Next Article in Journal
No Cooperative Lamb Shift in Response of Thin Slab to cw Beam of Resonant Light
Next Article in Special Issue
Probing the Atomic Structure of Californium by Resonance Ionization Spectroscopy
Previous Article in Journal
Superstructure and Distorted-Wave Codes and Their Applications
Previous Article in Special Issue
Update of Atomic Data for the First Three Spectra of Actinium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Structure of Lr+ (Z = 103) from Ab Initio Calculations

by
Harry Ramanantoanina
1,2,*,
Anastasia Borschevsky
3,
Michael Block
1,2,4 and
Mustapha Laatiaoui
1,2
1
Department Chemie, Johannes Gutenberg-Universität, Fritz-Strassmann Weg 2, 55128 Mainz, Germany
2
Helmholtz-Institut Mainz, Staudingerweg 18, 55128 Mainz, Germany
3
Van Swinderen Institute for Particle Physics and Gravity, University of Groningen, Nijenborgh 4, 9747 Groningen, The Netherlands
4
GSI Helmholtzzentrum für Schwerionenforschung, Planckstrasse 1, 64291 Darmstadt, Germany
*
Author to whom correspondence should be addressed.
Submission received: 14 April 2022 / Revised: 5 May 2022 / Accepted: 6 May 2022 / Published: 9 May 2022
(This article belongs to the Special Issue Atomic Structure of the Heaviest Elements)

Abstract

:
The four-component relativistic Dirac–Coulomb Hamiltonian and the multireference configuration interaction (MRCI) model were used to provide the reliable energy levels and spectroscopic properties of the Lr + ion and the Lu + homolog. The energy spectrum of Lr + is very similar to that of the Lu + homolog, with the multiplet manifold of the 7s 2 , 6d 1 7s 1 and 7s 1 7p 1 configurations as the ground and low-lying excited states. The results are discussed in light of earlier findings utilizing different theoretical models. Overall, the MRCI model can reliably predict the energy levels and properties and bring new insight into experiments with superheavy ions.

1. Introduction

A new development in the field of atomic spectroscopy and ion mobility has been recently proposed under the name of Laser Resonance Chromatography (LRC) [1], a method that gained interest in particular because of its potential applicability to superheavy elements. In this method, optical resonances are identified based on resonant optical pumping of ions drifting in diluted helium [1,2]. The optical pumping process exploits strong ground state transitions to feed metastable electronic states, causing relative changes in the transport properties, which can be measured using drift time spectrometers [2]. However, in the perspective of an application of the LRC method in the field of superheavy elements, the question of how well optical lines are defined becomes important, because atomic levels are simply missing from conventional tables. In this context, theoretical models play a significant role in calculating the electronic structure and predicting energies. Additionally, calculations of the transport properties involving the interaction between metal ions and rare gas elements are very useful in assessing experimental parameters such as the required detector sensitivities and beamtimes [3,4].
High-accuracy theoretical predictions for the heaviest elements should be based on atomic calculations involving relativistic methods and the many-body theory. For spectra, these problems are often solved by using the Fock Space Coupled Cluster (FSCC) [5,6,7], configuration Interaction (CI) models based on multiconfigurational Dirac–Hartree–Fock (MCDHF) [8,9,10,11], a combination of CI and the many-body perturbation theory (CI+MBPT) [12,13], or multireference (MRCI) theory [14,15,16]. FSCC is one of the most powerful available approaches that provides very accurate results at a reasonable computational price, where applicable. The limitation of FSCC is in its formulation which, until recently, could only accommodate up to two holes or two electrons. Lately, this method has been extended to treat three valence electrons [17]. The use of MRCI techniques provides flexibility, allowing investigations of various configurations, and considerable effort has been invested in making the CI algorithm functional within realistic computational resources. The MRCI results for heavy and superheavy elements can be found in the literature [14,15,16,18,19,20].
In this work, calculations of the electronic structures and the properties of Lu + and Lr + ions using the MRCI model are reported. These results are compared with the experimental data [21,22] and earlier FSCC theoretical findings for the energy levels and earlier CI+MBPT data for the electronic transition rates [23]. These comparisons are used to evaluate the reliability of the calculated energy levels and to gain insight into the prospects of using this method to further study superheavy ions with multiple valence electrons (more than two) and also to evaluate the molecular systems of metal ions and rare gas elements.

2. Theoretical Method

The calculations were carried out using the 2019 release of the DIRAC code [24,25]. The electronic structure and wavefunctions were computed based on the four-component Dirac–Coulomb Hamiltonian to ensure full relativistic treatment of the superheavy elements. The nuclei were described within the finite-nucleus model in the form of a Gaussian charge distribution [26]. The Dyall basis set series of double- (cv2z), triple- (cv3z) and quadruple-zeta (cv4z) cardinal numbers for both the Lu and Lr elements [27,28] were used. All the properties were computed with these basis sets, thus allowing us to also extrapolate the energy levels at the complete basis set (CBS) limit. The small component wavefunctions were generated from the large component basis sets by strict kinetic balance [29]. Further augmentation of the basis sets with extra diffuse functions in an even-tempered manner would not significantly impact the results, as found in preliminary tests made with single- and double-augmented calculations.
We divided the theoretical procedure into three steps. The first step consisted of the atomic Dirac–Hartree–Fock (DHF) calculations that were conducted by using the average of configuration (AOC) method [30]. We considered the AOC-type calculation for consistency with an earlier study of Rf + ions [31]. We also note that based on the AOC electronic structure, we always obtained reliable energy levels and transitions in heavy metal ions and their molecular complexes [32,33,34,35]. The AOC method was used to distribute the two valence electrons of the Lu + and Lr + ions within 12 valence spinors of the s and d atomic characters. In other words, we used fractional occupation numbers (0.1667 = 2/12) for the merged Lu 6s and 5d orbitals as well as for the Lr 7s and 6d orbitals, allowing us to obtain a totally symmetrical wavefunction that was isomorphic with the configuration system under which the MRCI model (see below) was operated.
The second step consisted of the MRCI calculations that were conducted based on the AOC DHF wavefunctions. The MRCI calculations were performed by using the Kramers-restricted configuration interaction module in the DIRAC code [24]. Table 1 shows the theoretical scheme for the generalized active space (GAS) [15,18,19] that was defined in the model. In total, 34 electrons were activated that formed the basis of the valence 5d, 6s and 6p spinors of Lu (and similarly, 6d, 7s and 7p of Lr) and the semi-core 4d, 5s, 5p and 4f spinors of Lu (and similarly, 5d, 6s, 6p and 5f of Lr). No excitations were allowed in GAS 1 in order to reduce the computational demands, whereas single- and double-electron excitations were allowed in GAS 2 and GAS 3, respectively, to complete the CI expansion (see Table 1). Virtual spinors with energies below 30 atomic units were also added in the CI expansion. The numbers of the requested roots in the MRCI calculations were adjusted to contain all the multiplet manifolds of the Lu (and Lr) 6s 2 (7s 2 ), 5d 1 6s 1 (6d 1 7s 1 ) and 6s 1 6p 1 (7s 1 7p 1 ) configurations. In order to correct the energy levels for the Breit (transverse photon interaction) and the lowest-order quantum electrodynamics (QED) contributions (vacuum polarization and the self-energy terms) [36,37], we used the GRASP program package [38], which is based on the Dirac–Coulomb–Breit Hamiltonian and the multiconfiguration Dirac–Hartree–Fock (MCDHF) model. Aside from that, in the GRASP program [38], the self-energy terms are treated within the Welton approach, where the screening coefficients are approximated by the ratio of the Dirac wavefunction density in a small region around the nucleus to the same density obtained for hydrogenoic orbitals [39]. The reference spaces for the MCDHF calculations were the 4f 14 (5d6s6p) 2 and 5f 14 (6d7s7p) 2 multiplet manifolds of Lu + and Lr + , respectively. For Lu + , the virtual space for the CI expansion consisted of one extra spinor for each l quantum number from 0 to 4 (i.e., 7s7p6d5f5g). For Lr + on the other hand, the virtual space consisted of one extra spinor for each l quantum number from 0 to 2, together with two extra spinors for each l quantum number from 3 to 5 and the 6h function (i.e., 8s8p7d6f7f5g6g6h). The Lu core 5s and 5p and Lr 6s and 6p electrons were also correlated. For each energy level, the quantities Δ B and Δ B + Q E D were calculated, representing the differences in the MCDHF energy without and with the Breit contributions and the differences in the MCDHF energy without and with the Breit+QED contributions, respectively.
The third step consisted of the calculation of the spectroscopic properties based on the transition dipole moment between the levels. We used the relativistic transition moment operator within the MRCI method [40,41] to derive the oscillator strengths of the electronic transitions at the electric dipole ( E 1 ) level. We considered the multiplet manifolds of the Lu 6s 2 and 5d 1 6s 1 (Lr 7s 2 and 6d 1 7s 1 ) configurations as the lower levels and the multiplet manifold of the Lu 6p 1 6s 1 (Lr 7p 1 7s 1 ) configurations as the upper ones. The calculations of the Einstein coefficients and branching ratios were also conducted by following the standard equations [42].

3. Results

Table 2 lists the energies of the ground and the low-lying excited states of the Lu + ions as obtained from the MRCI calculations. We used the natural orbital occupation numbers of the CI vectors to deduce the dominant electron configuration of each electronic state. The electronic states were predominantly the multiplet manifold of the 6s 2 , 5d 1 6s 1 and 6s 1 6p 1 configurations (see Table 2). Note that the multiplets that originated from configurations 5d 2 or 5d 1 6p 1 were omitted for convenience because they were found to be higher in energy. For comparison, Table 2 also shows the reference energies that were taken from the literature (i.e., the experimental data collected within the framework of the National Institute of Standards and Technology (NIST) atomic spectra database [21] and previous calculations based on the FSCC model [23]).
The four columns that are depicted in the MRCI results section of Table 2 show the calculated energy levels for the three different basis sets and, subsequently, the energy extrapolated to the MRCI complete basis set limit ( E ( ) ). To derive ( E ( ) ), we used the polynomial ( n 3 ) complete basis set scheme [43] for the correlation energies with cardinal number n = 3 and 4 for triple and quadruple zeta, respectively. In Table 2, the calculated energy corrections for the Breit and Breit+QED contributions are also listed, together with the final values that add up the MRCI CBS limit and the energy corrections. For these final values (see Table 2), the numbers in brackets indicate the likely uncertainties due to the computational protocol in the least significant digits of the energy values. The uncertainties consist of the absolute value of the difference in energy between the data obtained with the triple-zeta and quadruple-zeta basis sets. Table 3 lists the calculated energies for the ground and low-lying excited states of Lr + ion as obtained from the MRCI calculations, together with the energy corrections due to the Breit interactions and QED. Similar to Table 2, the final values also include the likely uncertainties due to the computational protocol. The energy spectrum of Lr + was very similar to that of the Lu + homolog. The ground and low-lying excited states belonged to the multiplet manifold of configurations 7s 2 , 6d 1 7s 1 and 7s 1 7p 1 . For comparison, Table 3 also shows the earlier FSCC results [23].
Table 4 lists the spectroscopic properties obtained for the Lu + ions. The upper energy electronic states that belong to the configuration 6s 1 6p 1 decayed via the electric dipole E 1 mechanism to the lower energy states from configurations 5d 1 6s 1 and 6s 2 . To obtain the oscillator strengths, the Einstein coefficients and the branching ratios, we used the transition dipole moments obtained with the MRCI model [24,25], while we considered the extrapolated energy for the complete basis set limit in Table 2 for the Δ E between the upper and the lower energy levels. In Table 4, we also report the available experimental data for the Lu + ions for comparison [21,22].
Finally, Table 5 lists the spectroscopic properties obtained for the Lr + ions, alongside theoretical predictions that were taken from the literature (CI+MBPT model) [23]. Similar to the Lu + homolog, we show the transition rates for the multiplet manifolds of configuration 7s 1 7p 1 that decay via the electric dipole E 1 mechanism to the multiplet manifolds of the configurations 6d 1 7s 1 and 7s 2 . We observed that the calculated electronic transition rates of Lr + were slightly larger than those calculated for the Lu + homolog, and they were in good agreement with the earlier theoretical data taken from the literature [23].

4. Discussion

The Lu + and Lr + ions exhibited the same closed shell ground states and very similar energy spectra. The low-lying excited states of both the Lu + and Lr + ions belonged to the configurations 5d 1 6s 1 and 6d 1 7s 1 , respectively. The multiplet manifolds of the Lu 6s 1 6p 1 and Lr 7s 1 7p 1 configurations were higher in energy for both ions. The energy splitting in Lr + was larger than Lu + because of the larger spin–orbit interaction expected for the heavier element. We found that the Breit and QED corrections were relatively small, being to the order of 100 and 200 cm 1 for Lu + and Lr + , respectively, with energy values comparable to the Breit and QED effects calculated for analogous elements [23,31,44].
The MRCI energies of the Lu + ions were in good agreement with the experimental data [21,22] and previous theoretical findings [23] (see Table 2). The relative errors of most of the tabulated energy levels with respect to the experimental values were less than 5%. For the 5d 1 6s 1 configuration, the term with J = 2 ( 1 D) had the highest error (4.6%), making the term more susceptible to the interaction with the higher energy levels of the 5d 2 configuration. For configuration 6s 1 6p 1 , the term with J = 0 ( 3 P) had the highest uncertainty (5.5%), which might also be due to mixing with higher energy-excited electronic states.
The MRCI energies of the Lr + ions were also in good agreement with the previous theoretical findings [23] (see Table 3), with slightly larger deviations than for Lu + within the range of 2.9% to 8.7%. For the 6d 1 7s 1 configuration, the term with J = 2 ( 1 D) had the largest deviation (8.7%), a level which is more susceptible to interaction with the 6d 2 configuration which, in its multiplet manifold, possesses a term with the same symmetry (6d 2 1 S + 3 P + 1 D + 3 F + 1 G). For the 7s 1 7p 1 configuration, the terms with J = 0 and 1 ( 3 P) had the highest deviations (7.2% and 6.9%, respectively).
We note that the discrepancies from the reference values were larger for the odd parity states (Lu 6s 1 6p 1 and Lr 7s 1 7p 1 ) than those for the even parity states (Lu 5d 1 6s 1 and Lr 6d 1 7s 1 ). These might result from the choice of the reference spinors for the MRCI calculation, since the Lu 6p (as well as Lr 7p) spinors were left outside of the AOC occupation scheme. A possible way to improve the odd parity energy levels would be to build another AOC occupation scheme by changing the occupation number to two electrons in the Lu 6s and 6p (as well as Lr 7s 7p) and therefore run the MRCI calculation of the even and odd parity energy levels individually. The calculated Einstein coefficients for the Lu + inter-configurational 6s 2 ⟶ 6s 1 6p 1 transitions were in good agreement with the experimental data [21,22], where three electric dipole transitions were reported. The strongest transition corresponded to the 1 S 0 (6s 2 ) ⟶ 1 P 1 (6s 1 6p 1 ), in line with the experimental data [21,22], but we noted the slight overestimation of the MRCI results (see Table 4). The calculated Einstein coefficients for the Lr + inter-configurational 7s 2 ⟶ 7s 1 7p 1 transitions were also in agreement with the previously reported CI+MBPT values [23]. The strongest transition corresponded to the 1 S 0 (7s 2 ) ⟶ 1 P 1 (7s 1 7p 1 ), as was previously predicted [23]. The second strongest transitions corresponded to the 3 D 3 (6d 1 7s 1 ) ⟶ 3 P 2 (7s 1 7p 1 ) according to our MRCI calculation, unlike the 1 S 0 (7s 2 ) ⟶ 3 P 1 (7s 1 7p 1 ) transition predicted in [23].

5. Conclusions

In this work, we calculated the electronic energy levels and spectroscopic properties of the Lr + ions and of the homolog Lu + ions. We used a multireference model and configuration interaction approach to obtain the electronic structure and to compute the transition probabilities. The theoretical results were compared with the experimental data for the Lu + ions and previous theoretical findings for the Lr + ions. For Lu + , the results were remarkably very close to the experimental data, allowing us to translate the theoretical procedure to treat the heavier Lr + ions. For this, the calculated energy levels were also consistent later with the previous theoretical findings based on the Fock Space-coupled cluster method. We conclude that MRCI is a reliable theoretical model in computing energy levels for heavy and superheavy elements. MRCI is potentially of interest for systems with more than two valence electrons and also for the calculation of the interaction between the metal ions and rare gas atoms. The latter will be used to describe the transport properties of these ions in our next theoretical development.
Our results support the conclusions from previous theoretical work. (1) The energy spectrum of the Lr + ion was predicted to be similar to the one obtained for the Lu + homolog, and (2) both the Lr + and Lu + ions are good candidates for future Laser Resonance Chromatography experiments. In fact, their energy spectra present a case for experiments based on a metastable electronic state that is too long-lived for spectroscopy experiments. A potential LRC route consists of pumping the ground state 1 S 0 (6s 2 and 7s 2 ) to the excited state 3 P 1 (6s 1 6p 1 and 7s 1 7p 1 ), which radiatively decays to the metastable 3 D 1 (5d 1 6s 1 and 6d 1 7s 1 ) state with a sizeable branching ratio.

Author Contributions

Conceptualization, H.R.; methodology, H.R. and A.B.; software, H.R. and A.B.; validation, H.R., A.B., M.B. and M.L.; formal analysis, H.R., A.B. and M.L.; investigation, H.R.; resources, A.B. and M.B.; data curation, H.R. and A.B.; writing—original draft preparation, H.R.; writing—review and editing, A.B., M.B. and M.L.; visualization, H.R.; supervision, A.B. and M.L.; project administration, M.B. and M.L.; funding acquisition, M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This project received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Programme (Grant Agreement No. 819957).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

We gratefully acknowledge the high-performance computing (HPC) support, time and infrastructure from the Center for Information Technology of the University of Groningen (Peregrine), the Johannes Gutenberg University of Mainz (Mogon) and the HPC group of GSI. We are also indebted to the HPC-Europa3 program (host: Anastasia Borschevsky; guest: Harry Ramanantoanina) for HPC computer time (Netherlands) and for the funding of a short-term scientific mission in Groningen.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Laatiaoui, M.; Buchachenko, A.A.; Viehland, L.A. Laser Resonance Chromatography of Superheavy Elements. Phys. Rev. Lett. 2020, 125, 023002. [Google Scholar] [CrossRef] [PubMed]
  2. Laatiaoui, M.; Buchachenko, A.A.; Viehland, L.A. Exploiting transport properties for the detection of optical pumping in heavy ions. Phys. Rev. A 2020, 102, 013106. [Google Scholar] [CrossRef]
  3. Block, M.; Laatiaoui, M.; Raeder, S. Recent progress in laser spectroscopy of the actinides. Prog. Part. Nucl. Phys. 2021, 116, 103834. [Google Scholar] [CrossRef]
  4. Visentin, G.; Laatiaoui, M.; Viehland, L.A.; Buchachenko, A.A. Mobility of the Singly-Charged Lanthanide and Actinide Cations: Trends and Perspectives. Front. Chem. 2020, 8, 438. [Google Scholar] [CrossRef] [PubMed]
  5. Kaldor, U.; Eliav, E. High-Accuracy Calculations for Heavy and Super-Heavy Elements. In Advances in Quantum Chemistry; Academic Press: Cambridge, UK, 1998; Volume 31, pp. 313–336. [Google Scholar] [CrossRef]
  6. Dzuba, V.A.; Safronova, M.S.; Safronova, U.I. Atomic properties of superheavy elements No, Lr, and Rf. Phys. Rev. A 2014, 90, 012504. [Google Scholar] [CrossRef] [Green Version]
  7. Eliav, E.; Fritzsche, S.; Kaldor, U. Electronic structure theory of the superheavy elements. Nucl. Phys. A 2015, 944, 518–550. [Google Scholar] [CrossRef]
  8. Radžiūtė, L.; Gaigalas, G.; Jönsson, P.; Bieroń, J. Electric dipole moments of superheavy elements: A case study on copernicium. Phys. Rev. A 2016, 93, 062508. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, J.; Shen, X.; Wang, K.; Sang, C. Ionization potentials of the superheavy element livermorium (Z = 116). J. Chem. Phys. 2020, 152, 204303. [Google Scholar] [CrossRef]
  10. Fischer, C.F.; Senchuk, A. Numerical Procedures for Relativistic Atomic Structure Calculations. Atoms 2020, 8, 85. [Google Scholar] [CrossRef]
  11. Fritzsche, S. Level Structure and Properties of Open f-Shell Elements. Atoms 2022, 10, 7. [Google Scholar] [CrossRef]
  12. Porsev, S.G.; Safronova, M.S.; Safronova, U.I.; Dzuba, V.A.; Flambaum, V.V. Nobelium energy levels and hyperfine-structure constants. Phys. Rev. A 2018, 98, 052512. [Google Scholar] [CrossRef] [Green Version]
  13. Allehabi, S.O.; Dzuba, V.A.; Flambaum, V.V. Theoretical study of the electronic structure of hafnium (Hf,Z = 72) and rutherfordium (Rf,Z = 104) atoms and their ions: Energy levels and hyperfine-structure constants. Phys. Rev. A 2021, 104, 052811. [Google Scholar] [CrossRef]
  14. Visscher, L.; Saue, T.; Nieuwpoort, W.C.; Faegri, K.; Gropen, O. The electronic structure of the PtH molecule: Fully relativistic configuration interaction calculations of the ground and excited states. J. Chem. Phys. 1993, 99, 6704–6715. [Google Scholar] [CrossRef]
  15. Fleig, T.; Olsen, J.; Visscher, L. The generalized active space concept for the relativistic treatment of electron correlation. II. Large-scale configuration interaction implementation based on relativistic 2- and 4-spinors and its application. J. Chem. Phys. 2003, 119, 2963–2971. [Google Scholar] [CrossRef]
  16. Fleig, T. Invited review: Relativistic wave-function based electron correlation methods. Chem. Phys. 2012, 395, 2–15. [Google Scholar] [CrossRef]
  17. Skripnikov, L.V.; Oleynichenko, A.V.; Zaitsevskii, A.V.; Maison, D.E.; Barzakh, A.E. Relativistic Fock space coupled-cluster study of bismuth electronic structure to extract the Bi nuclear quadrupole moment. Phys. Rev. C 2021, 104, 034316. [Google Scholar] [CrossRef]
  18. Fleig, T.; Olsen, J.; Marian, C.M. The generalized active space concept for the relativistic treatment of electron correlation. I. Kramers-restricted two-component configuration interaction. J. Chem. Phys. 2001, 114, 4775–4790. [Google Scholar] [CrossRef]
  19. Fleig, T.; Jensen, H.J.A.; Olsen, J.; Visscher, L. The generalized active space concept for the relativistic treatment of electron correlation. III. Large-scale configuration interaction and multiconfiguration self-consistent-field four-component methods with application to UO2. J. Chem. Phys. 2006, 124, 104106. [Google Scholar] [CrossRef]
  20. Knecht, S.; Jensen, H.J.A.; Fleig, T. Large-scale parallel configuration interaction. II. Two- and four-component double-group general active space implementation with application to BiH. J. Chem. Phys. 2010, 132, 014108. [Google Scholar] [CrossRef]
  21. Kramida, A.; Ralchenko, Y.; Reader, J.; NIST ASD Team. NIST Atomic Spectra Database (Ver. 5.8); National Institute of Standards and Technology: Gaithersburg, MD, USA, 2022. Available online: https://physics.nist.gov/asd (accessed on 25 March 2022).
  22. Sansonetti, J.E.; Martin, W.C. Handbook of Basic Atomic Spectroscopic Data. J. Phys. Chem. Ref. Data 2005, 34, 1559–2259. [Google Scholar] [CrossRef] [Green Version]
  23. Kahl, E.V.; Berengut, J.C.; Laatiaoui, M.; Eliav, E.; Borschevsky, A. High-precision ab initio calculations of the spectrum of Lr+. Phys. Rev. A 2019, 100, 062505. [Google Scholar] [CrossRef] [Green Version]
  24. Saue, T.; Bast, R.; Gomes, A.S.P.; Jensen, H.J.A.; Visscher, L.; Aucar, I.A.; Di Remigio, R.; Dyall, K.G.; Eliav, E.; Fasshauer, E.; et al. The DIRAC code for relativistic molecular calculations. J. Chem. Phys. 2020, 152, 204104. [Google Scholar] [CrossRef] [PubMed]
  25. DIRAC, a Relativistic Ab Initio Electronic Structure Program, Release DIRAC19 (2019). Available online: http://www.diracprogram.org (accessed on 17 March 2022).
  26. Visscher, L.; Dyall, K. DIRAC–Fock Atomic Electronic Structure Calculations Using Different Nuclear Charge Distributions. At. Data Nucl. Data Tables 1997, 67, 207–224. [Google Scholar] [CrossRef]
  27. Dyall, K.G. Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the 5d elements Hf–Hg. Theor. Chem. Acc. 2004, 112, 403. [Google Scholar] [CrossRef]
  28. Dyall, K.G. Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the 6d elements Rf–Cn. Theor. Chem. Acc. 2011, 129, 603. [Google Scholar] [CrossRef]
  29. Stanton, R.E.; Havriliak, S. Kinetic balance: A partial solution to the problem of variational safety in Dirac calculations. J. Chem. Phys. 1984, 81, 1910–1918. [Google Scholar] [CrossRef]
  30. Slater, J.C. Quantum Theory of Atomic Structure, Vol. 1, 2; International Series of Monographs on Physics; McGraw-Hill: New York, NY, USA, 1960. [Google Scholar]
  31. Ramanantoanina, H.; Borschevsky, A.; Block, M.; Laatiaoui, M. Electronic structure of Rf+(Z = 104) from ab initio calculations. Phys. Rev. A 2021, 104, 022813. [Google Scholar] [CrossRef]
  32. Ramanantoanina, H. On the calculation of multiplet energies of three-open-shell 4f135fn6d1 electron configuration by LFDFT: Modeling the optical spectra of 4f core-electron excitation in actinide compounds. Phys. Chem. Chem. Phys. 2017, 19, 32481–32491. [Google Scholar] [CrossRef] [Green Version]
  33. Ramanantoanina, H. A DFT-based theoretical model for the calculation of spectral profiles of lanthanide M4,5-edge X-ray absorption. J. Chem. Phys. 2018, 149, 054104. [Google Scholar] [CrossRef] [Green Version]
  34. Ramanantoanina, H.; Studniarek, M.; Daffé, N.; Dreiser, J. Non-empirical calculation of X-ray magnetic circular dichroism in lanthanide compounds. Chem. Commun. 2019, 55, 2988–2991. [Google Scholar] [CrossRef] [Green Version]
  35. Tobin, J.G.; Ramanantoanina, H.; Daul, C.; Roussel, P.; Yu, S.W.; Nowak, S.; Alonso-Mori, R.; Kroll, T.; Nordlund, D.; Weng, T.C.; et al. Unoccupied electronic structure of actinide dioxides. Phys. Rev. B 2022, 105, 125129. [Google Scholar] [CrossRef]
  36. Indelicato, P.; Santos, J.P.; Boucard, S.; Desclaux, J.P. QED and relativistic corrections in superheavy elements. Eur. Phys. J. D 2007, 45, 155–170. [Google Scholar] [CrossRef]
  37. Ginges, J.S.M.; Berengut, J.C. QED radiative corrections and many-body effects in atoms: Vacuum polarization and binding energy shifts in alkali metals. J. Phys. B At. Mol. Opt. Phys. 2016, 49, 095001. [Google Scholar] [CrossRef] [Green Version]
  38. Fischer, C.F.; Gaigalas, G.; Jönsson, P.; Bieroń, J. GRASP2018—A Fortran 95 version of the General Relativistic Atomic Structure Package. Comput. Phys. Commun. 2019, 237, 184–187. [Google Scholar] [CrossRef]
  39. Lowe, J.; Chantler, C.; Grant, I. Self-energy screening approximations in multi-electron atoms. Radiat. Phys. Chem. 2013, 85, 118–123. [Google Scholar] [CrossRef]
  40. Knecht, S.R. Parallel Relativistic Multiconfiguration Methods: New Powerful Tools for Heavy-Element Electronic-Structure Studies; Heinrich-Heine-Universität: Düsseldorf, Germany, 2009. [Google Scholar]
  41. Denis, M.; Nørby, M.S.; Jensen, H.J.A.; Gomes, A.S.P.; Nayak, M.K.; Knecht, S.; Fleig, T. Theoretical study on ThF, a prospective system in search of time-reversal violation. New J. Phys. 2015, 17, 043005. [Google Scholar] [CrossRef]
  42. Cowan, R.D. The Theory of Atomic Structure and Spectra; International Series of Monographs on Physics; University of California Press, Ltd.: Berkeley, CA, USA, 1981. [Google Scholar]
  43. Helgaker, T.; Klopper, W.; Koch, H.; Noga, J. Basis-set convergence of correlated calculations on water. J. Chem. Phys. 1997, 106, 9639–9646. [Google Scholar] [CrossRef]
  44. Pašteka, L.; Eliav, E.; Borschevsky, A.; Kaldor, U.; Schwerdtfeger, P. Relativistic Coupled Cluster Calculations with Variational Quantum Electrodynamics Resolve the Discrepancy between Experiment and Theory Concerning the Electron Affinity and Ionization Potential of Gold. Phys. Rev. Lett. 2017, 118, 023002. [Google Scholar] [CrossRef]
Table 1. Specification of the generalized active space (GAS) scheme used for the calculations for the Lu + and Lr + ions (see the text for details).
Table 1. Specification of the generalized active space (GAS) scheme used for the calculations for the Lu + and Lr + ions (see the text for details).
GASAccumulated ElectronsNumber of SpinorsCharacters a
Min b Max
110 − x1010(n − 2)d
218 − y188(n − 1)s, (n − 1)p
332 − z3214(n − 2)f
4323418ns, (n − 1)d, np
53434(<30 au) c Virtual
a For Lu+ and Lr+, n = 6 and 7, respectively. b x, y and z are variables that control the electron excitation process attributed to the selective GAS. In the calculations, we defined the following: x = 0, y = 2 and z = 1, in line with the previous presentation [31]. c This includes all the virtual spinors up to an energy of 30 atomic units.
Table 2. Calculated energies (in cm 1 ) of the ground and the low-lying excited states of the Lu + ions obtained from the MRCI model using the double- (2), triple- (3) and quadruple-zeta (4) basis sets and the energy values derived at the complete basis set limit (), together with the final energy values (Final) that take into consideration the energy corrections obtained for the Breit ( Δ B ) and QED ( Δ B + Q E D ) contributions, compared with the experimental data (Exp.) and the FSCC results.
Table 2. Calculated energies (in cm 1 ) of the ground and the low-lying excited states of the Lu + ions obtained from the MRCI model using the double- (2), triple- (3) and quadruple-zeta (4) basis sets and the energy values derived at the complete basis set limit (), together with the final energy values (Final) that take into consideration the energy corrections obtained for the Breit ( Δ B ) and QED ( Δ B + Q E D ) contributions, compared with the experimental data (Exp.) and the FSCC results.
LevelsMRCICorrectionsFinalReference
Config.StateJ(2)(3)(4)() Δ B Δ B + Q E D Exp. a FSCC b
6s 2 1 S00000 000
5d 1 6s 1 3 D112,22712,21312,17212,14592−10412,041 (41)11,79612,354
212,69812,66912,62612,59891−8812,510 (43)12,43512,985
313,94613,90713,86613,83888−2413,814 (41)14,19914,702
1 D216,81716,65616,58316,53598−4416,491 (73)17,33317,892
6s 1 6p 1 3 P027,71228,00428,46228,75263−8828,664 (456)27,26427,091
128,88629,20829,64629,92364−7729,846 (438)28,50328,440
232,65033,12733,59933,89960−3633,863 (472)32,45332,294
1 P138,07138,40238,45338,484101−5138,433 (51)38,22338,464
a Taken from [21,22]. b Taken from [23].
Table 3. Calculated energies (in cm 1 ) of the ground and the low-lying excited states of the Lr + ion obtained from the MRCI model using the double- (2), triple (3) and quadruple-zeta (4) basis sets, and the energy values derived at the complete basis set limit (), together with the final energy values (Final) that take into consideration the energy corrections obtained for the Breit ( Δ B ) and QED ( Δ B + Q E D ) contributions, compared with the FSCC results.
Table 3. Calculated energies (in cm 1 ) of the ground and the low-lying excited states of the Lr + ion obtained from the MRCI model using the double- (2), triple (3) and quadruple-zeta (4) basis sets, and the energy values derived at the complete basis set limit (), together with the final energy values (Final) that take into consideration the energy corrections obtained for the Breit ( Δ B ) and QED ( Δ B + Q E D ) contributions, compared with the FSCC results.
LevelsMRCICorrectionsFinalReference
Config.StateJ(2)(3)(4)() Δ B Δ B + Q E D FSCC a
7s 2 1 S00000 00
6d 1 7s 1 3 D121,79621,76821,69621,649219−8621,563 (72)20,265
222,49422,45922,37522,320218−6122,259 (84)21,623
324,76124,72324,63324,5742115624,630 (90)26,210
1 D228,88328,72128,57028,4722303228,504 (151)31,200
7s 1 7p 1 3 P029,82530,07231,00631,600144−8131,519 (934)29,487
132,11432,36033,22233,770150−6033,710 (862)31,610
243,42843,80944,78345,4021524945,451 (974)43,513
1 P147,90848,13548,79449,2122053349,245 (659)47,819
a Taken from [23].
Table 4. Calculated Einstein coefficients A E 1 (in 1/s) and branching ratios β for the electric dipole’s allowed transitions in Lu + , obtained from the MRCI transition dipole moment matrix and the Δ E (in cm 1 ) from the complete basis set limit, compared with the reference experimental values.
Table 4. Calculated Einstein coefficients A E 1 (in 1/s) and branching ratios β for the electric dipole’s allowed transitions in Lu + , obtained from the MRCI transition dipole moment matrix and the Δ E (in cm 1 ) from the complete basis set limit, compared with the reference experimental values.
LevelsMRCIReference a
UpperLower Δ E A E 1 β A E 1 (NIST)
3 P 1 (6s 1 6p 1 ) 1 S 0 (6s 2 )29,9246.10 × 10 6 0.081.25 × 10 7
1 P 1 (6s 1 6p 1 ) 38,4743.74 × 10 8 0.904.53 × 10 8
3 P 0 (6s 1 6p 1 ) 3 D 1 (5d 1 6s 1 )16,6094.38 × 10 7 1.00
3 P 1 (6s 1 6p 1 ) 17,7791.39 × 10 7 0.19
3 P 2 (6s 1 6p 1 ) 21,7501.14 × 10 6 <0.01
1 P 1 (6s 1 6p 1 ) 26,3294.00 × 10 5 <0.01
3 P 1 (6s 1 6p 1 ) 3 D 2 (5d 1 6s 1 )17,3265.36 × 10 7 0.729.90 × 10 6
3 P 2 (6s 1 6p 1 ) 21,2972.06 × 10 7 0.16
1 P 1 (6s 1 6p 1 ) 25,8763.48 × 10 7 0.08
3 P 2 (6s 1 6p 1 ) 3 D 3 (5d 1 6s 1 )20,0581.09 × 10 8 0.82
3 P 1 (6s 1 6p 1 ) 1 D 2 (5d 1 6s 1 )13,3851.04 × 10 6 0.01
3 P 2 (6s 1 6p 1 ) 17,3561.82 × 10 6 0.01
1 P 1 (6s 1 6p 1 ) 21,9357.95 × 10 6 0.01
a Taken from [21,22].
Table 5. Calculated Einstein coefficients A E 1 (in 1/s) and branching ratios β for the electric dipole’s allowed transitions in Lr + , obtained from the MRCI transition dipole moment matrix and the Δ E (in cm 1 ) from the complete basis set limit, compared with the reference theoretical data.
Table 5. Calculated Einstein coefficients A E 1 (in 1/s) and branching ratios β for the electric dipole’s allowed transitions in Lr + , obtained from the MRCI transition dipole moment matrix and the Δ E (in cm 1 ) from the complete basis set limit, compared with the reference theoretical data.
LevelsMRCIReference a
UpperLower Δ E A E 1 β A E 1 (CI+MBPT)
3 P 1 (7s 1 7p 1 ) 1 S 0 (7s 2 )33,7832.97 × 10 7 0.496.36 × 10 7
1 P 1 (7s 1 7p 1 ) 49,2217.93 × 10 8 0.878.34 × 10 8
3 P 0 (7s 1 7p 1 ) 3 D 1 (6d 1 7s 1 )99661.54 × 10 7 1.005.44 × 10 6
3 P 1 (7s 1 7p 1 ) 12,1346.91 × 10 6 0.112.42 × 10 6
3 P 2 (7s 1 7p 1 ) 23,7642.44 × 10 6 <0.019.41 × 10 5
1 P 1 (7s 1 7p 1 ) 27,5721.07 × 10 6 <0.011.36 × 10 6
3 P 1 (7s 1 7p 1 ) 3 D 2 (6d 1 7s 1 )11,4632.38 × 10 7 0.394.66 × 10 6
3 P 2 (7s 1 7p 1 ) 23,0934.03 × 10 7 0.179.70 × 10 6
1 P 1 (7s 1 7p 1 ) 26,9014.98 × 10 7 0.061.63 × 10 7
3 P 2 (7s 1 7p 1 ) 3 D 3 (6d 1 7s 1 )20,8391.93 × 10 8 0.813.43 × 10 7
3 P 1 (7s 1 7p 1 ) 1 D 2 (6d 1 7s 1 )53072.51 × 10 4 <0.01
3 P 2 (7s 1 7p 1 ) 16,9372.68 × 10 6 0.013.19 × 10 5
1 P 1 (7s 1 7p 1 ) 20,7456.60 × 10 7 0.071.68 × 10 7
a Theoretical values obtained by using CI plus many-body perturbation theory (MBPT) in [23].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramanantoanina, H.; Borschevsky, A.; Block, M.; Laatiaoui, M. Electronic Structure of Lr+ (Z = 103) from Ab Initio Calculations. Atoms 2022, 10, 48. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms10020048

AMA Style

Ramanantoanina H, Borschevsky A, Block M, Laatiaoui M. Electronic Structure of Lr+ (Z = 103) from Ab Initio Calculations. Atoms. 2022; 10(2):48. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms10020048

Chicago/Turabian Style

Ramanantoanina, Harry, Anastasia Borschevsky, Michael Block, and Mustapha Laatiaoui. 2022. "Electronic Structure of Lr+ (Z = 103) from Ab Initio Calculations" Atoms 10, no. 2: 48. https://0-doi-org.brum.beds.ac.uk/10.3390/atoms10020048

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