data for Si/6+/, S/8+/ and Ar/10+/. Astron. Astrophys. 1979, 80, 22–26. [Google Scholar]
  166. McLaughlin, B.M.; Bell, K.L. Analytical FITS for Electron Collisionally Excited Effective Collision Strengths of O II. Astrophys. J. Suppl. Ser. 1994, 94, 825. [Google Scholar] [CrossRef]
  167. Loulergue, M.; Mason, H.E.; Nussbaumer, H.; Storey, P.J. Fe XIX transitions within the N = 2 complex. Astron. Astrophys. 1985, 150, 246–250. [Google Scholar]
  168. Blackford, H.M.S.; Hibbert, A. Transitions in Fluorine-Like Ions. At. Data Nucl. Data Tables 1994, 58, 101–164. [Google Scholar] [CrossRef]
  169. Butler, K.; Zeippen, C.J. Atomic data from the IRON Project. V. Effective collision strengths for transitions in the ground configuration of oxygen-like ions. Astron. Astrophys. Suppl. Ser. 1994, 108, 1–8. [Google Scholar]
  170. Sampson, D.H.; Zhang, H.L.; Fontes, C.J. Relativistic Distorted-Wave Collision Strengths and Oscillator Strengths for F-like Ions with 22 <= Z <= 92. At. Data Nucl. Data Tables 1991, 48, 25. [Google Scholar] [CrossRef]
  171. Cornille, M.; Dubau, J.; Loulergue, M.; Bely-Dubau, F.; Faucher, P. Theoretical analysis of the Fe XVIII X-ray spectrum and application to solar coronal observations. Astron. Astrophys. 1992, 259, 669–681. [Google Scholar]
  172. Bhatia, A.K.; Feldman, U.; Seely, J.F. Atomic Data and Spectral Line Intensities for the Neon Isoelectronic Sequence (Si V through Kr XXVII). At. Data Nucl. Data Tables 1985, 32, 435. [Google Scholar] [CrossRef]
  173. Hibbert, A.; Ledourneuf, M.; Mohan, M. Energies, Oscillator Strengths, and Lifetimes for Neon-like Ions Up to Kr XXVII. At. Data Nucl. Data Tables 1993, 53, 23–112. [Google Scholar] [CrossRef]
  174. Saraph, H.E.; Tully, J.A. Atomic data from the IRON project. IV. Electron excitation of the 2P3/2∘–2P1/2∘ fine structure transition in fluorine-like ions. Astron. Astrophys. Suppl. Ser. 1994, 107, 29–38. [Google Scholar]
  175. Mohan, M.; Baluja, K.L.; Hibbert, A. Electron-impact excitation of MG IV using the R-matrix method. Phys. Scr. 1988, 38, 699–701. [Google Scholar] [CrossRef]
  176. Zhang, H.; Sampson, D.H.; Clark, R.E.H.; Mann, J.B. Collision Strengths and Oscillator Strengths for Excitation to the n = 3 and 4 Levels of Neon-like Ions. At. Data Nucl. Data Tables 1987, 37, 17. [Google Scholar] [CrossRef]
  177. Bhatia, A.K.; Doschek, G.A. Atomic Data and Spectral Line Intensities for Ne-like Fe XVII. At. Data Nucl. Data Tables 1992, 52, 1. [Google Scholar] [CrossRef]
  178. Mohan, M.; Le Dourneuf, M. Electron-impact excitation from the Si(5+) ground state, using the R-matrix method. Astron. Astrophys. 1990, 227, 285–288. [Google Scholar]
  179. Mohan, M.; Hibbert, A.; Baluja, K.L. Electron impact excitation of S VIII from its ground state. J. Phys. B At. Mol. Phys. 1987, 20, 2565–2570. [Google Scholar] [CrossRef]
  180. Sampson, D.H.; Zhang, H.L.; Fontes, C.J. Relativistic Distorted-Wave Collision Strengths and Oscillator Strengths for the 71 Na-like Ions with 22 <= Z <= 92. At. Data Nucl. Data Tables 1990, 44, 209. [Google Scholar] [CrossRef]
  181. Dufton, P.L.; Hibbert, A.; Kingston, A.E.; Doschek, G.A. Si III line ratios in the Sun. Astrophys. J. 1983, 274, 420–428. [Google Scholar] [CrossRef]
  182. Mason, H.E. The excitation of several iron and calcium lines in the visible spectrum of the solar corona. Mon. Not. R. Astron. Soc. 1975, 170, 651–689. [Google Scholar] [CrossRef]
  183. Christensen, R.B.; Norcross, D.W.; Pradhan, A.K. Electron-impact excitation of ions in the magnesium sequence. II. S v, Ar vii, Ca ix, Cr xiii, and Ni xvii. Phys. Rev. A 1986, 34, 4704–4715. [Google Scholar] [CrossRef]
  184. Lanzafame, A.C. Si II resonance multiplets in the Sun. Astron. Astrophys. 1994, 287, 972–981. [Google Scholar]
  185. Mohan, M.; Hibbert, A.; Berrington, K.A.; Burke, P.G. Electron impact excitation of neon-like S VII from its ground state. J. Phys. B At. Mol. Phys. 1990, 23, 989–996. [Google Scholar] [CrossRef]
  186. Bhatia, A.K.; Doschek, G.A.; Feldman, U. Atomic data for S IV and solar observations of the 3s23p2P–3s3p24P multiplet. Astron. Astrophys. 1980, 86, 32–35. [Google Scholar]
  187. Dufton, P.L.; Hibbert, A.; Kingston, A.E.; Doschek, G.A. S IV emission-line ratios in the sun. Astrophys. J. 1982, 257, 338–344. [Google Scholar] [CrossRef]
  188. Fischer, C.F.; Liu, B. Energy Levels, Lifetimes, Wavelengths, and Transition Probabilities of Fe XIV. At. Data Nucl. Data Tables 1986, 34, 261. [Google Scholar] [CrossRef]
  189. Czyzak, S.J.; Krueger, T.K. On the Excited Levels of Fe VIII. Astrophys. J. 1966, 144, 381. [Google Scholar] [CrossRef]
  190. Fawcett, B.C. Computed Oscillator Strengths and Energy Levels for Potassium-like Ions V V, Cr VI, Mn VII, Fe VIII, Co IX, Ni X, and Cu XI with Calculated Wavelengths and Wavelengths Derived from Established Data. At. Data Nucl. Data Tables 1989, 43, 71. [Google Scholar] [CrossRef]
  191. Dufton, P.L.; Kingston, A.E. Effective collision strengths for Si III. Mon. Not. R. Astron. Soc. 1989, 241, 209–214. [Google Scholar] [CrossRef] [Green Version]
  192. Nussbaumer, H.; Storey, P.J.; Storey, P.J. Forbidden emission lines of Fe VII. Astron. Astrophys. 1982, 113, 21–26. [Google Scholar]
  193. Dufton, P.L.; Kingston, A.E. Effective collision strengths for Si II. Mon. Not. R. Astron. Soc. 1991, 248, 827–828. [Google Scholar] [CrossRef] [Green Version]
  194. Bhadra, K.; Henry, R.J.W. Oscillator strengths and collision strengths for S IV. Astrophys. J. 1980, 240, 368–373. [Google Scholar] [CrossRef]
  195. Fawcett, B.C.; Mason, H.E. Collision Strengths and Oscillator Strengths for Fe IX. At. Data Nucl. Data Tables 1991, 47, 17. [Google Scholar] [CrossRef]
  196. Mason, H.E. An Assessment of Theoretical Electron Excitation Data for Fe IX-Fe XIV. At. Data Nucl. Data Tables 1994, 57, 305–328. [Google Scholar] [CrossRef]
  197. Pindzola, M.S.; Griffin, D.C.; Bottcher, C. Coupling effects for electron-impact excitation in the potassium isoelectronic sequence. Phys. Rev. A 1989, 39, 2385–2391. [Google Scholar] [CrossRef]
  198. Fawcett, B.C.; Mason, H.E. A Slater Parameter Optimization Method Applied to the Computation of Collision Strengths for Fe XIII. At. Data Nucl. Data Tables 1989, 43, 245. [Google Scholar] [CrossRef]
  199. Aggarwal, K.M.; Berrington, K.A.; Burke, P.G.; Kingston, A.E.; Pathak, A. Electron collision cross sections at low energies for all transitions between the n=1, 2, 3, 4 and 5 levels of atomic hydrogen. J. Phys. B At. Mol. Phys. 1991, 24, 1385–1410. [Google Scholar] [CrossRef]
  200. Chen, G.X.; Pradhan, A.K. Atomic data from the Iron Project. XXXVII. Electron impact excitation collision strengths and rate coefficients for Fe VI. Astron. Astrophys. Suppl. Ser. 1999, 136, 395–403. [Google Scholar] [CrossRef]
  201. Eissner, W.; Galavís, M.E.; Mendoza, C.; Zeippen, C.J. Atomic data from the IRON Project. XXXIV. Electron impact excitation of Fe XVI. Astron. Astrophys. Suppl. Ser. 1999, 136, 385–394. [Google Scholar] [CrossRef] [Green Version]
  202. Eissner, W.; Galavís, M.E.; Mendoza, C.; Zeippen, C.J. Atomic data from the IRON Project. XXXVIII. Electron impact excitation of the fine-structure transitions in the n = 3 complex of Fe XV. Astron. Astrophys. Suppl. Ser. 1999, 137, 165–173. [Google Scholar] [CrossRef]
  203. Garstang, R.H. Transition probabilities for forbidden lines of Fe II. Mon. Not. R. Astron. Soc. 1962, 124, 321. [Google Scholar] [CrossRef]
  204. Nussbaumer, H.; Storey, P.J. Transition probabilities for Fe II infrared lines. Astron. Astrophys. 1988, 193, 327–333. [Google Scholar]
  205. Giridhar, S.; Arellano Ferro, A. A Critical Compilation of Oscillator Strengths for Fe II Lines. Rev. Mex. Astron. Astrofis. 1995, 31, 23–37. [Google Scholar]
  206. Nahar, S.N. Atomic data from the Iron Project. VII. Radiative dipole transition probabilities for Fe II. Astron. Astrophys. 1995, 293, 967–977. [Google Scholar]
  207. Pelan, J.; Berrington, K.A. Atomic data from the IRON Project. IX. Electron excitation of the 2P3/2∘–2P1/2∘ fine-structure transition in chlorine-like ions, from AR II to NI XII. Astron. Astrophys. Suppl. Ser. 1995, 110, 209. [Google Scholar]
  208. Berrington, K.A.; Tully, J.A. Atomic data from the IRON Project. XXIV. Electron excitation of Li-like Fe XXIV between the N = 2 and n = 2, 3, 4 fine-structure levels. Astron. Astrophys. Suppl. Ser. 1997, 126, 105–111. [Google Scholar] [CrossRef] [Green Version]
  209. Garstang, R.H. Energy levels and transition probabilities of Fe IV. Mon. Not. R. Astron. Soc. 1958, 118, 572. [Google Scholar] [CrossRef] [Green Version]
  210. Kono, A.; Hattori, S. Accurate oscillator strengths for neutral helium. Phys. Rev. A 1984, 29, 2981–2988. [Google Scholar] [CrossRef]
  211. Goett, S.J.; Sampson, D.H.; Clark, R.E.H. Intermediate-Coupling Collision Strengths for Transitions between Doubly Excited Levels Produced by Electron Impact on Highly Charged He-like Ions. At. Data Nucl. Data Tables 1983, 28, 279. [Google Scholar] [CrossRef]
  212. Kaufman, V.; Sugar, J. Forbidden Lines in ns2npk Ground Configurations and nsnp Excited Configurations of Beryllium through Molybdenum Atoms and Ions. J. Phys. Chem. Ref. Data 1986, 15, 321–426. [Google Scholar] [CrossRef]
  213. Nussbaumer, H.; Rusca, C. Forbidden transitions in the C I sequence. Astron. Astrophys. 1979, 72, 129–133. [Google Scholar] [CrossRef]
  214. Mendoza, C. Recent advances in atomic calculations and experiments of interest in the study of planetary nebulae. In Planetary Nebulae; Flower, D.R., Ed.; D. Reidel Publishing Co.: Dordrecht, Holland, 1983; IAU Symposium 103; pp. 143–172. [Google Scholar]
  215. Zeippen, C.J. Transition probabilities for forbidden lines in the 2p3 configuration. Mon. Not. R. Astron. Soc. 1982, 198, 111–125. [Google Scholar] [CrossRef] [Green Version]
  216. Nahar, S.N.; Pradhan, A.K. Atomic data from the Iron Project. XVII. Radiative transition probabilities for dipole allowed and forbidden transitions in Fe III. Astron. Astrophys. Suppl. Ser. 1996, 119, 509–522. [Google Scholar] [CrossRef]
  217. Butler, K.; Zeippen, C.J. N I forbidden lines revisited. Astron. Astrophys. 1984, 141, 274–278. [Google Scholar]
  218. Bautista, M.A. Atomic data from the Iron Project. XXXIX. Photoionization cross sections and oscillator strengths for NI II. Astron. Astrophys. Suppl. Ser. 1999, 137, 529–535. [Google Scholar] [CrossRef] [Green Version]
  219. Bautista, M.A. Continuum Fluorescence Excitation of [N I] and [O I] Lines in Gaseous Nebulae. Astrophys. J. 1999, 527, 474–478. [Google Scholar] [CrossRef]
  220. Moore, C.E. Selected Tables of Atomic Spectra. Atomic Energy Levels and Multiplet Table O III; US Government Printing Office: Washington, DC, USA, 1985. [Google Scholar]
  221. Thomas, L.D.; Nesbet, R.K. Low-energy electron scattering by atomic carbon. Phys. Rev. A 1975, 12, 2378–2382. [Google Scholar] [CrossRef]
  222. Pequignot, D.; Aldrovandi, S.M.V. Effective collision strengths for the forbidden emission lines of neutral carbon, nitrogen and oxygen. Astron. Astrophys. 1976, 50, 141–144. [Google Scholar]
  223. Johnson, C.T.; Burke, P.G.; Kingston, A.E. Electron scattering from the fine structure levels within the 1s22s22p23Pe ground state of C I. J. Phys. B At. Mol. Phys. 1987, 20, 2553–2563. [Google Scholar] [CrossRef]
  224. Dopita, M.A.; Mason, D.J.; Robb, W.D. Atomic nitrogen as a probe of physical conditions in the interstellar medium. Astrophys. J. 1976, 207, 102–109. [Google Scholar] [CrossRef]
  225. Berrington, K.A.; Burke, P.G. Effective collision strengths for forbidden transitions in e-N and e-O scattering. Planet. Space Sci. 1981, 29, 377–381. [Google Scholar] [CrossRef]
  226. LeDourneuf, M.; Nesbet, R.K. Electron impact excitation of fine-structure levels of atomic oxygen. J. Phys. B At. Mol. Phys. 1976, 9, L241–L243. [Google Scholar] [CrossRef]
  227. Zeippen, C.J.; Butler, K.; Le Bourlot, J. Effective collision strengths for fine-structure forbidden transitions in the 3p3 configuration of AR IV. Astron. Astrophys. 1987, 188, 251–257. [Google Scholar]
  228. Krueger, T.K.; Czyzak, S.J. Excitation of Forbidden Lines in Gaseous Nebulae. II. Calculations for 3pq Ions. Proc. R. Soc. Lon. Ser. A 1970, 318, 531–539. [Google Scholar] [CrossRef]
  229. Mendoza, C.; Zeippen, C.J. Transition probabilities for forbidden lines in the 3p2 configuration-II. Mon. Not. R. Astron. Soc. 1982, 199, 1025–1032. [Google Scholar] [CrossRef] [Green Version]
  230. Hayes, M.A. The S III intercombination multiplet—Radiative and electron excitation data. J. Phys. B At. Mol. Phys. 1986, 19, 1853–1862. [Google Scholar] [CrossRef]
  231. Mendoza, C.; Zeippen, C.J. Transition probabilities for forbidden lines in the 3p3 configuration. Mon. Not. R. Astron. Soc. 1982, 198, 127–139. [Google Scholar] [CrossRef] [Green Version]
  232. Fuhr, J.R.; Wiese, W.L. RCR Handbook of Chemistry and Physics, 71st ed.; CRC Press: Boca Raton, FL, USA, 1990. [Google Scholar]
  233. Ramsbottom, C.A.; Bell, K.L.; Stafford, R.P. Effective Collision Strengths for Electron Impact Excitation of Singly Ionized Sulfur. At. Data Nucl. Data Tables 1996, 63, 57. [Google Scholar] [CrossRef]
  234. Cai, W.; Pradhan, A.K. Excitation Rate Coefficients and Line Ratios for the Optical and Ultraviolet Transitions in S II. Astrophys. J. Suppl. Ser. 1993, 88, 329. [Google Scholar] [CrossRef]
  235. Storey, P.J.; Mason, H.E.; Young, P.R. Atomic data from the IRON Project. XL. Electron impact excitation of the Fe XIV EUV transitions. Astron. Astrophys. Suppl. Ser. 2000, 141, 285–296. [Google Scholar] [CrossRef] [Green Version]
  236. Saraph, H.E.; Storey, P.J. Atomic data from the IRON Project. XI. The 2P1/2∘–2P3/2∘ fine-structure lines of AR VI, K VII and CA VIII. Astron. Astrophys. Suppl. Ser. 1996, 115, 151. [Google Scholar]
  237. Zhang, H. Atomic data from the Iron Project. XVIII. Electron impact excitation collision strengths and rate coefficients for Fe III. Astron. Astrophys. 1996, 119, 523–528. [Google Scholar] [CrossRef]
  238. Zhang, H.L.; Pradhan, A.K. Atomic data from the Iron Project. XXVII. Electron impact excitation collision strengths and rate coefficients for Fe IV. Astron. Astrophys. Suppl. Ser. 1997, 126, 373–378. [Google Scholar] [CrossRef] [Green Version]
  239. Fernley, J.A.; Seaton, M.J.; Taylor, K.T. Atomic data for opacity calculations. VII. Energy levels, f values and photoionisation cross sections for He-like ions. J. Phys. B At. Mol. Phys. 1987, 20, 6457–6476. [Google Scholar] [CrossRef]
  240. Peach, G.; Saraph, H.E.; Seaton, M.J. Atomic data for opacity calculations. IX. The lithium isoelectronic sequence. J. Phys. B At. Mol. Phys. 1988, 21, 3669–3683. [Google Scholar] [CrossRef]
  241. Tully, J.A.; Seaton, M.J.; Berrington, K.A. Atomic data for opacity calculations. XIV. The beryllium sequence. J. Phys. B At. Mol. Phys. 1990, 23, 3811–3837. [Google Scholar] [CrossRef]
  242. Fernley, J.A.; Hibbert, A.; Kingston, A.E.; Seaton, M.J. Atomic data for opacity calculations: XXIV. The boron-like sequence. J. Phys. B At. Mol. Phys. 1999, 32, 5507–5522. [Google Scholar] [CrossRef]
  243. Luo, D.; Pradhan, A.K. Atomic data for opacity calculations. XI. The carbon isoelectronic sequence. J. Phys. B At. Mol. Phys. 1989, 22, 3377–3395. [Google Scholar] [CrossRef]
  244. Hibbert, A.; Scott, M.P. Atomic data for opacity calculations. XXI. The neon sequence. J. Phys. B At. Mol. Phys. 1994, 27, 1315–1323. [Google Scholar] [CrossRef]
  245. Butler, K.; Mendoza, C.; Zeippen, C.J. Atomic data for opacity calculations. XIX. The magnesium isoelectronic sequence. J. Phys. B At. Mol. Phys. 1993, 26, 4409–4423. [Google Scholar] [CrossRef]
  246. Mendoza, C.; Eissner, W.; LeDourneuf, M.; Zeippen, C.J. Atomic data for opacity calculations. XXIII. The aluminium isoelectronic sequence. J. Phys. B At. Mol. Phys. 1995, 28, 3485–3504. [Google Scholar] [CrossRef]
  247. Nahar, S.N.; Pradhan, A.K. Atomic data for opacity calculations. XVIII. Photoionization and oscillator strengths of Si-like ions Si0, S2+, Ar4+, Ca6+. J. Phys. B At. Mol. Phys. 1993, 26, 1109–1127. [Google Scholar] [CrossRef]
  248. Bautista, M.A. Atomic data from the Iron Project. XVI. Photoionization cross sections and oscillator strengths for Fe V. Astron. Astrophys. Suppl. Ser. 1996, 119, 105–110. [Google Scholar] [CrossRef] [Green Version]
  249. Bautista, M.A.; Pradhan, A.K. Atomic data from the Iron Project. XXVI. Photoionization cross sections and oscillator strengths for Fe IV. Astron. Astrophys. Suppl. Ser. 1997, 126, 365–371. [Google Scholar] [CrossRef] [Green Version]
  250. Nahar, S.N. Photoionization cross sections and oscillator strengths for Fe III. Phys. Rev. A 1996, 53, 1545–1552. [Google Scholar] [CrossRef] [Green Version]
  251. Nahar, S.N.; Pradhan, A.K. Atomic data for opacity calculations: XX. Photoionization cross sections and oscillator strengths for Fe II. J. Phys. B At. Mol. Phys. 1994, 27, 429–446. [Google Scholar] [CrossRef] [Green Version]
  252. Bautista, M.A. Atomic data from the IRON Project. XX. Photoionization cross sections and oscillator strengths for Fe I. Astron. Astrophys. Suppl. Ser. 1997, 122, 167–176. [Google Scholar] [CrossRef] [Green Version]
  253. Gorczyca, T.W. Auger decay of the photoexcited 1s-1np Rydberg series in neon. Phys. Rev. A 2000, 61, 024702. [Google Scholar] [CrossRef]
Figure 1. hitomi spectrum of the Perseus Cluster in the Fe xxv “triplet” spectral region (6.50–6.60 keV) [47] showing the xstar fit (red line). With the higher resolution, the He-like “triplet” becomes a quartet comprising the resonance (w), magnetic quadrupole (x), intercombination (y), and magnetic dipole (z) lines. Other Fe satellite lines begin to be detectable (e.g., the bump at ∼6.5 keV).
Figure 1. hitomi spectrum of the Perseus Cluster in the Fe xxv “triplet” spectral region (6.50–6.60 keV) [47] showing the xstar fit (red line). With the higher resolution, the He-like “triplet” becomes a quartet comprising the resonance (w), magnetic quadrupole (x), intercombination (y), and magnetic dipole (z) lines. Other Fe satellite lines begin to be detectable (e.g., the bump at ∼6.5 keV).
Atoms 09 00012 g001
Figure 2. Content of file d 454 t 006 of the xstar database listing the atomic model (data type d t = 06 and rate type r t = 13 ) for Mg-like Zn xix ( i o n = 454 ).
Figure 2. Content of file d 454 t 006 of the xstar database listing the atomic model (data type d t = 06 and rate type r t = 13 ) for Mg-like Zn xix ( i o n = 454 ).
Atoms 09 00012 g002
Figure 3. Excerpt of file d 454 t 050 of the xstar database listing the radiative transitions with upper level k 10 (data type d t = 50 and rate type r t = 4 ) in Mg-like Zn xix ( i o n = 454 ).
Figure 3. Excerpt of file d 454 t 050 of the xstar database listing the radiative transitions with upper level k 10 (data type d t = 50 and rate type r t = 4 ) in Mg-like Zn xix ( i o n = 454 ).
Atoms 09 00012 g003
Figure 4. Content of file d 454 t 086 of the xstar database listing the radiative and Auger widths (data type d t = 86 and rate type r t = 4 ) of Mg-like Zn xix ( i o n = 454 ).
Figure 4. Content of file d 454 t 086 of the xstar database listing the radiative and Auger widths (data type d t = 86 and rate type r t = 4 ) of Mg-like Zn xix ( i o n = 454 ).
Atoms 09 00012 g004
Figure 5. Photoabsorption cross-section of the 1 s 2 2 s 2 2 p 6 1 S ground state of Ne-like Fe xvii in the K-edge region. Top panel: undamped cross-section. Bottom panel: damped (radiation and Auger) cross-section.
Figure 5. Photoabsorption cross-section of the 1 s 2 2 s 2 2 p 6 1 S ground state of Ne-like Fe xvii in the K-edge region. Top panel: undamped cross-section. Bottom panel: damped (radiation and Auger) cross-section.
Atoms 09 00012 g005
Figure 6. K-edge photoabsorption (full curve) and photoionization (dotted curve) cross-sections of Ca xv (C-like) as a function of photon energy. The dashed curve gives the photoabsorption cross-section convolved with a Gaussian of 0.001 width. Reproduced from Figure 3 of [18] with permission of the AAS.
Figure 6. K-edge photoabsorption (full curve) and photoionization (dotted curve) cross-sections of Ca xv (C-like) as a function of photon energy. The dashed curve gives the photoabsorption cross-section convolved with a Gaussian of 0.001 width. Reproduced from Figure 3 of [18] with permission of the AAS.
Atoms 09 00012 g006
Figure 7. Probability density of the Hartree–Fock 2p orbital in the initial O i 1 s 2 2 s 2 2 p 4 ground state compared to the 2p orbital in the final O ii 1 s 2 s 2 2 p 4 K-vacancy state. The 2p electrons of the latter are screened from the nucleus by one fewer inner-shell 1s electron and are therefore “relaxed” to more tightly bound orbitals.
Figure 7. Probability density of the Hartree–Fock 2p orbital in the initial O i 1 s 2 2 s 2 2 p 4 ground state compared to the 2p orbital in the final O ii 1 s 2 s 2 2 p 4 K-vacancy state. The 2p electrons of the latter are screened from the nucleus by one fewer inner-shell 1s electron and are therefore “relaxed” to more tightly bound orbitals.
Atoms 09 00012 g007
Figure 8. Ionization potential lowering in Fe ix–Fe xxv as a function of effective charge Z eff = Z N + 1 obtained with the mcdf method using a Debye–Hückel screening potential. Red circles: μ = 0.1 a.u. Black circles: μ = 0.25 a.u. Triangles: Debye–Hückel limit Δ E 0 DH = 27.2116 μ Z eff eV. Reproduced from Figure 1 of [43] with permission from Astronomy & Astrophysics, ©ESO.
Figure 8. Ionization potential lowering in Fe ix–Fe xxv as a function of effective charge Z eff = Z N + 1 obtained with the mcdf method using a Debye–Hückel screening potential. Red circles: μ = 0.1 a.u. Black circles: μ = 0.25 a.u. Triangles: Debye–Hückel limit Δ E 0 DH = 27.2116 μ Z eff eV. Reproduced from Figure 1 of [43] with permission from Astronomy & Astrophysics, ©ESO.
Atoms 09 00012 g008
Figure 9. Cooling- and heating-rate ratios with and without DR obtained from an xstar model with solar abundances, a power-law spectrum, density n e = 10 10 cm 3 , and luminosity L = 10 32 erg s 1 . Black dotted curve: thermal temperature with DR. Red dotted curve: thermal temperature without DR.
Figure 9. Cooling- and heating-rate ratios with and without DR obtained from an xstar model with solar abundances, a power-law spectrum, density n e = 10 10 cm 3 , and luminosity L = 10 32 erg s 1 . Black dotted curve: thermal temperature with DR. Red dotted curve: thermal temperature without DR.
Atoms 09 00012 g009
Figure 10. Reflected X-ray spectrum from an optically thick atmosphere modeled using xillver with and without DR. A plane-parallel atmosphere with solar abundances and density log ( n e ) = 18 cm 3 was assumed.
Figure 10. Reflected X-ray spectrum from an optically thick atmosphere modeled using xillver with and without DR. A plane-parallel atmosphere with solar abundances and density log ( n e ) = 18 cm 3 was assumed.
Atoms 09 00012 g010
Figure 11. Grid tool in uaDB.
Figure 11. Grid tool in uaDB.
Atoms 09 00012 g011
Table 1. Comparison of astronomically observed, laboratory measured, and computed O i K α and K β line energies (eV).
Table 1. Comparison of astronomically observed, laboratory measured, and computed O i K α and K β line energies (eV).
MethodSource E ( 1 s 2 p ) E ( 1 s 3 p ) Δ E
Astronomical observationsXMM-Newton, Mrk 421 [23]527.30(5)541.95(28)14.65(33)
Chandra, 7 sources [23]527.44(9)541.72(18)14.28(21)
Chandra, shifted [23]527.26(9)
Chandra, 11 sources [77]527.39(2)
Chandra, 6 sources [78]527.41(18)541.77(40)14.36(58)
Laboratory measurementsHZB [52]527.26(4)541.645(12)
ALS [79,80]526.79(4)541.19(4)14.40(8)
ALS [76]526.79(4)541.20(4)14.41(8)
WSRC [81]527.85(10)541.27(15)13.41(25)
Auger spectroscopy [82]527.20(30)
mchf converged resultmchf [23]527.49
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mendoza, C.; Bautista, M.A.; Deprince, J.; García, J.A.; Gatuzz, E.; Gorczyca, T.W.; Kallman, T.R.; Palmeri, P.; Quinet, P.; Witthoeft, M.C. The XSTAR Atomic Database. Atoms 2021, 9, 12. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms9010012

AMA Style

Mendoza C, Bautista MA, Deprince J, García JA, Gatuzz E, Gorczyca TW, Kallman TR, Palmeri P, Quinet P, Witthoeft MC. The XSTAR Atomic Database. Atoms. 2021; 9(1):12. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms9010012

Chicago/Turabian Style

Mendoza, Claudio, Manuel A. Bautista, Jérôme Deprince, Javier A. García, Efraín Gatuzz, Thomas W. Gorczyca, Timothy R. Kallman, Patrick Palmeri, Pascal Quinet, and Michael C. Witthoeft. 2021. "The XSTAR Atomic Database" Atoms 9, no. 1: 12. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms9010012

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop