This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Precision calculation of hyperfine structure of 7,9Be2+ ions

Xiao-Qiu Qi1,2    Pei-Pei Zhang2,4,†    Zong-Chao Yan3,2    Ting-Yun Shi2    G. W. F. Drake4    Ai-Xi Chen1    Zhen-Xiang Zhong2 1 Key Laboratory of Optical Field Manipulation of Zhejiang Province and Physics Department of Zhejiang Sci-Tech University, Hangzhou 310018, China 2 State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430071, China 3 Department of Physics, University of New Brunswick, Fredericton, New Brunswick, Canada E3B 5A3 4 Department of Physics, University of Windsor, Windsor, Ontario, Canada N9B 3P4
Abstract

The hyperfine structures of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of the 7Be2+ and 9Be2+ ions are investigated within the framework of the nonrelativistic quantum electrodynamics (NRQED). The uncertainties of present hyperfine splitting results of 9Be2+ are in the order of several tens of ppm, where two orders of magnitude improvement over the previous theory and experiment values has been achieved. The contribution of nuclear electric quadrupole moment to hyperfine splitting of 7Be2+ has been studied. A scheme for determining the properties of Be nuclei in terms of Zemach radius or the electric quadrupole moment based on precise spectra is proposed, and it opens a new window for the study of Be nuclei.

I Introduction

Light helium and helium-like ions are among simplest atomic systems where theoretical approaches are well advanced to calculate their electronic structures with high precision. However, there are still unsolved interesting problems [1, 2, 3, 4, 5, 6, 7]. Among various theoretical methods, the nonrelativistic quantum electrodynamics (NRQED) is the most effective approach designed to calculate the electronic structure of light atomic systems [8, 9, 10, 11]. For the helium 23PJ2\,^{3}\!P_{J} fine-structure, for example, the NRQED-based calculation has achieved a precision of about 1.7 kHz, far exceeding all other theoretical approaches that are based on Dirac-like methods [12]. Experimentally, Clausen et al. [13] have recently reported a much improved new determination of the He 21S2\,^{1}\!S ionization energy at the level of 32 kHz, which is in good accord with theory. However, the derived experimental ionization energies of the 23S2\,^{3}\!S and 23P2\,^{3}\!P states are in disagreement with theoretical prediction by 6.5σ6.5\sigma and 10σ10\sigma, respectively. Li+{\rm Li}^{+} is very similar to helium with a higher ZZ, and its QED effect is more significant than helium. For the 23P12\,^{3}\!P_{1}-23P22\,^{3}\!P_{2} fine structure interval for example, the contribution from order mα6\alpha^{6} and higher in Li+{\rm Li}^{+} is a factor of 26 larger than for helium [14]. The hyperfine structure splittings (hfs) of Li+{\rm Li}^{+} have been studied in our previous work [5] using the NRQED theory. The theoretical uncertainty is reduced to be less than 100 kHz by a complete calculation of all the corrections up to mα6\alpha^{6}. The so-called Zemach radius, which describes the distribution of magnetic moment inside the nucleus, can be extracted by combining precision measurements [14]. The obtained Zemach radius for 7Li is in good agreement with previous values, while the value for 6Li disagrees with the nuclear physics value [15] by more than 6σ6\sigma, indicating an anomalous nuclear structure for 6Li.

For further testing QED effect with low-ZZ ions, the helium-like Be2+\rm{Be}^{2+} is a suitable candidates [16, 17, 18], since the transition wavelength of 23S23P2\,^{3}\!S-2\,^{3}\!P of 372 nm is still close to the visible region. Beryllium has many isotopes 6Be-14Be [19, 20, 21, 22], including one-neutron halo 11Be and two-neutron halo 14Be. There are some recent spectral experiments to explore Be\rm{Be} nuclear structure [20, 22, 23, 24]. Puchalski et al. calculated the hyperfine splittings of 9Be using explicitly correlated Gaussian function (ECG), and accurately determined the nuclear electric quadrupole moment [25], although it was inconsistent with the previous value. The advantage of studying Be2+\rm{Be}^{2+} ion rather than neutral Be is that it is a three-body system for which the corresponding QED theory is relatively simpler. Compared with helium and Li+{\rm Li}^{+} ions, the current research on Be2+\rm{Be}^{2+} is rare. yanyan In 1993, Scholl et al. measured the 1s2s3S11s2p3PJ1s2s^{3}\!S_{1}-1s2p^{3}\!P_{J} transition of Be2+9{}^{9}\mathrm{Be}^{2+} ion by applying fast ion beam laser fluorescence method with an accuracy of 10810^{-8} [16], which is three orders of magnitude improvement over previous measurements. The fine and hyperfine splittings extracted are, respectively, in the order of tens of ppm and 10410^{-4}. Theoretically, Johnson et al. in 1997 [26] calculated 23PJ2\,^{3}\!P_{J} hfs of Be2+9{}^{9}\mathrm{Be}^{2+} by the relativistic configuration interaction method with only four significant digits. With the development of experimental technology, especially the emergence of new light sources of narrow linewidth in the XUV area [27, 28, 29], it is now possible to improve the measurement of Be2+\mathrm{Be}^{2+} to reach a new accuracy level.

In this paper, we intend to present a systematic calculation of hfs of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of the 7,9Be2+ ions by including QED corrections up to mα6m\alpha^{6} order. The possibility of determining the Zemach radius and the electric quadrupole moment of a Be isotope based on Be2+\mathrm{Be}^{2+} spectroscopy is discussed. The present paper is organized as follows. Sec. II outlines the basic theoretical framework for our calculations. Sec. III details various QED contributions to the hfs of 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of the 7,9Be2+. Finally, discussions and conclusions are given in Sec. IV.

Refer to caption
Figure 1: Hyperfine energy levels (not drawn to scale) of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of Be2+9{}^{9}\mathrm{Be}^{2+} [16], in cm-1. The nuclear spin of 9Be is 3/2.

II Theoretical method

The NRQED theory for quasidegenerate states is used to calculate fine and hyperfine structure splittings [30, 31, 4, 32]. The theory has been used in calculation of Li+ hyperfine structure [5]. Here we simply outline the framework for calculating relativistic and QED corrections to an energy level. Figure 1 shows the energy level diagram of hfs for Be2+9{}^{9}\mathrm{Be}^{2+} (The diagram for Be2+7{}^{7}\mathrm{Be}^{2+} is similar to Be2+9{}^{9}\mathrm{Be}^{2+} since they have the same nuclear spin 3/2). In order to obtain the energies of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states, we need to diagonalize the effective Hamiltonian HH with its matrix elements being

EJJFJFMF|H|JFMF,\displaystyle E_{JJ^{\prime}}^{F}\equiv\langle JFM_{F}|H|J^{\prime}FM_{F}\rangle, (1)

where MFM_{F} is the projection of the total angular momentum FF, which can be fixed arbitrarily since the energies are independent of it. For convenience, we treat the 23PJ2\,^{3}P_{J} centroid as the zero level. The above matrix elements Eq. (1) can be expanded in powers of the fine structure constant α\alpha

EJJF=\displaystyle E_{JJ^{\prime}}^{F}= HfsJδJJ+Hhfs(4+)+Hhfs(6)\displaystyle\langle H_{\mathrm{fs}}\rangle_{J}\delta_{JJ^{\prime}}+\langle H_{\mathrm{hfs}}^{(4+)}\rangle+\langle H_{\mathrm{hfs}}^{(6)}\rangle (2)
+2Hhfs(4),[Hnfs(4)+Hfs(4)]+Hhfs(4),Hhfs(4)\displaystyle+2\langle H_{\mathrm{hfs}}^{(4)},[H_{\mathrm{nfs}}^{(4)}+H_{\mathrm{fs}}^{(4)}]\rangle+\langle H_{\mathrm{hfs}}^{(4)},H_{\mathrm{hfs}}^{(4)}\rangle
+HQED(6)+HQEDho+Hnucl+Heqm,\displaystyle+\langle H_{\mathrm{QED}}^{(6)}\rangle+\langle H_{\mathrm{QED}}^{\mathrm{ho}}\rangle+\langle H_{\mathrm{nucl}}\rangle+\langle H_{\mathrm{eqm}}\rangle,

where A,BA1(E0H0)B\langle A,B\rangle\equiv\langle A\frac{1}{(E_{0}-H_{0})^{\prime}}B\rangle, with H0H_{0} and E0E_{0} being the nonrelativistic Hamiltonian and its eigenvalue. HfsH_{\mathrm{fs}} is the effective operator that does not depend on the nuclear spin and is responsible for the fine structure splittings [12, 33]. The other terms in Eq. (2) are the nuclear spin dependent contributions. Hhfs(4+)H_{\mathrm{hfs}}^{(4+)} is the leading-order hyperfine Hamiltonian of mα4m\alpha^{4}, where the superscript ‘+’ means the higher-order terms from the recoil and anomalous magnetic moment effects. Hhfs(6)H_{\mathrm{hfs}}^{(6)} is the effective operator for the hyperfine splittings of order mα6m\alpha^{6}. Hfs(4)H_{\mathrm{fs}}^{(4)} and Hnfs(4)H_{\mathrm{nfs}}^{(4)} are the Breit Hamiltonians of order mα4m\alpha^{4} with and without electron spin. The fifth term in Eq. (2) is the second-order hyperfine correction, which contributes to the isotope shift, fine and hyperfine splittings. HQED(6)H_{\mathrm{QED}}^{(6)} and HQEDhoH_{\mathrm{QED}}^{\mathrm{ho}} are the two effective operators for the QED corrections of order mα6m\alpha^{6} and higher mα7\sim m\alpha^{7}. Finally, HnuclH_{\mathrm{nucl}} and HeqmH_{\mathrm{eqm}} represent the nuclear effects due to the Zemach radius and the nuclear electric quadrupole moment.

We solve the eigenvalue problem of H0H_{0} variationally in Hylleraas coordinates. The relativistic and QED corrections as well as the corrections due to nuclear structure are evaluated perturbatively. The Hylleraas basis set [34] is constructed according to

ψmn(r1,r2)=r1r2mrneαr1βr2γr𝒴12LM(r^1,r^2),\displaystyle\psi_{\ell mn}(\vec{r}_{1},\vec{r}_{2})=r^{\ell}_{1}r^{m}_{2}r^{n}e^{-\alpha r_{1}-\beta r_{2}-\gamma r}\mathcal{Y}^{LM}_{{\ell}_{1}{\ell}_{2}}(\hat{r}_{1},\hat{r}_{2}), (3)

where r=r1r2\vec{r}=\vec{r}_{1}-\vec{r}_{2} and 𝒴12LM(r^1,r^2)\mathcal{Y}^{LM}_{{\ell}_{1}{\ell}_{2}}(\hat{r}_{1},\hat{r}_{2}) is the vector coupled product of spherical harmonics for the electrons. In order to deal with the nonrelativistic finite nuclear mass effect, according to whether the mass polarization operator is explicitly included in the nonrelativistic Hamiltonian, two different types of wave functions can be generated. For Hhfs(4+)\langle H_{\mathrm{hfs}}^{(4+)}\rangle, HQED(6)\langle H_{\mathrm{QED}}^{(6)}\rangle, HQEDho\langle H_{\mathrm{QED}}^{\mathrm{ho}}\rangle, Hnucl\langle H_{\mathrm{nucl}}\rangle, and Heqm\langle H_{\mathrm{eqm}}\rangle, we use the wave functions with the mass polarization, whereas for other terms we use the wave functions corresponding to the infinite nuclear mass limit. The coupling of intermediate states of different symmetries should be included in the second-order terms, where some singular integrals need to be handled by including more singular terms in the intermediate states [35]. The necessary angular momentum operators, which can be evaluated analytically [5], are [31] SiLiS^{i}L^{i}, IiLiI^{i}L^{i}, IiSiI^{i}S^{i}, {SiSj}{LiLj}\{S^{i}S^{j}\}\{L^{i}L^{j}\}, IiSj{LiLj}I^{i}S^{j}\{L^{i}L^{j}\}, IiLj{SiSj}I^{i}L^{j}\{S^{i}S^{j}\}, {IiIj}{SiSj}\{I^{i}I^{j}\}\{S^{i}S^{j}\}, {IiIj}{LiLj}\{I^{i}I^{j}\}\{L^{i}L^{j}\}, {IiIj}LiSj\{I^{i}I^{j}\}L^{i}S^{j}, and {IiIj}[{SmSn}{LkLl}]ij\{I^{i}I^{j}\}[\{S^{m}S^{n}\}\{L^{k}L^{l}\}]^{ij}, where {SiSj}12SiSj+12SjSi13S2δij\{S^{i}S^{j}\}\equiv\frac{1}{2}S^{i}S^{j}+\frac{1}{2}S^{j}S^{i}-\frac{1}{3}\vec{S}^{2}\delta^{ij} and the summation over the repeated indices is assumed.

III The hfs of 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states

The hfs operators responsible for relativistic and QED corrections to the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of helium-like ions are defined in our previous paper [5]. The first-order perturbation results of mα4m\alpha^{4} and mα6m\alpha^{6} corrections are listed in Table 1.

Table 1: Expectation values for the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of 7Be2+ and 9Be2+. The listed numerical values are uncertain only at the last digits. In atomic units.
State Operator Be2+7{}^{7}\mathrm{Be}^{2+} Be2+9{}^{9}\mathrm{Be}^{2+}
23S12\,^{3}\!S_{1} 4πδ3(r1)4\pi\delta^{3}(\vec{r}_{1}) 137.731960 137.739110
KK^{\prime} –157.232037 –157.232037
23PJ2\,^{3}\!P_{J} 4πδ3(r1)4\pi\delta^{3}(\vec{r}_{1}) 126.232881 126.239606
(r1×p1)/r13(\vec{r}_{1}\times\vec{p}_{1})/r_{1}^{3} 1.814143 1.814146
(r1×p2)/r13(\vec{r}_{1}\times\vec{p}_{2})/r_{1}^{3} –2.602150 –2.602181
(δij3r1ir1j/r12)/r13(\delta^{ij}-3r_{1}^{i}r_{1}^{j}/r_{1}^{2})/r_{1}^{3} –0.671762 –0.671769
KK^{\prime} 55.3670854 55.3670854
K\vec{K} –145.86034 –145.86034
K^\hat{K} –82.07829 –82.07829

The second-order corrections of mα6m\alpha^{6} can be divided into several parts according to the symmetries of the intermediate states. For the 23S12\,^{3}\!S_{1} state, the intermediate states are S3{}^{3}\!S, P3{}^{3}\!P, and D3{}^{3}\!D. And for 23PJ2\,^{3}\!P_{J} state the intermediate states are P3{}^{3}P, P1{}^{1}P, D3{}^{3}D, D1{}^{1}D, and F1{}^{1}F. Numerical results of various operators for the radial parts are presented in Table 2. Since the second-order hyperfine correction Hhfs(4),Hhfs(4)\langle H_{\mathrm{hfs}}^{(4)},H_{\mathrm{hfs}}^{(4)}\rangle is divergent, we calculate only the dominant contribution from the 21P12\,^{1}\!P_{1} intermediate state. It should be noted that the uncertainty of PA,PA\langle P_{A},P_{A}\rangle^{\circ} in Table 2 is only computational. We also use the method in Ref. [31] to estimate the uncertainty due to this approximation, i.e., calculating the second-order perturbation for the operator PA,PA\langle P^{\prime}_{A},P^{\prime}_{A}\rangle, and taking the difference between PA,PA\langle P_{A},P_{A}\rangle^{\circ} and PA,PA\langle P^{\prime}_{A},P^{\prime}_{A}\rangle as the uncertainty, which is 10000 a.u. for 23S12\,^{3}\!S_{1} and 15000 a.u. for 23PJ2\,^{3}\!P_{J} respectively.

Table 2: Second-order matrix elements for all possible intermediate states connected to the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states. The listed numerical values are uncertain only at the last digits when not given explicitly. In atomic units.
State Symmetry A,B\langle A,B\rangle Value
23S12\,^{3}\!S_{1} S3{}^{3}\!S P,G\langle P^{\prime},G^{\prime}\rangle 4592.8
P3{}^{3}\!P P,G\langle\vec{P},\vec{G}\rangle 0.140
D3{}^{3}\!D P^,G^\langle\hat{P},\hat{G}\rangle 0.87
S1{}^{1}\!S PA,PA\langle P_{A},P_{A}\rangle^{\circ} –839249.282
PA,PA\langle P^{\prime}_{A},P^{\prime}_{A}\rangle –848800(200)
23PJ2\,^{3}\!P_{J} P3{}^{3}\!P P,G\langle P^{\prime},G^{\prime}\rangle 4024.6(5)
P,G\langle\vec{P},G\rangle 86.9(5)
P^,G\langle\hat{P},G\rangle 40(5)
P,G\langle P,\vec{G}\rangle 26.678
P,G\langle\vec{P},\vec{G}\rangle –64.1
P^,G\langle\hat{P},\vec{G}\rangle –39.8427
P,G^\langle P,\hat{G}\rangle 6.714(5)
P,G^\langle\vec{P},\hat{G}\rangle 19.323
P^,G^\langle\hat{P},\hat{G}\rangle 8.882
P1{}^{1}\!P PA,GA\langle P_{A},\vec{G}_{A}\rangle 9054.88(5)
PA,GA\langle P_{A},\vec{G}_{A}\rangle^{\circ} 9021.158
P^A,GA\langle\hat{P}_{A},\vec{G}_{A}\rangle –176.7(5)
P^A,GA\langle\hat{P}_{A},\vec{G}_{A}\rangle^{\circ} –139.044
D3{}^{3}\!D P,G\langle\vec{P},\vec{G}\rangle 0.044
P^,G\langle\hat{P},\vec{G}\rangle –0.0149824
P,G^\langle\vec{P},\hat{G}\rangle 1.563(5)
P^,G^\langle\hat{P},\hat{G}\rangle –0.0688
D1{}^{1}\!D P^A,GA\langle\hat{P}_{A},\vec{G}_{A}\rangle 0.348(5)
F3{}^{3}\!F P^,G^\langle\hat{P},\hat{G}\rangle 0.628(5)
P1{}^{1}\!P PA,PA\langle P_{A},P_{A}\rangle^{\circ} –1785103.485
PA,PA\langle P^{\prime}_{A},P^{\prime}_{A}\rangle –1797300(200)
PA,P^A\langle P_{A},\hat{P}_{A}\rangle^{\circ} 27527.773
P^A,P^A\langle\hat{P}_{A},\hat{P}_{A}\rangle^{\circ} –2547.006

We calculate the hfs of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states using the values in Tables 1 and 2. Since the contribution from the 1s1s electron dominates higher-order QED correction, the assumption that HQEDho(1s2p)HQEDho(1s)H_{\mathrm{QED}}^{\mathrm{ho}}(1s2p)\simeq H_{\mathrm{QED}}^{\mathrm{ho}}(1s) is adopted for the hfs calculation of the 23PJ2\,^{3}\!P_{J} state, while the HQEDho(1s2s)H_{\mathrm{QED}}^{\mathrm{ho}}(1s2s) of 23S12\,^{3}\!S_{1} state is approximated by the weighted average of HQEDho(1s)H_{\mathrm{QED}}^{\mathrm{ho}}(1s) and HQEDho(2s)H_{\mathrm{QED}}^{\mathrm{ho}}(2s). The uncertainty of the correction HQEDhoH_{\mathrm{QED}}^{\mathrm{ho}} is estimated as 20% of its contribution. According to Eq. (2), the hfs calculation of the 23PJ2\,^{3}\!P_{J} state requires the results of the fine structure splittings, which are HfsJ=0=(8f01+5f12)/9\langle H_{fs}\rangle_{J=0}=(8f_{01}+5f_{12})/9, HfsJ=1=(f01+5f12)/9\langle H_{fs}\rangle_{J=1}=(-f_{01}+5f_{12})/9, and HfsJ=2=(f014f12)/9\langle H_{fs}\rangle_{J=2}=(-f_{01}-4f_{12})/9, relative to the 23PJ2\,^{3}\!P_{J} centroid, with f01=11.5586(5)f_{01}=11.5586(5) cm-1 and f12=14.8950(4)f_{12}=-14.8950(4) cm-1 for 9Be2+ [16]. The fine structure splittings of 7Be2+ are obtained by changing the reduced mass accordingly i.e., f01=11.558(2)f_{01}=11.558(2) cm-1 and f12=14.895(2)f_{12}=-14.895(2) cm-1. For 7Be2+ and 9Be2+, the magnetic moments are 1.39928(2)μN-1.39928(2)~{}\mu_{N} [36] and 1.177432(3)μN-1.177432(3)~{}\mu_{N} [20], and the nuclear electric quadrupole moments 6.11fm2-6.11~{}\rm{fm}^{2} (the theoretical result from  [37]) and 5.350(14)fm25.350(14)~{}\rm{fm}^{2} [25], respectively. The contributions of Zemach radii are at 615(8)-615(8) ppm [25] for 9Be2+ and 521(16)-521(16) ppm for 7Be2+. This contribution of 7Be2+ is calculated by 2ZRem/a0-2ZR_{\rm{em}}/a_{0}, where Rem=4Re/3πR_{\rm{em}}=4R_{\rm{e}}/\sqrt{3\pi} (Gaussian distributions) and the nuclear charge radius ReR_{\rm{e}} is 2.647(17) fm [20]. The hfs of 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states can be obtained by diagonalizing the matrix in Eq. (2) and the results relative to the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} centroids are listed in Table 3.

Table 3: Theoretical results for individual 23S1F2\,^{3}\!S^{F}_{1} and 23PJF2\,^{3}\!P^{F}_{J} levels in 7Be2+ and 9Be2+, relative to the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} centroid respectively, where the first error in 23P2^{3}P state is due to the fine structure and the second error is due to the hyperfine structure, in cm-1.
State (J,F)(J,F) Be2+7{}^{7}\mathrm{Be}^{2+} Be2+9{}^{9}\mathrm{Be}^{2+}
23S2^{3}S (1,1/2)(1,1/2)    0.68251(1)    0.574282(6)
(1,3/2)(1,3/2)    0.27300(1)    0.229708(3)
(1,5/2)(1,5/2) –0.40950(1) –0.344566(4)
23P2^{3}P (2,1/2)(2,1/2)    5.90767(100)(1)    5.817172(190)(4)
(2,3/2)(2,3/2)    5.72041(100)(1)    5.658805(190)(3)
(2,5/2)(2,5/2)    5.40467(100)(1)    5.392683(190)(1)
(2,7/2)(2,7/2)    4.95513(100)(1)    5.015548(190)(3)
(0,3/2)(0,3/2)    2.01174(200)(1)    2.008479(500)(1)
(1,1/2)(1,1/2) –9.23556(110)(1) –9.287087(230)(3)
(1,3/2)(1,3/2) –9.44648(110)(1) –9.461891(230)(1)
(1,5/2)(1,5/2) –9.75933(110)(1) –9.727037(230)(2)

For the 23PJ2\,^{3}\!P_{J} states, the 21P123P12\,^{1}\!P_{1}-2\,^{3}\!P_{1} mixing effect should be taken into consideration carefully. Here we follow two methods used in our previous calculation [5]. Method 1. Do an exact diagonalization only within the 23PJ2\,^{3}\!P_{J} manifold and treat the 21P123P12\,^{1}\!P_{1}-2\,^{3}\!P_{1} mixing effect by perturbation theory up to second order. Method 2. Extend the 23PJ2\,^{3}\!P_{J} manifold by including the 21P12\,^{1}\!P_{1} state and do an exact diagonalization of the extended matrix. Both the methods only include the relativistic correction of order mα4\alpha^{4}. The second-order matrix elements involving the intermediate state 21P12\,^{1}\!P_{1} and the hyperfine structure coefficients [33] for the 21P12\,^{1}\!P_{1} and 23PJ2\,^{3}\!P_{J} states are listed in Tables 2 and 4, as inputs for applying Methods 1 and 2. The hfs of 23PJ2\,^{3}\!P_{J} are evaluated using these two methods and the results are presented in Table 5. The modification of the mixing effect alters the hyperfine intervals (1,1/2)(1,3/2)(1,1/2)-(1,3/2) and (1,3/2)(1,5/2)(1,3/2)-(1,5/2) by 0.000322 cm-1 and 0.000516 cm-1 for 9Be2+, whereas for 7Be2+ they are 0.00038 cm-1 and 0.00061 cm-1, respectively. These shifts are about three orders of magnitude larger than that of 7Li+. Our final results of 23PJ2\,^{3}\!P_{J} hfs for 7Be2+ and 9Be2+ are shown in Tables 6 and Table 7.

Table 4: Calculated values of 7Be2+ and 9Be2+ hfs coefficients for the 21P12\,^{1}\!P_{1} and 23PJ2\,^{3}\!P_{J} states, in cm-1. These coefficients are defined in Eqs. (10)-(12) of Ref. [33]. The listed numerical values are uncertain only at the last digits.
Coefficient Be2+7{}^{7}\mathrm{Be}^{2+} Be2+9{}^{9}\mathrm{Be}^{2+}
C1,1(0)C^{(0)}_{1,1} –0.24990 –0.210283
C1,0(0)C^{(0)}_{1,0} –0.25126 –0.211428
D1(0)D^{(0)}_{1} –0.00440 –0.003701
D0(0)D^{(0)}_{0} –0.00317 –0.002663
E1,1(0)E^{(0)}_{1,1}    0.00112    0.000938
E1,0(0)E^{(0)}_{1,0}    0.00097    0.000815
Table 5: Hyperfine splittings in 23PJ2\,^{3}\!P_{J} of 7Be2+ and 9Be2+, in cm-1. Only the relativistic correction of order mα4\alpha^{4} is included. The listed numerical values are uncertain only at the last digits.
(J,F)(J,F)(J,F)-(J^{\prime},F^{\prime}) Method 1 Method 2 Difference
7Be2+ (2,1/2)(2,3/2)(2,1/2)-(2,3/2) 0.18729 0.18729
(2,3/2)(2,5/2)(2,3/2)-(2,5/2) 0.31552 0.31552
(2,5/2)(2,7/2)(2,5/2)-(2,7/2) 0.44873 0.44872 –0.00001
(1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.21031 0.21069    0.00038
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.31267 0.31328    0.00061
9Be2+ (2,1/2)(2,3/2)(2,1/2)-(2,3/2) 0.157960 0.157964    0.000004
(2,3/2)(2,5/2)(2,3/2)-(2,5/2) 0.265658 0.265660    0.000002
(2,5/2)(2,7/2)(2,5/2)-(2,7/2) 0.376899 0.376891 –0.000008
(1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.174864 0.175186    0.000322
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.264669 0.265185    0.000516
Table 6: Theoretical hyperfine intervals in the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of 7Be2+ with quadrupole moment Qd=6.11fm2Q_{\mathrm{d}}=-6.11~{}\rm{fm}^{2}. The listed numerical values are uncertain only at the last digits.
State\rm{State} (J,F)(J,F)(J,F)-(J^{\prime},F^{\prime}) EaE_{\rm{a}} 105X10^{5}X 106δX10^{6}\delta X η\eta EdE_{\mathrm{d}}
cm-1 cm/1fm2{}^{-1}/\rm{fm}^{2} cm/1fm2{}^{-1}/\rm{fm}^{2} ppm cm-1
23P22\,^{3}\!P_{2} (2,1/2)(2,3/2)(2,1/2)-(2,3/2) 0.18751    4.12051    0.30 1354 0.18726
(2,3/2)(2,5/2)(2,3/2)-(2,5/2) 0.31591    2.94322 –2.04    530 0.31574
(2,5/2)(2,7/2)(2,5/2)-(2,7/2) 0.44928 –4.12051    1.02    546 0.44953
23P12\,^{3}\!P_{1} (1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.21097 –5.29780 –0.43 1544 0.21130
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.31365    2.94322    2.17    616 0.31346
23S12\,^{3}\!S_{1} (1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.40951 0.40951
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.68250 0.68250
Table 7: Experimental and theoretical hyperfine intervals in the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of 9Be2+, in cm-1.
Experiment Theory
State\rm{State} (J,F)(J,F)(J,F)-(J^{\prime},F^{\prime}) Scholl et al. Johnson et al. This work
[16] [26]
23P22\,^{3}\!P_{2} (2,1/2)(2,3/2)(2,1/2)-(2,3/2) 0.1585(10) 0.1581 0.158371(7)
(2,3/2)(2,5/2)(2,3/2)-(2,5/2) 0.2659(11) 0.2659 0.266123(4)
(2,5/2)(2,7/2)(2,5/2)-(2,7/2) 0.3768(14) 0.3773 0.377128(4)
23P12\,^{3}\!P_{1} (1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.1751(10) 0.1754 0.175126(4)
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.2654(10) 0.2654 0.265662(3)
23S12\,^{3}\!S_{1} (1,1/2)(1,3/2)(1,1/2)-(1,3/2) 0.3448(10) 0.344574(9)
(1,3/2)(1,5/2)(1,3/2)-(1,5/2) 0.5740(11) 0.574275(6)

IV Discussion and conclusion

The radioactive 7Be is a special atomic nucleus whose magnetic moment cannot be obtained by the βγ\beta\gamma-NMR method, and optical spectroscopy is the only method to measure the nuclear moment. Although Okada et al. [36] determined the magnetic dipole moment of 7Be to high accuracy, its charge radius has not been determined until now. In addition, there is no published value for the nuclear electric quadrupole moment of 7Be. Fortunately, although Be7{}^{7}\mathrm{Be} is not stable, its half-life is about 53 days, which is helpful for the experimental measurement of 7Be. Since the quadrupole moment QdQ_{\mathrm{d}} of 7Be has not been measured and the results obtained by theoretical calculations differ noticeably from each other (–6.11 fm2\rm{fm}^{2} [37], –5.50(48) fm2\rm{fm}^{2} and –4.68(28) fm2\rm{fm}^{2}  [38]), these values are not yet conclusive. Here we study the contribution of the quadrupole moment QdQ_{\mathrm{d}} to hfs of 7Be2+ by ignoring its higher-order nonlinear correction,

Ed=Ea+Qd(X+δX),\displaystyle E_{\rm{d}}=E_{\rm{a}}+Q_{\rm{d}}(X+\delta X), (4)

where EdE_{\rm{d}} and EaE_{\rm{a}} represent the hfs obtained by diagonalization of Eq. (2) with and without the contribution of QdQ_{\rm{d}} (Heqm\langle H_{\mathrm{eqm}}\rangle term). XX and δX\delta X are tow linear coefficients independent of QdQ_{\rm{d}}, where XX is obtained from the diagonal element of the Heqm\langle H_{\mathrm{eqm}}\rangle term, and δX\delta X comes from the linear correction caused by the Heqm\langle H_{\mathrm{eqm}}\rangle term included in the diagonalization process. In order to reflect the sensitivity of the transitions to QdQ_{\rm{d}} intuitively, one can define the relative accuracy,

η=|Qd(X+δX)Ed|.\displaystyle\eta=\left\lvert\frac{Q_{\rm{d}}(X+\delta X)}{E_{\rm{d}}}\right\rvert. (5)

In other words, the η\eta is the precision required to detect the contribution of QdQ_{\rm{d}} in the experiment. Using the theoretical value Qd=6.11fm2Q_{\mathrm{d}}=-6.11~{}\rm{fm}^{2} chosen in the Ref. [30], we calculated the results as shown in Table 6. The results show that the transitions (1,1/2)(1,3/2)(1,1/2)-(1,3/2) and (2,1/2)(2,3/2)(2,1/2)-(2,3/2) are more suitable to be used to determine the QdQ_{\rm{d}}. According to the theoretical values, the value of QdQ_{d} is likely to exist between 7fm2-7\;\rm{fm}^{2} and 4fm2-4\;\rm{fm}^{2}. In this range, once the experiment reaches the same accuracy as the theory, the result of the QdQ_{\rm{d}} can be determined according to the Eq. (5), and the accuracy can have two significant digits.

Table 7 lists the experimental and theoretical hyperfine intervals in the 23PJ2\,^{3}\!P_{J} state of 9Be2+, where the uncertainties are mainly due to the mα7m\alpha^{7} contribution and the nuclear structure. It is worth noting that these theoretical uncertainties are propagated only from the errors displayed in Table 3. The uncertainty from the fine structure is canceled for the same-JJ transitions. Table 7 also shows the measured results obtained through weighted average of all the values in Ref. [16], and the only available theoretical values of Johnson et al. [26] for the 23PJ2\,^{3}\!P_{J} state. Our results are in good agreement with these previous values and are about two orders of magnitude more precise. Our theoretical calculation has reached the level of ten or so ppm, which is sensitive to some of the major nuclear electromagnetic structure effect.

In summary, we have studied the hfs of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states of the 7Be2+ and 9Be2+ ions, including the relativistic and QED corrections up to order mα6m\alpha^{6}. The 21P123P12\,^{1}\!P_{1}-2\,^{3}\!P_{1} single-triple mixing effect has been treated rigorously. Compared to Li+, the 21P123P12\,^{1}\!P_{1}-2\,^{3}\!P_{1} mixing effect is about three orders of magnitude larger, indicating that this procedure becomes more and more essential with increasing ZZ. The uncertainties of present calculations are in the order of tens of ppm for 9Be2+, mainly from the error of mα7m\alpha^{7} and nuclear contribution (the Zemach radius). The results for the hfs of the 23S12\,^{3}\!S_{1} and 23PJ2\,^{3}\!P_{J} states have been improved by two orders of magnitude. The contribution of nuclear electric quadrupole moment to the hyperfine splittings of 7Be2+ has also been studied. In order to observe the influence of QdQ_{\mathrm{d}}, the precision of experimental measurements on hfs needs to be better than 104cm110^{-4}\;\rm{cm}^{-1}. If the experiments reach the same accuracy as the present theoretical value, the two significant digits of QdQ_{\mathrm{d}} can be determined. Our results may stimulate further experimental activities to explore Be nuclear structure.

{\dagger}Email address: [email protected]

References

  • Yan and Drake [1995] Z.-C. Yan and G. W. F. Drake, High precision calculation of fine structure splittings in helium and He-like ions, Phys. Rev. Lett. 74, 4791 (1995).
  • Patkóš et al. [2016] V. c. v. Patkóš, V. A. Yerokhin, and K. Pachucki, Higher-order recoil corrections for triplet states of the helium atom, Phys. Rev. A 94, 052508 (2016).
  • Patkóš et al. [2017] V. c. v. Patkóš, V. A. Yerokhin, and K. Pachucki, Higher-order recoil corrections for singlet states of the helium atom, Phys. Rev. A 95, 012508 (2017).
  • Patkóš et al. [2019] V. c. v. Patkóš, V. A. Yerokhin, and K. Pachucki, Complete quantum electrodynamic α6m{\alpha}^{6}m correction to energy levels of light atoms, Phys. Rev. A 100, 042510 (2019).
  • Qi et al. [2020] X.-Q. Qi, P.-P. Zhang, Z.-C. Yan, G. W. F. Drake, Z.-X. Zhong, T.-Y. Shi, S.-L. Chen, Y. Huang, H. Guan, and K.-L. Gao, Precision calculation of hyperfine structure and the zemach radii of Li+6,7{{}^{6,7}\mathrm{Li}}^{+} ions, Phys. Rev. Lett. 125, 183002 (2020).
  • Patkóš et al. [2021a] V. c. v. Patkóš, V. A. Yerokhin, and K. Pachucki, Complete α7m{\alpha}^{7}m lamb shift of helium triplet states, Phys. Rev. A 103, 042809 (2021a).
  • Patkóš et al. [2021b] V. c. v. Patkóš, V. A. Yerokhin, and K. Pachucki, Radiative α7m{\alpha}^{7}m QED contribution to the helium lamb shift, Phys. Rev. A 103, 012803 (2021b).
  • Pachucki et al. [2004] K. Pachucki, U. D. Jentschura, and V. A. Yerokhin, Nonrelativistic QED approach to the bound-electron gg factor, Phys. Rev. Lett. 93, 150401 (2004).
  • Pachucki et al. [2005] K. Pachucki, U. D. Jentschura, and V. A. Yerokhin, Erratum: Nonrelativistic QED approach to the bound-electron gg factor [phys. rev. lett. 93, 150401 (2004)], Phys. Rev. Lett. 94, 229902 (2005).
  • Jentschura et al. [2005] U. D. Jentschura, A. Czarnecki, and K. Pachucki, Nonrelativistic QED approach to the lamb shift, Phys. Rev. A 72, 062102 (2005).
  • Pachucki et al. [2017] K. Pachucki, V. c. v. Patkóš, and V. A. Yerokhin, Testing fundamental interactions on the helium atom, Phys. Rev. A 95, 062510 (2017).
  • Pachucki and Yerokhin [2010] K. Pachucki and V. A. Yerokhin, Fine structure of heliumlike ions and determination of the fine structure constant, Phys. Rev. Lett. 104, 070403 (2010).
  • Clausen et al. [2021] G. Clausen, P. Jansen, S. Scheidegger, J. A. Agner, H. Schmutz, and F. Merkt, Ionization energy of the metastable 2 1S02\text{ }^{1}{S}_{0} state of He4{}^{4}\mathrm{He} from rydberg-series extrapolation, Phys. Rev. Lett. 127, 093001 (2021).
  • Guan et al. [2020] H. Guan, S. Chen, X.-Q. Qi, S. Liang, W. Sun, P. Zhou, Y. Huang, P.-P. Zhang, Z.-X. Zhong, Z.-C. Yan, G. W. F. Drake, T.-Y. Shi, and K. Gao, Probing atomic and nuclear properties with precision spectroscopy of fine and hyperfine structures in the Li+7{}^{7}\mathrm{Li}^{+} ion, Phys. Rev. A 102, 030801(R) (2020).
  • Yerokhin [2008] V. A. Yerokhin, Hyperfine structure of Li and Be+\mathrm{Be}^{+}Phys. Rev. A 78, 012513 (2008).
  • Scholl et al. [1993] T. J. Scholl, R. Cameron, S. D. Rosner, L. Zhang, R. A. Holt, C. J. Sansonetti, and J. D. Gillaspy, Precision measurement of relativistic and QED effects in heliumlike beryllium, Phys. Rev. Lett. 71, 2188 (1993).
  • Yan et al. [2008] Z.-C. Yan, W. Nörtershäuser, and G. W. F. Drake, High precision atomic theory for Li and Be+\mathrm{Be}^{+}: QED shifts and isotope shifts, Phys. Rev. Lett. 100, 243002 (2008).
  • Nörtershäuser et al. [2015a] W. Nörtershäuser, C. Geppert, A. Krieger, K. Pachucki, M. Puchalski, K. Blaum, M. L. Bissell, N. Frömmgen, M. Hammen, M. Kowalska, J. Krämer, K. Kreim, R. Neugart, G. Neyens, R. Sánchez, and D. T. Yordanov, Precision test of many-body QED in the Be+\mathrm{Be}^{+} 2P2{P} fine structure doublet using short-lived isotopes, Phys. Rev. Lett. 115, 033002 (2015a).
  • Labiche et al. [1999] M. Labiche, F. M. Marqués, O. Sorlin, and N. Vinh Mau, Structure of Be13{}^{13}\mathrm{Be} and Be14{}^{14}\mathrm{Be}Phys. Rev. C 60, 027303 (1999).
  • Nörtershäuser et al. [2009] W. Nörtershäuser, D. Tiedemann, M. Žáková, Z. Andjelkovic, K. Blaum, M. L. Bissell, R. Cazan, G. W. F. Drake, C. Geppert, M. Kowalska, J. Krämer, A. Krieger, R. Neugart, R. Sánchez, F. Schmidt-Kaler, Z.-C. Yan, D. T. Yordanov, and C. Zimmermann, Nuclear charge radii of Be7,9,10{}^{7,9,10}\mathrm{Be} and the one-neutron halo nucleus Be11{}^{11}\mathrm{Be}Phys. Rev. Lett. 102, 062503 (2009).
  • Fortune and Sherr [2012] H. T. Fortune and R. Sherr, Consistent description of 11Be and 12Be and of the 11Be(dd,pp)12Be reaction, Phys. Rev. C 85, 051303 (2012).
  • Krieger et al. [2012] A. Krieger, K. Blaum, M. L. Bissell, N. Frömmgen, C. Geppert, M. Hammen, K. Kreim, M. Kowalska, J. Krämer, T. Neff, R. Neugart, G. Neyens, W. Nörtershäuser, C. Novotny, R. Sánchez, and D. T. Yordanov, Nuclear charge radius of Be12{}^{12}\mathrm{Be}Phys. Rev. Lett. 108, 142501 (2012).
  • Lu et al. [2013] Z.-T. Lu, P. Mueller, G. W. F. Drake, W. Nörtershäuser, S. C. Pieper, and Z.-C. Yan, Colloquium: Laser probing of neutron-rich nuclei in light atoms, Rev. Mod. Phys. 85, 1383 (2013).
  • Nörtershäuser et al. [2015b] W. Nörtershäuser, C. Geppert, A. Krieger, K. Pachucki, M. Puchalski, K. Blaum, M. L. Bissell, N. Frömmgen, M. Hammen, M. Kowalska, J. Krämer, K. Kreim, R. Neugart, G. Neyens, R. Sánchez, and D. T. Yordanov, Precision test of many-body QED in the Be+\mathrm{Be}^{+} 2P2{P} fine structure doublet using short-lived isotopes, Phys. Rev. Lett. 115, 033002 (2015b).
  • Puchalski et al. [2021] M. Puchalski, J. Komasa, and K. Pachucki, Hyperfine structure of the 23P2^{3}{P} state in Be9{}^{9}\mathrm{Be} and the nuclear quadrupole moment, Phys. Rev. Research 3, 013293 (2021).
  • Johnson et al. [1997] W. R. Johnson, K. T. Cheng, and D. R. Plante, Hyperfine structure of 23P{2}^{3}{P} levels of heliumlike ions, Phys. Rev. A 55, 2728 (1997).
  • Jones et al. [2005] R. J. Jones, K. D. Moll, M. J. Thorpe, and J. Ye, Phase-coherent frequency combs in the vacuum ultraviolet via high-harmonic generation inside a femtosecond enhancement cavity, Phys. Rev. Lett. 94, 193201 (2005).
  • Cingöz et al. [2012] A. Cingöz, D. C. Yost, T. K. Allison, A. Ruehl, M. E. Fermann, I. Hartl, and J. Ye, Direct frequency comb spectroscopy in the extreme ultraviolet, Nature 482, 68 (2012).
  • Zhang et al. [2020] J. Zhang, L.-Q. Hua, Z. Chen, M.-F. Zhu, C. Gong, and X.-J. Liu, Extreme ultraviolet frequency comb with more than 100 μ\muw average power below 100nm, Chin. Phys. Lett. 37, 124203 (2020).
  • Puchalski and Pachucki [2009] M. Puchalski and K. Pachucki, Fine and hyperfine splitting of the 2P2{P} state in Li and Be+Phys. Rev. A 79, 032510 (2009).
  • Pachucki et al. [2012] K. Pachucki, V. A. Yerokhin, and P. Cancio Pastor, Quantum electrodynamic calculation of the hyperfine structure of 3He, Phys. Rev. A 85, 042517 (2012).
  • Haidar et al. [2020] M. Haidar, Z.-X. Zhong, V. I. Korobov, and J.-P. Karr, Nonrelativistic QED approach to the fine- and hyperfine-structure corrections of order mα6m{\alpha}^{6} and mα6(m/m)m{\alpha}^{6}(m/m): Application to the hydrogen atom, Phys. Rev. A 101, 022501 (2020).
  • Riis et al. [1994] E. Riis, A. G. Sinclair, O. Poulsen, G. W. F. Drake, W. R. C. Rowley, and A. P. Levick, Lamb shifts and hyperfine structure in Li+6{}^{6}\mathrm{Li}^{+} and Li+7{}^{7}\mathrm{Li}^{+}: Theory and experiment, Phys. Rev. A 49, 207 (1994).
  • Zhang et al. [2015] P.-P. Zhang, Z.-X. Zhong, Z.-C. Yan, and T.-Y. Shi, Precision calculation of fine structure in helium and Li+Chin. Phys. B 24, 033101 (2015).
  • Drake [2002] G. W. F. Drake, Progress in helium fine-structure calculations and the fine-structure constant, Can. J. Phys. 80, 1195 (2002).
  • Okada et al. [2008] K. Okada, M. Wada, T. Nakamura, A. Takamine, V. Lioubimov, P. Schury, Y. Ishida, T. Sonoda, M. Ogawa, Y. Yamazaki, Y. Kanai, T. M. Kojima, A. Yoshida, T. Kubo, I. Katayama, S. Ohtani, H. Wollnik, and H. A. Schuessler, Precision measurement of the hyperfine structure of laser-cooled radioactive Be+7{}^{7}\mathrm{Be}^{+} ions produced by projectile fragmentation, Phys. Rev. Lett. 101, 212502 (2008).
  • Koji et al. [2001] A. Koji, O. Yoko, S. Yasuyuki, and V. Kálmán, Microscopic multicluster description of light exotic nuclei with stochastic variational method on correlated gaussians, Prog. Theor. Phys. Suppl. 142, 97 (2001).
  • Forssén et al. [2009] C. Forssén, E. Caurier, and P. Navrátil, Charge radii and electromagnetic moments of Li and Be isotopes from the ab initio no-core shell model, Phys. Rev. C 79, 021303 (2009).