Correlation effects on the fine-structure splitting within the 3d 9 ground configuration in highly-charged Co-like ions
Guo Xue-Ling 1, 2 , Huang Min 1, 2 , Yan Jun 3, 4 , Li Shuang 1, 2 , Wang Kai 3, 4, 5 , Si Ran 1, 2 , Chen Chong-Yang 1, 2, †,
Applied Ion Beam Physics Laboratory of Key Laboratory of the Ministry of Education, Fudan University, Shanghai 200433, China
Shanghai EBIT Lab, Institute of Modern Physics, Department of Nuclear Science and Technology, Fudan University, Shanghai 200433, China
Institute of Applied Physics and Computational Mathematics, Beijing 100088, China
Center for Applied Physics and Technology, Peking University, Beijing 100871, China
Hebei Key Lab of Optic-electronic Information and Materials, the College of PhysicsScience and Technology, Hebei University, Baoding 071002, China

 

† Corresponding author. E-mail: chychen@fudan.edu.cn

Project supported by the National Natural Science Foundation of China (Grant Nos. 11076009 and 11374062), the Chinese Association of Atomic and Molecular Data, the Chinese National Fusion Project for ITER (Grant No. 2015GB117000), and the Leading Academic Discipline Project of Shanghai City, China (Grant No. B107).

Abstract
Abstract

A comprehensive theoretical study of correlation effects on the fine-structure splitting within the ground configuration 3d 9 of the Co-like Hf 45+ , Ta 46+ , W 47+ , and Au 52+ ions is performed by employing the multi-configuration Dirac–Hartree–Fock method in the active space approximation. It shows that the core-valence correlation with the inner-core 2p electron is more significant than with the outer 3p and 3s electrons, and the correlation with the 2s electron is also noticeable. The core–core correlation seems to be small and can be ignored. The calculated 2 D 3/2,5/2 splitting energies agree with the recent electron-beam ion-trap measurements [ Phys. Rev. A 83 032517 (2011), Eur. Phys. J. D 66 286 (2012)] to within the experimental uncertainties.

1. Introduction

Atomic spectroscopic properties of highly-charged ions (HCIs) are of great importance in atomic physics, astronomical physics, and fusion plasma research. [ 1 4 ] Among them the electric-dipole forbidden transition lines are of particular importance in plasma diagnostics due to their sensitive intensities to electron temperature and density and other plasma parameters. [ 5 11 ] The currently famous elements (for example, tungsten and gold) have inspired great research interest since their favorable physical properties, and many spectra lines have been measured with high accuracy. [ 12 22 ] Meanwhile, the accurately measured properties, such as wavelength, can provide crucial tests for atomic structure theory of both the electron correlation and relativistic corrections, [ 23 , 24 ] which are the major questions in theoretical calculations.

Ralchenko et al. [ 24 ] and Osin et al. [ 25 ] measured many forbidden lines from the HCIs of hafnium (Hf, Z = 72), tantalum (Ta, Z = 73), tungsten (W, Z = 74), and gold (Au, Z = 79) with an open 3d shell, with a wavelength uncertainty of 0.03%. Unfortunately, the calculated wavelengths, obtained by Ralchenko and Osin  et al. [ 24 , 25 ] using the relativistic configuration-interaction (RCI) method implemented in the flexible atomic code (FAC), [ 26 , 27 ] can only agree with the experimental ones in 0.8%. Quinet [ 28 ] later obtained the energy levels, wavelengths, and transition probabilities for tungsten ions between W 47+ and W 61+ by using the fully relativistic multi-configuration Dirac–Fock (MCDF) method with the latest version of GRASP, i.e., the general purpose relativistic atomic structure package by Norrington et al. [ 29 ] However, due to the limited configuration interaction considered, the calculated wavelengths differ from the measured ones by up to 0.7%. Fournier [ 30 ] also performed calculations for W 47+ by using the fully relativistic parametric potential code (RELAC), [ 31 ] but the quantum electrodynamic (QED) corrections were not included. [ 32 ] It is obvious that the differences between these calculated wavelengths and measured ones are much larger than the stated experimental uncertainties, which may mean that there is a long way to go on the theoretical side.

Recently, by means of the second-order relativistic many-body perturbation theory (RMBPT) approach, we have updated the energy levels and the transition data for these forbidden transitions among the levels within the ground configurations [Ar] 3d k ( k = 1,...,9) in Co-like through K-like ions of Hf, Ta, W, and Au elements. [ 33 ] This RMBPT approach is based on the work of Lindgren on Rayleigh–Schrödinger perturbation theory extended to a multi-configurational zero-order wave function, [ 34 ] which is also implemented in the FAC code. [ 26 , 27 ] The RMBPT calculation dramatically removes the existing systematic deviations for wavelengths in previous theoretical calculations and reproduces these measured wavelengths to within 0.2%, which greatly improves the existing theoretical accuracy mentioned above, though still fall outside the experimental uncertainties, 0.03%, remotely.

To date, another state-of-the-art method with high accuracy is the multi-configuration Dirac–Hartree–Fock (MCDHF) approach implemented in the GRASP2K-V3 package, [ 35 ] which is a revised version of GRASP2K by Jönsson et al. [ 37 ] This approach is suitable to study different physical effects step by step. [ 37 ] In this work, with the purpose of analyzing the correlation effects clearly and getting the calculated transition energies close to the experimental data, by means of the MCDHF approach, we focus on the forbidden transition within the relatively simple multi-valence cobalt-like ions, which have the ground configuration 3d 9 , with nine valence electrons, i.e., one-vacancy-electron outside the closed core subshell. The one-vacancy-electron system is similar to the one-valence-electron system, [ 38 , 39 ] but should involve more complex correlations.

Furthermore, to investigate the core–valence (CV) correlation effects clearly, two independent approaches, the separate core–valence approach (SCV) and the full core–valence approach (FCV), are adopted. By comparison and analysis, we prove that the SCV is beneficial to give reliable information about important correlation contribution from separate subshell quickly. It is found that the electron correlation from the inner core 2p makes significant contribution to the transition energy, even larger than from the outer electron 3p. At the same time, the contribution from deeper core electron 2s is larger than from the 3s, which also should be considered. The final calculated fine-structure separation are in excellent agreement with the experimental results to within 0.01%.

2. Theoretical approach

In the MCDHF method, the ion is represented by the atomic state functions (ASF) Ψ ( γJ ), which is a linear combination of a number of configuration state functions (CSF)

where the CSFs Φ ( γ i J ) are constructed from single-electron Dirac orbitals, which are based on the well-known rules of symmetry (parity and angular momenta). [ 40 ] The coefficients c i and the radial parts in the CSFs are determined by solving the MCDHF equations which are derived by using the variational approach, within the relativistic self-consistent field (RSCF) method. [ 41 ] The radial functions are numerically represented on a logarithmic grid and are orthogonalized by introducing the Lagrange multipliers. [ 41 , 42 ] In the variational part of the calculations, we include the Dirac–Coulomb (DC) Hamiltonian (in a.u.), which can be written as

Once we obtain a set of radial orbitals, relativistic configuration interaction calculations (RCI) [ 43 ] can be performed. In the RCI calculations we include the transverse photon interaction in the Hamiltonian [ 44 ]

The photon frequency ω i j used by the RCI program in calculating the matrix elements of the transverse photon interaction is taken to be the difference in the diagonal Lagrange multipliers and associated with orbitals. When using the low frequency limit ω i j → 0, the interaction is usually referred to as the Breit interaction. The calculation of the quantum electrodynamic corrections (QED) due to the other leading part of self-energy (SE) and vacuum polarization (VP) are also included in the RCI calculation. [ 45 ]

One main advantage of the multiconfigurational method is the fact that correlation can be included in a systematic way, by using the notion of an active set (AS) of orbitals, that are in turn used to generate CSFs. [ 46 ] The foundation of this approach is a restricted active space (RAS) method, [ 47 ] where the active set of orbits is systematically increased. Considering the relative importance of different excitations from the view point of order-by-order expansion of energy in perturbation theory, we allow single (S) and double (D) replacement from the reference configurations to an active set. According to this approach, the atomic electrons are divided into two types: valence electrons and core electrons. The correlations are then classified into three types, i.e., the valence–valence (VV), core–valence (CV), and core–core (CC) correlations. It should be mentioned that the CC correlation can be further classified according to (a) intra-CC, i.e., correlation within one core subshell; and (b) inter-CC, i.e., correlation between two core subshells. The details of this approach can be found elsewhere. [ 37 , 48 ]

3. Calculation procedure
3.1. DHF calculation

At first, the DHF calculation is performed for the single 3d 9 configuration. After a set of radial orbitals have been obtained, the RCI calculations can been done, by which the relativistic corrections are included. Table  1 shows the specific contributions to the transition energy (Δ E = E ( 2 D 3/2 ) – E ( 2 D 5/2 )) between the fine-structure levels of 3d 9 2 D 3/2 and 2 D 5/2 from the DHF calculation. As expected, the DC contribution is the dominating part, and the Breit interaction giving a significant correction, decreasing the fine-structure splitting by about 2%. Among the QED effects, the SE is more important and increased by about 0.17%, whereas the VP could almost be ignored. In the following calculations, all results are obtained from the RCI calculations.

Table 1.

Relativistic and QED effects on the fine-structure splitting (Δ E ) (in cm –1 ) between the 3d 9 2 D 3/2 and 2 D 5/2 levels from the DHF calculations.

.
3.2. VV correlation calculation

On the base of the DHF calculation, the contributions of the different correlations are obtained by way of increasing the active sets (AS) layer by layer. In this VV model, we set the Ar-like 1s 2 2s 2 ··· 3p 6 core inactive and only allow the outside 3d valence electron to be excited to the AS by SD replacement. Then, the AS is increased as follows:

We note that the main VV contribution comes from n = 4 complex, and the contributions from n = 7 and 8 vary very little. The effects of high- l ( h and i ) orbitals on the fine-structure splitting have also been tested and found to be negligible (6 to 10 cm –1 ). Hence we conclude that the AS 4 mentioned above is enough to reach convergence (see Fig.  2 ). The splitting energies for the four ions of Hf 45+ , Ta 46+ , W 47+ , and Au 52+ from the VV model, that is Δ E VV , are given in Table  2 , and also in Tables  3 and 5 for further comparison.

Table 2.

The fine-structure splitting (Δ E SCV ( nl )) and the net core–valence contributions ( δE SCV ( nl ), see text) from each core orbital ( nl ) from the SCV model for the Hf 45+ , Ta 46+ , W 47+ , Au 52+ ions along with AS n . The Δ E VV from VV model given for comparisons. The total CV contributions from SCV are also given. Values are given in cm –1 .

.
Table 3.

The fine-structure splitting (Δ E FCV ( nl )) and the net core–valence contributions ( δE FCV ( nl ), see text) from each core orbital ( nl ) from the FCV model for the Hf 45+ , Ta 46+ , W 47+ , Au 52+ ions along with AS n . The Δ E VV obtained from VV model is given for comparison. The total CV contributions of FCV are also listed. Values are given in cm –1 .

.
Table 4.

The fine-structure splitting (Δ E SCC ( nl )) and the net core–core correlations ( δE SCC ( nl ), see text) from each core orbital ( nl ) from the SCC model for the Hf 45+ , Ta 46+ , W 47+ , Au 52+ ions along with AS n . The total CC contributions from SCC are also given. Values are given in cm –1 .

.
Table 5.

The fine-structure splitting (Δ E , in cm –1 ) calculated from the present DHF, VV, SCV, FCV, FCV+SCC model, compared with the other available theoretical results [ 24 , 25 , 28 , 33 ] and experimental data with error bars for Hf 45+ , Ta 46+ , W 47+ , Au 52+ ions.

.
3.3. CV correlation calculation

In the subsequent calculations, the CV correlation is successively included. We applied two strategies in this work to explore the importance of CV correlations, i.e., separate core valence (SCV) (see Section 3.3.1) and full core valence (FCV) (see Section 3.3.2). The SCV method can be used to probe the influence of one subshell at the time, on the fine-structure of the ground term. The FCV method is the final large-scale CV calculation by including full-core-valence correlation by SD excitation from all the involved core subshells simultaneously, which can be used to check the degree of approximation of SCV and the accuracy of the whole calculations.

3.3.1. Separate core–valence correlation

The separate approach for the CV correlation (SCV) was first proposed for studying Ag-like W 27+ ion by Fei. [ 37 ] In this method, the effect on the fine structure splitting is investigated by including correlation with one core subshell at a time. For example, CV with a 2p electron is included via the CSFs of the form 1s 2 2s 2 2p 6 3s 2 3p 6 3d 7 nln l ′ and 1s 2 2s 2 2p 5 3s 2 3p 6 3d 8 nln l ′ (this model also includes VV correlation), and the nl and n l ′ belong to the AS mentioned above (see Section 3.2).

In Table  2 , we present the fine-structure splitting (Δ E SCV ( nl )) for each subshell ( nl ) from the SCV model for the four target ions, along with the net CV contributions δE SCV ( nl ) ( δE SCV ( nl ) = Δ E SCV ( nl ) – Δ E VV ). To clearly show the convergence behavior for the CV correlation as a function of the AS, we plot the δE SCV ( nl ) in Fig.  1 . It is clear that δE SCV ( nl ) converges as the active set increases. It shows that except for the outer-core-3p correlation, all the others increase the fine-structure energy separations. The contributions from deep-core 2p are larger than from the outer-core 3p (by a factor of 2–3), while the contributions from deeper-core 2s are larger than from 3s (by 2 times). The CV correlations with the p subshells are considerably larger than with the s subshells. Hence, it is important to include core correlation down to the n = 2 shells to account for all important correlation effects.

Fig. 1. Convergence plot for each core (3p, 3s, 2p, 2s, 1s) CV correlation and the total CV correlation to the transition energy for the four (a) Hf 45+ , (b) Ta 46+ , (c) W 47+ , (d) Au 52+ ions in the SCV model. The dashed lines with down triangles (▽), circles (○), up triangles (△), crosses (+) and stars (★) represent the correlation from 3p, 3s, 2p, 2s, 1s in order, while the solid line with dots (●) are the sum of these CV correlation from the SCV model.

The sum of all δE SCV ( nl ) is the total CV correlation contributions resulting from the SCV model, which are also given in Table  2 and plotted in Fig.  1 . It shows that good convergence is obtained using the CSFs-AS 4 expansion, with difference between the AS 3 and AS 4 splitting being only about 40 cm –1 (0.008%). The sum of the net CV contribution and the corresponding Δ E VV , both obtained in the AS 4 calculations, referred to as Δ E SCV , can be used as a preliminary estimation on the fine-structure splitting from the SCV model, which are also given in Table  5 .

It is worth noting that, this surprising result also occurs in Ag-like tungsten ion, [ 37 ] where correlation effects from deep core 3d is larger than the outer valence electron 4d. Similarly, the 3d gives a positive contribution, but 4d presents a negative effect to the fine structure between the ground state 4d 10 4f 2 F 5/2 and 2 F 7/2 . The reason behind this is worth further study.

3.3.2. Full core–valence correlation

The second approach for the CV correlation study is the full core–valence (FCV) approach, that means correlations would be included in one step for all the involved core electrons, simultaneously. However, for comparison with the SCV model and to see the detailed correlation effects from different core electrons, we include the core electron one-by-one in the order from the outer to inner subshell (i.e., 3p→3s→2p→2s→1s). In practical terms, the core-electron 3p first interacts with valence electrons of active sets, then the inner core electrons of 3s, 2p, 2s, 1s are gradually added one-by-one. Hence, once the innermost core 1s is considered, the total CV correlation effect is obtained. Here the number of CSFs reaches 238277, which is much larger than the number from SCV for 1s (53731).

Similar to the results listed in Table  2 from the SCV model, the transition energies from the FCV approach (Δ E FCV ) are also presented in Table  3 , together given are also the net CV contributions δE FCV ( nl ) from each core-subshell. Obviously, due to different calculation procedures from the SCV, there is a difference in the way in which we obtain the δE FCV ( nl ), that is, the net CV effect from 3p is the energy difference relative to the VV model, i.e., δE FCV (3p) = Δ E FCV (3p) – Δ E VV , while the 3s CV contribution is relative to Δ E FCV (3p) ( δE FCV (3s) = Δ E FCV (3s) – Δ E FCV (3p)), and so on.

By comparing with Tables  2 and 3 , we can obtain similar conclusions for the correlation effects on the splitting from the FCV approach as well as from SCV model. The CV contributions from each core orbital calculated in the FCV model are in excellent agreement with the SCV ones. Therefore, to make an exploratory investigation of core–valence effects quickly, the SCV calculation is a good estimate, while to show all CV correlation results at one time, the FCV is a good choice.

3.4. CC correlation calculation

In this step, the CC correlation could be taken into account by applying a similar approach to that for the SCV and FCV, which are referred to as the SCC and FCC model. However, when including all the core–core correlation, the size of the CSF expansions increases rapidly with the AS, and therefore convergence is hard to achieve.

Here we only investigate the intra-CC effects [ 37 ] in the SCC model, and the resulting CC correlation contributions (Δ E SCC ( nl )) are presented in Table  4 . The net CC contribution ( δE SCC ( nl )) from each core subshell nl is the corresponding differences between Δ E SCC ( nl ) and Δ E SCV ( nl ), which are also given in Table  4 . From Table  4 , it is easy to see that the largest contribution due to the CC correlation is from the 3p orbital and the second one is from 2p, though the contributions from both of them are tiny. The sum of all the net CC correlation only decreases the transition energies around 15 cm –1 (by 30 per million) for the four ions. These net CC values are added to the results from FCV calculations, the resulting energies referred to as Δ E FCV+SCC and given in Table  5 , could be regarded as the most accurate fine-structure splitting in the present work, as all kinds of electron correlation have been taken into account.

3.5. Results and discussion

Convergence is one of the most important factors to state the accuracy of theoretical results. Figure  1 shows the convergence of the results calculated in the SCV model. We also display the convergence behavior of energy differences (Δ E Theo – Δ E Expt ) as a function of AS n for the four ions from the ab inito results of the VV and FCV model (see Fig.  2 ). It shows that the VV model is much easier to reach convergency than the FCV model. That is, the VV model converges when the AS grows to AS 2 , while the results converge in AS 4 for the FCV model. It is clear that the CV correlations are critical, increasing significantly the splitting and bringing the calculated results within the experimental uncertainties.

Fig. 2. Convergence plot of the different approaches (FCV and VV) of the fine-structure splitting compared to the experimental data for (a) Hf 45+ , (b) Ta 46+ , (c) W 47+ , and (d) Au 52+ ion. The solid lines and the dashed lines respectively represent the differences of FCV and VV from experimental data. The horizontal dashed lines stand for the experimental errors.

In Table  5 , we collect our fine-structure splitting energies obtained from the above-mentioned DHF, VV, SCV, FCV, and FCV+SCC calculations, together with previously published theoretical and experimental data. [ 24 , 25 , 28 , 33 ] The splitting energies of VV, SCV, FCV, and FCV+SCC are all calculated using the AS 4 -CSFs expansion. The Δ E VV added by the total CV contribution from each subshell in the SCV model is the Δ E SCV , and the Δ E FCV is just the Δ E FCV (1s). The final FCV+SCC results are obtained by adding the total CC correlation contributions from the SCC model to Δ E FCV (1s), as mentioned above.

By analyzing Table  5 , we can see, firstly, the VV-correlation contributions (Δ E VV – Δ E DHF ) are about 384 cm –1 and 557 cm –1 , while the CV-correlation contributions (Δ E SCV – Δ E VV ) are between 685 cm –1 and 985 cm –1 for the four Z = 72, 73, 74, 79 ions. Thus, based on the iso-electronic sequence law, we may conclude that the CV correlations are larger than the VV ones for these Co-like high- Z (70 < Z < 80) ions. Secondly, the results from SCV, FCV, and FCV+SCC agree with each other very well, and all of them agree with the measured ones to within the experimental uncertainty (0.03%), which are much better than the previous RCI, [ 24 , 25 ] (up to 0.4%) RMBPT (within 0.1%) [ 33 ] using the FAC code, [ 26 , 27 ] and MCDF results (0.6%). [ 28 ] It also implies that the CC correlations are small and could be discarded.

4. Conclusions

With the revised GRASP2K-V3 code, we have made a detailed treatment of the electron correlation for the multi-valence-electron systems of Co-like (3d 9 ) ions using the MCDHF approach. Highly accurate fine-structure energies between 2 D 3/2 and 2 D 5/2 are obtained for Hf 45+ , Ta 46+ , W 47+ , and Au 52+ , being within the experimental uncertainties. [ 24 , 25 ] It shows that it is critical to include the deep core-valence correlation down to the n = 2 subshell, to achieve spectroscopic accuracy. In particular, the CV correlation from the deep core 2p is larger than 3p, 2s is larger than 3s, and CV from 1s could be ignored. The CC correction is so small that it could be ignored in the latter large-scale calculations especially for other higher-Z Co-like ions.

Among the SCV and FCV methods used to study CV correlation, the SCV method is reliable enough to get similar accurate results to the FCV model, but with much smaller amount of CSFs and less computation efforts. The SCV method provides a simple and wise way to investigate the core–valence correlation from the individual subshell. The strategies applied here have been successfully employed to investigate the fine-structure splitting within the ground state along with the whole Co I isoelectronic sequence, as well as for the other one-valence sequence (K I and Rb I) and one-vacancy sequence (Rh I). These results are being prepared to be published elsewhere. We believe that the present strategies can also be potentially useful in even more complicated sequences, such as the n d k ( k = 2,...,8) ones.

Reference
1 Lundqvist M Nilsson H Wahlgren G M Lundberg H Xu H L Jang Z K Leckrone D S 2006 Astron. Astrophys. 450 407
2 Fischer U Pereslavtsev P Möslang A Rieth M 2009 J. Nucl. Mater 386 789
3 Seely J F Ekberg J O Brown C M Feldman U Behring W E Reader J Richardson M C 1986 Phys. Rev. Lett. 57 2924
4 Li B Higashiguchi T Otsuka T Yugami N Dunne P Kilbane D Sokell E O’ullivan G 2014 J. Phys. B: At., Mol. Opt. Phys. 47 075001
5 Fan J Z Zhang D H Chang Z W Shi Y L Dong C Z 2012 Chin. Phys. Lett 29 073102
6 Eidelsberg M Crifo-Magnant F Zeippen C J 1981 Astro. Astrophys. Suppl. Ser. 43 455
7 Edlén B 1984 Phys. Scr. 5 11
8 Kaufman V Sugar J 1986 J. Phys. Chem. Ref. Data 15 321
9 Biémont E Zeippen C J 1996 Phys. Scr. 1996 192
10 Feldman U Seely J Landi E Ralchenko Y 2008 Nucl. Fusion 48 045004
11 Fan J Z Wang Q M Chang Z W Dong C Z 2012 Chin. Phys. B 21 063102
12 Brown G V Hansen S B Träbert E Beiersdorfer P Widmann K Chen H Chung H K Clementson J H T Gu M F Thorn D B 2008 Phys. Rev. E 77 066406
13 Ralchenko Y Draganic I N Tan J N Gillaspy J D Pomeroy J M Reader J Feldman U Holland G E 2008 J. Phys. B 41 021003
14 Kramida A E Shirai T 2009 At. Data Nucl. Data Tables 95 305
15 Träbert E Clementson J Beiersdorfer P Santana J A Ishikawa Y 2010 Phys. Rev. A 82 062519
16 Kramida A 2011 Can. J. Phys. 89 551
17 Safronova U I Safronova A S 2012 J. Phys. B: At. Mol. Opt. Phys. 45 185002
18 Kilbane D O’ullivan G Gillaspy J D Ralchenko Y Reader J 2012 Phys. Rev. 86 042503
19 Qiu M L Zhao R F Guo X L Zhao Z Z Li W X Du S Y Xiao J Yao K Chen C Y Hutton R Zou Y M 2014 J. Phys. B: At., Mol. Opt. Phys. 47 175002
20 Zhao R Grumer J Li W Xiao J Brage T Huldt S Hutton R Zou Y 2014 J. Phys. B: At., Mol. Opt. Phys. 47 185004
21 Qiu M L Li W X Zhao Z Z Yang Y Xiao J Brage T Hutton R Zou Y 2015 J. Phys. B: At., Mol. Opt. Phys. 48 144029
22 Zhao Z Z Qiu M L Zhao R F Li W X Guo X L Xiao J Chen C Y Zou Y M Hutton R 2015 J. Phys. B: At., Mol. Opt. Phys. 48 115004
23 Artemyev A N Shabaev V M Tupitsyn I I Plunien G Yerokhin V A 2007 Phys. Rev. Lett. 98 173004
24 Ralchenko Y Draganić I N Osin D Gillaspy J D Reader 2011 Phys. Rev. A 83 032517
25 Osin D Gillaspy J Reader J Ralchenko Y 2012 Eur. Phys. J. D 66 286
26 Gu M F 2003 Astrophys. J. 582 1241
27 Gu M F 2008 Can. J. Phys. 86 675
28 Quinet P 2011 J. Phys. B: At. Mol. Opt. Phys. 44 195007
29 http://www.am.qub.ac.uk/darc/
30 Fournier K B 1998 At. Data Nucl. Data Tables 68 1
31 Klapisch M 1971 Comput. Phys. Commun. 2 239
32 Fischer C F Gaigalas G 2012 Phys. Rev. A 85 042501
33 Guo X L Huang M Yan J Li S Si R Li C Y Chen C Y Wang Y S Zou Y M 2015 J. Phys. B: At. Mol. Opt. Phys 48 144020
34 Lindgren I 1974 J. Phys. B: At. Mol. Opt. Phys. 7 2441
35 Jönsson P Gaigalas G Bieroń J Fischer C F Grant I 2013 Comput. Phys. Commun. 184 2197
36 Jönsson P He X Fischer C F Grant I 2007 Comput. Phys. Comm. 177 597
37 Fei Z Zhao R Shi Z Xiao J Qiu M Grumer J Andersson M Brage T Hutton R Zou Y 2012 Phys. Rev. A 86 062501
38 Brenner G López-Urrutia J R C Harman Z Mokler P H Ullrich J 2007 Phys. Rev. A 75 032504
39 Hao L H Jiang G Hou H J 2010 Phys. Rev. A 81 022502
40 Grant I P McKenzie B J Norrington P H Mayers D F Pyper N C 1980 Comput. Phys. Commun. 21 207
41 Grant I P 2007 Relativistic Quantum Theory of Atoms and Molecules New York Springer 13 10.1007%2F978-0-387-35069-1
42 Dyall K 1989 Comput. Phys. Commun. 55 425
43 McKenzie B J Grant I P Norrington P H 1980 Comput. Phys. Commun 21 233
44 Ekman J Jönsson P Gustafsson S Hartman H Gaigalas G Godefroid M R Fischer C F 2014 Astron. Astrophys. 564 24
45 Mohr P J 1982 Phys. Rev. A 26 2338
46 Olsen J Roos B O Jøgensen P Jensen H J A 1988 The Journal of Chemical Physics 89 2185
47 Brage T Fischer C F 1993 Phys. Scr. 47 18
48 Ding X B Koike F Murakami I Kato D Sakaue H A Dong C Z Nakamura NKA Sakoda J 2011 J. Phys. B: At. Mol. Opt. Phys. 44 145004