Theoretical investigation on forbidden transition properties of fine-structure splitting in 2D state for K-like ions with 26Z36
Liu Jian-Peng1, Li Cheng-Bin2, †, Zou Hong-Xin1, ‡
Interdisciplinary Center of Quantum Information, National University of Defense Technology, Changsha 410073, China
State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, Wuhan 430071, China

 

† Corresponding author. E-mail: cbli@wipm.ac.cn hxzou@nudt.edu.cn

Abstract

Excitation energies, magnetic dipole, and electric quadrupole transition probabilities of the transition in the potassium-like (K-like) sequence with are investigated by using the multi-configuration Dirac–Hartree–Fock (MCDHF) method. The contributions of the electron correlations, Breit interaction, and the leading-order quantum electrodynamic (QED) effects on the transition properties are analyzed. The present results are interested in the laboratory tokamak and the astronomical observations. Furthermore, the feasibility of these ions for the highly charged ion (HCI) clocks is discussed. Considering the wavelength of lasers and manipulation process of the atomic clocks, Cu10+ and Zn11+ are recommended as promising candidates with achievable quality factors at the 1015 level.

1. Introduction

The forbidden magnetic dipole (M1) and electric quadrupole (E2) transition properties of ions are of great importance in the diagnostic processes of the plasmas, due to the fact that the increment of populations in metastable states is caused by the low rates of collisional de-excitations in the upper states of the forbidden transitions. This is vital to infer the plasma temperature and dynamics by observing the transitions between the fine-structure levels of the ground state of HCIs in the laboratory tokamak.[1,2] The intensities of the forbidden transitions are also used in the identification of the observed spectra, which are important to acquire the abundances of the atomic elements in astronomical objects.[3,4] Since the spectral range is wide, the accurate calculations of the atomic parameters along the isoelectronic sequences are imperative. On the other hand, the HCIs are proposed as the candidates for making new high-precision atomic clocks with accuracy below 10−19, which will be useful in the applications on the time service, navigation, communication, fundamental physics, and so forth.[59] The breakthrough of the cooling and trapping techniques for the HCIs has shed light on the realization of the new ultra-precise optical frequency standards.[10] However, finding a suitable ion is still under investigation, which is compatible with the experimental feasibility, simplicity, and high index.[11]

With the development of the theoretical methods in atomic physics and the capacity of the available computers, the transition properties in different isoelectronic sequences have been widely studied for the astrophysical and laboratory interests in recent years (see Refs. [12]–[18] and the references therein). For the K-like ions, Biémont et al. have calculated the energy levels and transition properties of the 3d configuration with .[19] Similarly, Ali et al. have presented their results with in the framework of the Dirac–Fock single-configuration approximation in the late 1980’s.[20] Charro et al. performed calculations on the M1 and E2 transition probabilities of the transition in the K-like sequences with by using the relativistic quantum defect orbital method.[21] The K-like isoelectronic sequence ions, such as Zn11+, Se15+, have already been produced and studied in the laboratory plasmas.[20] The M1 transition rate of the transition in Kr17+ has been measured in two experiments using the electron beam ion trap (EBIT).[22,23] This result was not reproducible by an ab initio calculation, but combining experimental energies with the calculated transition matrix elements, the results were matching with the measured values within the 1% level.[23] In contrast to K and Ca+ of the K-like isoelectronic sequence, higher charged ions have the first two low-lying states as states. This fine-structure splitting becomes larger with increasing in atomic number (Z). Hence, these ions with the forbidden transitions of fine-structure splitting could serve as potential candidates for the HCI clocks. As has been demonstrated in Ref. [9], the monovalent (ns) (np) or divalent (np)2 ions are suitable for clock transitions. The above fine structure splitting with longer lifetime of the upper state can be considered for atomic clocks. The obvious advantages are the low degeneracy and the simplicity of their structures, with respect to other candidates of the nf valence electrons.[68,2428] Following this idea, Yu et al. carried out the calculations on various systematics for the clock transitions in the Al-like ions with . These transitions were predicted to provide the quality factors .[11] Therefore, it would be desirable to learn whether the K-like ions can be apt for the clock transition in the optical spectral region with higher quality factors.

In this work, we carry out the calculations on the M1 and E2 transition probabilities of the fine-structure splitting of the transition along the K-like sequence with employing the MCDHF method. We analyze contributions from the electron pair correlations and choose an appropriate correlation model for the calculations. Contributions from the Breit and the leading-order QED interactions are also estimated. The lifetimes, linewidth and quality factors are given.

2. Theory and computational details
2.1. MCDHF method

In the MCDHF method, the Dirac-Coulomb Hamiltonian has the form as

Here, and β are the Dirac matrices. ViN is the interaction part with the nucleus. The atomic state wave functions (ASFs) are built from the linear combination of the different configuration state wave functions (CSFs) with the same party P, the total angular momentum J and its component along the z direction MJ, which is written as

Here, Nc is the number of CSFs, ci is the mixing coefficient of the corresponding CSF, and γi is the additional appropriate labeling to the state i. CSFs are formed by products of one-electron Dirac orbitals; it can be optimized in the self-consistent field (SCF) procedure to minimize the energy eigenvalue as well as the ci. In the subsequent relativistic configuration interaction (RCI) calculations, only the ci can be optimized. The Breit interaction in the low-frequency approximation can be taken into account in the procedure, which has the form

The leading-order QED effects can be included from the vacuum polarization and self-energy interactions (details can be found in Ref. [29]). All calculations are performed by using the GRASP2K package.[30,31]

2.2. Transition parameters

The transition probabilities Aij between the upper state and the lower state can be expressed as[32]

Here, is an electromagnetic miltipole transition operator of the order L. ω represents the transition frequency. c is the light speed in atomic units. The forms of the E2 transition operators can be expressed as the Coulomb (length) gauge and Babushkin (velocity) gauge.[33] Furthermore, the line strength can be defined as

where λij is the transition wavelength in the unit Å. When using the Racah algebra to get the reduced matrix element, it has the assumption that ASFs are built from the same orthonormal radial orbital basis. Since the states of the transitions are always optimized separately to get more accurate results, this constraint is severe. This issue can be overcome by transforming the ASFs of the two states of the transitions into biorthonormal terms.[34]

Furthermore, the lifetime τi of the upper state i can be determined by

where, the notation “o” represents the transition channels from the state i to the state j,and j represents all the states below the state i.

2.3. Computational model

In order to consider the electron correlations systematically, the active space approach is applied to build the ASFs.[35] According to the perturbation theory, the dominant reference configuration state should be included in the zero-order correlation wavefunction. The set of the first-order ASFs contains all the CSFs obtained by the single and double (SD) substitution of electrons from the occupied orbitals to the virtual orbitals, which can be classified by the substitution of the specific electron pairs.[36] Furthermore, the higher-order set contains the correlations of triple, quadruple or more excitations.

In this work, {[Ne]3s23p63d} is chosen as the reference state for the two concerned states. To describe effectively the important correlations in the first step, we calculated the contributions of the individual electron pairs separately in the SCF procedure. The calculations started from Dirac–Hartree–Fock (DHF) approximation, and all occupied orbitals were optimized, but keeping frozen in the subsequent steps. The extended optimal levels (EOL) scheme was applied here. The relaxation effects were eliminated by the simultaneous optimization of the ground and excited states in our calculations. The configurations were generated by the excitations of the specific electron pairs in each step of the active space enlargement. In this manipulation, the contributions of the corresponding pair correlations were obtained.

In order to get the reliable results, contributions of 20 kinds of electron pairs are systematically included, where the active space was taken up to n = 6 and l = 5. We illustrate the contributions of the different electron-pair correlation in the Zn11+ as shown in Fig. 1. In this figure, the y axis shows the kind of electron pair being considered. For example, the label “2s2” represents two electrons in the 2s orbital. The contributions are expressed as the difference between the specific electron pair and the DHF results. One can see that the 2p3d and 3s3p pairs have considerable contributions to the transition properties. Although the 1s2 pair has negligible influence, the pair involving the core shells 1s and 2s can also have an indispensable contribution to the correlation. It is worth noting that the contributions from different electron pairs to the and have a similar pattern. Similar conclusions were also found in the other ions, which are not shown in this paper.

Fig 1. The contributions of the different electron-pair correlations in Zn11+: (a) to the excitation energy, in unit cm−1; (b) to the M1 transition probability, in unit s−1.

According to the analysis of correlations of the individual electron pairs, the most important electron correlation effects can be captured. Moreover, to describe efficiently the electron correlations in the orbital basis set, the virtual orbitals are optimized considering the specific electron correlations of these two electron pairs 2p3d and 3s3p. The configuration space is expanded by the substitutions of these two electron pairs from the reference configuration, which means at most one excitation of the electrons is allowed in the 2p, 3d or 3s, 3p orbitals, with no more than two excitations in the corresponding two orbitals. The space is enlarged up to n = 7 and l = 5, which is labeled as n. The operation of DHF calculation is the same as above and the virtual orbitals are optimized in each new added layer. Subsequently, the other first-order correlations are considered in the RCI procedure. Here, the involving electron correlations of the pairs 2s3s, 1s2p, 2s2p, 2p2, 2s3d, 3p2, 1s3p, 2s3p, 3s2, 3p3d, which have large contributions as shown in Fig. 1, are taken into calculations. The approximation is labeled as nrSD.

The higher-order correlations were not fully included in the calculations, because the number of CSFs increases rapidly, which is beyond the limitation of the current capacity of our computational facility. Hence, we use the multi-reference (MR) approach to involve the important higher-order correlations into the RCI procedure. In this approach, the dominant configurations {[Ne]3s23p63d, [Ne]3s23p43d3} in “7rSD” will be selected as the multi-reference set. It should be noted that the latter configuration is degenerate with the former for both the J = 3/2 and J = 5/2 states in the jj configuration. We performed the results by adding the higher-order correlations to the 7rSD involving the excitations from the important electron pairs 2p3d and 3s3p to the virtual orbitals with n = 4. This model is labeled as MR-rSD.

Finally, the Breit interaction and the leading QED effects were considered through RCI computation in the configuration space, which are labeled as “+B” and “+Q”, respectively.

3. Results and discussions

In order to verify the validity of the computational model, we show the calculated excitation energy and forbidden transition probabilities of Kr17+ as the functions of the computational model in Table 1. Here, . Comparing the results obtained from DHF and MR-rSD calculations, and are not quite sensitive to the electron correlations, leading to changes less than 0.6% and 1.8% on the bases of the DHF results, respectively. Although is smaller than by about 5 orders of magnitude, considerations of the electron correlations make the changes about 17% on it. Comparing the results obtained from the 7rSD and MR-rSD calculations, it can be found that the higher-order correlations cause the relative changes of about 0.07% on the excitation energy, 0.2% on , 0.5% on in the Coulomb gauge and 0.2% on in the Babushkin gauge, respectively. The differences between the DHF/DHF+X, 7rSD/7rSD+X, and MR-rSD/MR-rSD+X (X = B, Q) results show that the contribution of the leading QED effects on the transition properties is as small as about 0.15% on the excitation energy, 0.5% on , 0.4% on , and 0.7% on , respectively. Contrary to the QED effects, the Breit interaction influences the results considerably on the excitation energy about 6%, 16% on , 31% on and 25% on . It is noted that the line strength of M1 transition, , is insensitive to the electron correlations, the Breit interaction and the leading QED effects. It implies that the reliability of can be determined by the accuracy of the calculated .

Table 1.

Excitation energies in unit cm−1, M1 transition probabilities in unit s−1 with the corresponding line strength in unit a.u. and E2 transition probabilities in unit s−1 of the transition for Kr17+ as the functions of the computational model. The notations “C” and “B” represent the results of in the Coulomb and Babushkin gauge, respectively. represents the difference between the (B) and (C). The number in square brackets represents the power of 10.

.

The calculated excitation energies of Kr17+ in Table 1 using 7rSD+BQ and MR-rSD+BQ models differ from the NIST value[39] by about 0.1% and 0.03%. It is implied that, although some of the first-order correlations are neglected, the 7rSD+BQ model can still produce accurate results. The MR-rSD+BQ model can effectively capture the main contributions of the higher-order correlations. As for , our result is in agreement with the recent experimental values by Guise et al.[23] at the level of 1%, and almost equals the adjusted values in their work. It is reasonable that the adjusted values of are extrapolated by using the experimental energy values,[38] while our calculated excitation energy has a good agreement with it. In Table 1, the results of other theoretical calculations are listed. Our results on the excitation energy agree with other calculations, with the differences less than 1%. The transition rates and are also consistent with other calculations.

The results for the excitation energies and the forbidden transition probabilities of the transition in the K-like ions with are presented in tables 2 and 3. Comparing the results of different computational models, it can be found that, among the contributions to the transition properties, the Breit interaction is the most important one. The influences of the electron correlations to the excitation energies are smaller than the Breit interaction, but it is still significant. The leading-order QED effects are quite small. It is shown that the excitation energies obtained using the MR-rSD+BQ model are in good agreement with the NIST data[39] and the differences are within 0.4%. The relative contribution of QED effects is around 0.15% to the excitation energies. The transition rates and increase rapidly along the sequence, and the M1 transition is the dominant one. It can be said that for all the ions considered in this work are around 2.40 a.u. and they are insensitive to the electron correlations, Breit interaction and the leading QED effects. The electron correlations have more influences on than .

Table 2.

Excitation energies in unit cm−1 and the M1 transition probabilities in unit s−1 of the transition for the K-like ions with . For convenience, the notations “D” and “M” are used to represent the models “DHF” and “MR-rSD”, respectively. The number in square brackets represents the power of 10.

.
Table 3.

The E2 transition probabilities in unit s−1 of the transition for the K-like ions with . The same notations “D” and “M” in Table 2 are adopted. is the relative difference between and in the model “M+BQ”. The number in square brackets represents the power of 10.

.

The transition properties for the ions are summarized in Table 4. Here, the linewidth , , and ν is the transition frequency. Since the contribution of the E2 channel to the total transition rate A is quite small, comparing to the M1 channel, we neglect it in the estimation of the lifetime of the upper states. One can see that A increases rapidly along the Z, so the lifetimes (τ) of the state decrease nearly three orders of magnitude from Fe7+ to Kr17+, and the transition linewidths (Γ) increase with the same orders. According to the definition of Q, we calculated the values and found that it decreases slowly along the sequence. The Q values of all ions are larger than . For the promising candidates of HCI clocks, it would be better that the clock transition locates in the optical regime with a high Q value. Considering the manipulation process of the atomic clock experiments, the longer lifetime of the upper state of the clock transition is preferable. Although the fine-structure splitting in the 2D state of Kr17+ is 637 nm, which is in the optical region, the lifetime of the state of it is only 24 ms, which is not good enough for the clock experiment. The state of Fe7+ has the long lifetime as 15 s and the Q value is also large, but the transition wavelength is 5457 nm, which is not easy to build proper lasers for experiments. As the practical and balanced choice, Cu10+ and Zn11+ are recommended. The transitions of both ions are in the near infrared region and the lasers could be available. The lifetimes of the state are about 1 s, which is long enough for experiments.

Table 4.

The wavelength λ, total transition probabilities A, lifetimes τ, linewidth Γ, and quality factor Q of the transition for the K-like ions with . The number in square brackets represents the power of 10.

.
4. Conclusion

For the interests of the plasma diagnostic processes and the candidates of the HCI clocks, we have investigated the M1 and E2 transition properties of the fine-structure splitting in 2D state for the K-like ions for with the MCDHF method. The results of the excitation energies have good agreements with NIST data and the differences are within 0.4%. The contributions of the electron correlations, Breit interaction, and the leading QED effects are analyzed. It is shown that the good treatment of electron correlations and inclusion of the Breit interaction are essential for the reliable prediction of the transition properties for these ions. In choosing candidates for HCI clocks from the K-like ions in , Cu10+ and Zn11+ are recommended for the balanced consideration of the Q factor, the lifetime of the clock state and the feasibility of the lasers.

Reference
[1] Suckewer S Hinnov E 1978 Phys. Rev. Lett. 41 756
[2] Suckewer S Hinnov E 1979 Phys. Rev. 20 578
[3] Lodders K 2003 Astrophys. J. 591 1220
[4] François P Matteucci F Cayrel R Spite M Chiappini C 2004 Astron & Astronphys 421 613
[5] Ovsiannikov V D Derevianko A Gibble A 2011 Phys. Rev. Lett. 107 093003
[6] Safronova M S Dzuba V A Flambaum V V Safronova U I Porsev S G Kozlov M G 2014 Phys. Rev. Lett. 113 030801
[7] Derevianko A Dzuba V A Flambaum V V 2012 Phys. Rev. Lett. 109 180801
[8] Berengut J C Dzuba V A Flambaum V V 2010 Phys. Rev. Lett. 105 120801
[9] Yudin V I Taichenachev A V Derevianko A 2014 Phys. Rev. Lett. 113 233003
[10] Schmöger L Versolato O O Schwarz M Kohnen M Windberger A Piest B Feuchtenbeiner S Pedregosa-Gutierrez J Leopold T Micke P Hansen A K Baumann T M Drewsen M Ullrich J Schmidt P O López-Urrutia J R C 2015 Science 347 1233
[11] Yu Y M Sahoo B K 2016 Phys. Rev. 94 062502
[12] Rynkun P Jönsson P Gaigalas G Fischer C F 2013 Astron & Astronphys. 557 A136
[13] Jönsson L Gaigalas G Godefroid M R Marques J P Brage T Fischer C F Grant I P 2016 Astron. & Astronphys. 585 A26
[14] Rynkun P Jönsson P Gaigalas G Fischer C F 2014 At. Data Nucl. Data Tables 100 315
[15] Jönsson P Alkauskas A Gaigalas G 2013 At. Data Nucl. Data Tables 99 431
[16] Wang K Chen Z B Si R Jönsson P Ekman J Guo X L Li S Long F Y Dang W Zhao X H Hutton R Chen C Y Yan J Yang X 2016 Astro. J. Supp. 226 14
[17] Ekman J Jönsson P Gustafsson S Hartman H Gaigalas G Godefroid M R Fischer C F 2014 Astron. & Astronphys. 564 A24
[18] Rynkun P Jönsson P Gaigalas G Fischer C F 2012 At. Data Nucl. Data Tables 98 481
[19] Biémont E Hansen J E 1989 Phys. Scr. 39 308
[20] Ali M A Kim Y K 1988 Phys. Rev. 38 3992
[21] Charro E Curiel Z Martín I 2002 Astron. & Astronphys. 387 1146
[22] Träbert E Beiersdorfer P Brown G V Chen H Thorn D B Biémont E 2001 Phys. Rev. 64 042511
[23] Guise N D Tan J N Brewer S M Fischer C F Jönsson P 2014 Phys. Rev. 89 040502
[24] Nandy D K Sahoo B K 2016 Phys. Rev. 94 032504
[25] Dzuba V A Derevianko A Flambaum V V 2012 Phys. Rev. 86 054501
[26] Safronova M S Dzuba V A Flambaum V V Safronova U I Porsev S G Kozlov M G 2014 Phys. Rev. 90 042513
[27] Safronova M S Dzuba V A Flambaum V V Safronova U I Porsev S G Kozlov M G 2014 Phys. Rev. 90 052509
[28] Dzuba V A Flambaum V V Katori H 2015 Phys. Rev. 91 022119
[29] Dyall K G Grant I P Johnson C T Parpia F A Plummer E P 1989 Comput. Phys. Commun. 55 425
[30] Jönsson P He X Fischer C F Grant I P 2007 Comput. Phys. Commun. 177 597
[31] Jönsson P Gaigalas G Bieroń J Fischer C F Grant I P 2013 Comput. Phys. Commun. 184 2197
[32] Grant I P 2007 Relativistic Quantum Theory of Atoms and Molecules: Theory and Computation Springer
[33] Grant I P 1974 J. Phys. B: At. Mol. Opt. Phys. 7 1458
[34] Olsen J Godefroid M R Jönsson P Malmqvist P Å Fischer C F 1995 Phys. Rev. 52 4499
[35] Fischer C F Brage T Jönsson P 1997 Computational Atomic Structure: An MCHF Approach Bristol and Philadelphia Institute of Physics Publishing
[36] Zhou F Qu Y Li J Wang J 2015 Phys. Rev. 92 052505
[37] Ali M A Kim Y K 1992 J. Opt. Soc. Am. 9 185
[38] Kaufman V Sugar J Rowan W 1989 J. Opt. Soc. Am. 6 142
[39] A. Kramida, Yu. Ralchenko, J. Reader, NIST ADS TEAM (2016) NIST Atomic Spectra Database (version 5.4) http://physics.nist.gov/asd.
[40] López-Urrutia J R C Beiersdorfer P Widmann K Decaux V 1999 Phys Scr T80 448