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

Combination of the Perturbation Theory with Configuration Interaction Method

M. G. Kozlov1,2, I. I. Tupitsyn1,3, A. I. Bondarev1,4, and D. V. Mironova2 1Petersburg Nuclear Physics Institute of NRC “Kurchatov Institute”, 188300 Gatchina, Russia 2St. Petersburg Electrotechnical University “LETI”, Prof. Popov Str. 5, 197376 St. Petersburg, Russia 3Department of Physics, St. Petersburg State University, Ulianovskaya 1, Petrodvorets, 198504 St. Petersburg, Russia 4Center for Advanced Studies, Peter the Great St. Petersburg Polytechnic University, Polytekhnicheskaja 29, 195251 St. Petersburg Russia
Abstract

Present atomic theory provides accurate and reliable results for atoms with a small number of valence electrons. However, most current methods of calculations fail when the number of valence electrons exceeds four or five. This means that we can not make reliable predictions for more than a half of the periodic table. Here we suggest a modification of the CI+MBPT (configuration interaction plus many-body perturbation theory) method, which may be applicable to atoms and ions with filling dd and ff shells.

I Introduction

At present there are several methods of the relativistic correlation calculations of atoms, such as multiconfiguration Dirac-Fock Desclaux (1975); Jönsson et al. (2013); Froese Fischer et al. (2016), configuration interaction (CI) Gu (2008); Fritzsche et al. (2002); Jiang et al. (2016); Tupitsyn et al. (2003, 2005), many-body perturbation theory (MBPT) Dzuba et al. (1987); Blundell et al. (1987); Sapirstein (1998), CI+MBPT Dzuba et al. (1996); Kahl and Berengut (2019); Kozlov et al. (2015), coupled cluster Eliav and Kaldor (2010); Saue et al. (2020); Blundell et al. (1989); Safronova and Johnson (2008); Oleynichenko et al. (2020), and others. Calculations are usually done in the no-virtual-pair approximation using Dirac-Coulomb, or Dirac-Coulomb-Breit approximations Johnson (2007). QED corrections may be included using radiative potential method developed by Flambaum and Ginges (2005), Ginges and Berengut (2016), and QEDMOD potential Shabaev et al. (2013); Tupitsyn et al. (2016).

The coupled cluster method is one of the most popular and effective methods for calculation of atoms with a small number of open shell electrons (or holes). Calculations of the spectra of atoms and ions with many valence electrons (e. g. transition metals, lanthanides, and actinides) are very difficult and usually not very accurate. The reason for that is a combination of strong correlations and a very large configuration space. To account for strong correlations one needs non-perturbative methods, such as CI. On the other hand, a large configuration space makes such calculations very expensive. As a compromise one can try to combine CI with perturbation theory (PT). We will first assume that all closed atomic shells are considered frozen. Then we are treating only valence correlations and consider a combination of the valence CI with valence perturbation theory (VPT). Later we will see that this approach can also be used to treat core-valence correlations.

Recently there were several attempts Dzuba et al. (2017); Geddes et al. (2018); Dzuba et al. (2019); Imanbaeva and Kozlov (2019) to develop an effective and fast CI+VPT method to speed up calculations for such systems, where straightforward CI calculations are impossible. Application of these methods for systems with a large number of valence electrons was demonstrated in Refs. Cheung et al. (2020); Li and Dzuba (2020). A general idea of all these calculation schemes is to make CI in a smaller subspace PP and calculate corrections from a complementary subspace QQ using VPT. In Refs. Dzuba et al. (2017); Geddes et al. (2018); Dzuba et al. (2019) it is suggested to neglect non-diagonal blocks of the CI matrix in the subspace QQ, which is equivalent to using VPT. All these methods require summation over all determinants of the complementary subspace QQ. Though calculating this sum is much easier than calculating and diagonalizing the whole CI matrix, it is still too expensive for the number of valence electrons approaching, or exceeding ten.

In the paper Dzuba et al. (2019), the sum over determinants was partly substituted by the sum over configurations that led to a significant increase in calculation speed. Here we want to make another step in this direction. To this end, we will partly substitute VPT with many-body perturbation theory (MBPT). The method we propose here is similar to the old CI+MBPT method Dzuba et al. (1996) but uses different splitting of the problem into the CI and MBPT parts. In particular, we suggest to account for double excitations (D) from the subspace PP by means of the MBPT and treat single excitations (S) within VPT, or, if possible, include them directly in CI. We think that this variant is not only more efficient for treating valence correlations, but may also be used for the core-valence correlations.

II Formalism

II.1 Valence correlations

Consider many-electron atom, or ion with NN valence electrons, where N1N\gg 1. Let us first assume that other electrons always occupy closed core shells, which is known as a frozen core approximation. Our aim is to solve the NN-electron Schrödinger equation and find the spectrum of this system.

We start with splitting NN-electron configuration space in two orthogonal subspaces PP and QQ. The subspace PP, which we call valence, includes the most important shells. It may be not obvious from the start, which orbitals are ‘important’. We definitely must include into subspace PP all orbitals with occupation numbers of the order of unity in the physical states, we are interested in. Complementary subspace QQ includes S, D, and so on excitations from the valence shells to the virtual ones, thus, Q=QS+QD+Q=Q_{S}+Q_{D}+\dots. We start by solving the matrix equation in the subspace PP,

P^HP^Ψa\displaystyle\hat{P}H\hat{P}\Psi_{a} =EaP^Ψa,\displaystyle=E_{a}\hat{P}\Psi_{a}\,, (1)

where HH is the Hamiltonian for valence electrons and P^\hat{P} is the projector on the subspace PP. We can find a correction from the complementary subspace QQ using the second-order perturbation theory:

δEa\displaystyle\delta E_{a} =nQΨa|P^HQ^|nn|Q^HP^|ΨaEaEn,\displaystyle=\sum_{n\in Q}\frac{\langle\Psi_{a}|\hat{P}H\hat{Q}|n\rangle\langle n|\hat{Q}H\hat{P}|\Psi_{a}\rangle}{E_{a}-E_{n}}, (2)

where |n|n\rangle are NN-electron determinants in the complementary subspace QQ and En=n|Q^HQ^|nE_{n}=\langle n|\hat{Q}H\hat{Q}|n\rangle.

The wavefunction Ψa\Psi_{a} is a linear combination of the determinants:

Ψa\displaystyle\Psi_{a} =mPCma|m=p,mpCp,mpa|mp.\displaystyle=\sum_{m\in P}C^{a}_{m}|m\rangle=\sum_{p,m_{p}}C^{a}_{p,m_{p}}|m_{p}\rangle\,. (3)

Here and below indexes pp and qq run over configurations in the subspaces PP and QQ respectively and indexes mpm_{p} and nqn_{q} numerate determinants within one configuration. Now Eq. (2) takes the form:

δEa=p,mpp,mpCp,mpaCp,mpa×q,nqmp|H|nqnq|H|mpEaEnq,\delta E_{a}=\sum_{p,m_{p}}\sum_{p^{\prime},m_{p^{\prime}}}C^{a}_{p,m_{p}}C^{a}_{p^{\prime},m_{p^{\prime}}}\\ \times\sum_{q,n_{q}}\frac{\langle m_{p}|H|n_{q}\rangle\langle n_{q}|H|m_{p^{\prime}}\rangle}{E_{a}-E_{n_{q}}}, (4)

where the sum over the subspace QQ is also split in two.

For an atom with N10N\sim 10, the dimension of space QQ is very large, which makes evaluation of expression (4) very lengthy. Therefore, our aim is to substitute double sum over qq and nqn_{q} by a single sum over qq. To this end we do the following approximation: we substitute the energy EnqE_{n_{q}} in the denominator by the configuration average:

E¯q\displaystyle\bar{E}_{q} =1Nqnq=1NqEnq,\displaystyle=\frac{1}{N_{q}}\sum_{n_{q}=1}^{N_{q}}E_{n_{q}}\,, (5)

where NqN_{q} is the number of determinants in configuration qq. Using this approximation we rewrite (4) in a form:

δEa=p,mpp,mpCp,mpaCp,mpa×qmp|H(nq|nqnq|)H|mpEaE¯q.\delta E_{a}=\sum_{p,m_{p}}\sum_{p^{\prime},m_{p^{\prime}}}C^{a}_{p,m_{p}}C^{a}_{p^{\prime},m_{p^{\prime}}}\\ \times\sum_{q}\frac{\langle m_{p}|H\left({\sum_{n_{q}}|n_{q}\rangle\langle n_{q}|}\right)H|m_{p^{\prime}}\rangle}{E_{a}-\bar{E}_{q}}\,. (6)

Below we will show that in some very important cases one can get rid of the internal sum over nqn_{q}.

Hamiltonian HH includes one-particle and two-particle parts. The former consists of the kinetic term and the core potential, while the latter corresponds to the Coulomb (or Coulomb-Breit) interaction between valence electrons. Thus, in the sum over qq remain only configurations, which differ by no more than two electrons from configurations pp and pp^{\prime}. This means that within this approximation the subspace QQ is actually truncated to QS+QDQ_{S}+Q_{D}. All non-zero contributions correspond to the diagrams, shown in Fig. 1.

Refer to caption
Figure 1: Set of connected second-order diagrams. Black dots correspond to the core potential and wavy lines to the Coulomb interaction. Double and single lines denote electrons in valence and virtual orbitals respectively. Non-symmetric diagrams (b)(b) and (e)(e) have mirror twins.

According to our definition of the spaces PP and QQ, the latter must include at least one electron in the virtual shell. Diagrams (a)(a), (b)(b), and (e)(e) include only one intermediate line, so they describe single excitations from the subspace PP. Diagrams (c)(c) and (d)(d) include two intermediate lines, but only diagram (d)(d) describes double (D) excitations, as both intermediate lines correspond to the virtual shells.

Figure 1 shows that all many-electron matrix elements in Eq. (6) are reduced to the effective one-electron, two-electron, and three-electron contributions. Effective one-electron contributions are described by diagram (a)(a); diagrams (b)(b), (c)(c), and (d)(d) correspond to the two-electron contributions; finally, diagram (e)(e) describes effective three-electron contributions, see Figs. 2 and 3.

Refer to caption
Figure 2: Many-electron second-order expression in Eq. (7) (left) is reduced to the two-particle expression (middle), which, in turn, is reduced to the effective two-particle interaction (right).
Refer to caption
Figure 3: The case when many-electron second-order expression in Eq. (6) (left) is reduced to the three-particle expression (middle), which, in turn, is reduced to the effective three-particle interaction (right). The initial configuration on the left differs from the intermediate configuration by the upper two electrons. The final configuration differs from the intermediate one by the second and third electron from the top.

For combinatorial reasons the number of configurations with two excited electrons is much bigger, than the number of those with only one such electron. Therefore the vast majority of terms in Eq. (6) correspond to the two-electron excitations from configurations pp and pp^{\prime}. For these terms in the Hamiltonian HH only the two-electron interaction VV can contribute, so we can neglect the one-electron part and make substitution HVH\to V. As we saw above, all such terms are described by the single diagram (d)(d) from Fig. 1.

Let us consider the sum over doubly excited configurations. It can be written as:

δEaD=p,mpp,mpCp,mpaCp,mpa×qQDmp|V(nq|nqnq|)V|mpEaE¯q,\delta E_{a}^{D}=\sum_{p,m_{p}}\sum_{p^{\prime},m_{p^{\prime}}}C^{a}_{p,m_{p}}C^{a}_{p^{\prime},m_{p^{\prime}}}\\ \times\sum_{q\in Q_{D}}\frac{\langle m_{p}|V\left({\sum_{n_{q}}|n_{q}\rangle\langle n_{q}|}\right)V|m_{p^{\prime}}\rangle}{E_{a}-\bar{E}_{q}}\,, (7)

Non-zero contributions come from determinants |nq|n_{q}\rangle, which differ from both determinants |mp|m_{p}\rangle and |mp|m_{p^{\prime}}\rangle by two electrons. It is clear that it must be the same two electrons, see Fig. 2. In this case the second-order expression from (7) is reduced to the effective two-particle interaction Dzuba et al. (1996):

δEaD\displaystyle\delta E_{a}^{D} =p,mpp,mpCp,mpaCp,mpamp|Veff|mp.\displaystyle=\sum_{p,m_{p}}\sum_{p^{\prime},m_{p^{\prime}}}C^{a}_{p,m_{p}}C^{a}_{p^{\prime},m_{p^{\prime}}}\langle m_{p}|V_{\mathrm{eff}}|m_{p^{\prime}}\rangle\,. (8)

This effective interaction can be expressed in terms of the effective radial integrals, which are similar to the Coulomb radial integrals. The latter appear when we expand Coulomb interaction in spherical multipoles,

V=k,ϰVϰk.\displaystyle V=\sum_{k,\varkappa}V^{k}_{\varkappa}. (9)

The matrix element of each multipole component VϰkV^{k}_{\varkappa} has the form:

c,d|Vϰk|a,b=(1)mc+mb+1δp(2ja+1)(2jb+1)(2jc+1)(2jd+1)(jcjakmcmaϰ)(jbjdkmbmdϰ)(jcjak12120)(jbjdk12120)Ra,b,c,dk,\langle c,d|V_{\varkappa}^{k}|a,b\rangle=(-1)^{m_{c}+m_{b}+1}\delta_{p}\sqrt{(2j_{a}+1)(2j_{b}+1)(2j_{c}+1)(2j_{d}+1)}\\ \left(\!\begin{array}[]{ccc}j_{c}&j_{a}&k\\ -m_{c}&m_{a}&\varkappa\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{b}&j_{d}&k\\ -m_{b}&m_{d}&\varkappa\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{c}&j_{a}&k\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{b}&j_{d}&k\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)R^{k}_{a,b,c,d}\,, (10)

where round brackets denote 3j-symbols, Ra,b,c,dkR^{k}_{a,b,c,d} denotes the Coulomb radial integral and δp\delta_{p} ensures parity selection rule: δp=ξ(la+lc+k)ξ(lb+ld+k)\delta_{p}=\xi(l_{a}+l_{c}+k)\xi(l_{b}+l_{d}+k) and ξ(n)=1,0\xi(n)=1,0 for n=even, oddn=\mbox{even, odd}. A similar multipole expansion holds for the effective interaction VeffV_{\mathrm{eff}}, the effective radial integral being Dzuba et al. (1996):

Ra,b,c,dk,eff=k1,k2m,n(1)χ(2jm+1)(2jn+1)(2k+1){jcjakk1k2jm}{jbjdkk2k1jn}(jmjak112120)(jbjnk112120)(jcjmk212120)(jnjdk212120)(jcjak12120)1(jbjdk12120)1Ra,b,m,nk1Rc,d,m,nk2ΔE,R^{k,\mathrm{eff}}_{a,b,c,d}=\sum_{k_{1},k_{2}}\sum_{m,n}(-1)^{\chi}(2j_{m}+1)(2j_{n}+1)(2k+1)\left\{\!\begin{array}[]{ccc}j_{c}&j_{a}&k\\ k_{1}&k_{2}&j_{m}\\ \end{array}\!\right\}\left\{\!\begin{array}[]{ccc}j_{b}&j_{d}&k\\ k_{2}&k_{1}&j_{n}\\ \end{array}\!\right\}\left(\!\begin{array}[]{ccc}j_{m}&j_{a}&k_{1}\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{b}&j_{n}&k_{1}\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)\\ \left(\!\begin{array}[]{ccc}j_{c}&j_{m}&k_{2}\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{n}&j_{d}&k_{2}\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)\left(\!\begin{array}[]{ccc}j_{c}&j_{a}&k\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\!\right)^{-1}\!\!\left(\begin{array}[]{ccc}j_{b}&j_{d}&k\\ \frac{1}{2}&-\frac{1}{2}&0\\ \end{array}\right)^{-1}\frac{R^{k_{1}}_{a,b,m,n}R^{k_{2}}_{c,d,m,n}}{\Delta_{E}}\,, (11)

where curly brackets denote 6j-coefficients, the phase χ=ja+jb+jc+jd+jm+jn+k1+k2+k+1\chi={j_{a}+j_{b}+j_{c}+j_{d}+j_{m}+j_{n}+k_{1}+k_{2}+k+1}, and ΔE\Delta_{E} is energy denominator, which we will discuss later. For the effective interaction there is no link between parity and multipolarity kk, so for VeffV_{\mathrm{eff}} we do not have factor δp\delta_{p} as in Eq. (10). The sum in (11) runs over multipolarities k1k_{1} and k2k_{2}, which satisfies the triangle rule |k1k2|kk1+k2|k_{1}-k_{2}|\leq k\leq k_{1}+k_{2}.

All single excitations are described by the remaining diagrams from Fig. 1. The diagram (a)(a) has a form of the effective one-electron radial integral, while diagrams (b)(b) and (c)(c) are reduced to the two-electron effective radial integrals. In principle, these effective radial integrals can be calculated and stored. However, the diagram (e)(e) corresponds to the effective three-particle interaction. It is difficult to include such interactions into CI matrix for several reasons:

  • When N>3N>3 the number of such effective three-particle integrals is huge.

  • It is difficult to store them and find them.

  • The number of the non-zero matrix elements in the matrix drastically increases. The matrix becomes less sparse and its diagonalization is much more difficult and time-consuming.

Because of all that it is inefficient to use the MBPT approach for three-particle diagrams and it is much easier to treat them within the determinant-based PT. However, it is difficult then to separate them from other contributions, which correspond to single excitations. Thus, it is better not to use MBPT for single excitations at all. We suggest to use instead any form of the determinant-based VPT described in Refs. Dzuba et al. (2017); Geddes et al. (2018); Imanbaeva and Kozlov (2019); Dzuba et al. (2019). This means that we do VPT in the subspace QSQ_{S}. Note that the dimension of this subspace is incomparably smaller than the dimension of the QDQ_{D} subspace. In some cases it may be so small, that we can include QSQ_{S} in the subspace PP, where we do CI.

II.2 Core-valence correlations

Refer to caption
Figure 4: Set of one-electron second-order diagrams accounting for the excitations from the core. Diagrams (e)(e) and (f)(f) have mirror twins. Diagrams (c)(c) and (d)(d) describe double excitations from the core.

It is easy to use the scheme described above for the core-valence correlations as well. Now PP subspace corresponds to the frozen-core approximation and the subspaces QSQ_{S} and QDQ_{D} include single and double excitations from the core respectively. This means that these subspaces include many-electron states with one and two holes in the core. As before, the second-order MBPT corrections are described by one-electron, two-electron, and three-electron diagrams. All one-electron diagrams are given in Fig. 4. Excitations from the core correspond to the hole lines with arrows looking to the left. It is easy to see that only diagrams (c)(c) and (d)(d) describe double excitations. Therefore, we need to calculate them and store as one-electron effective radial integrals, see Fig. 5 (note that there are no one-electron contributions for the valence excitations). Expressions for these diagrams were given in Ref. Dzuba et al. (1996).

There is only one two-electron diagram, which corresponds to the double excitations from the core. This diagram must be calculated and added to the similar diagram for valence excitations, which was discussed in the previous section, see Fig. 6. Finally, in analogy with the valence correlations, the three-particle diagrams correspond to the single excitations from the core.

Refer to caption
Figure 5: Diagrams, which correspond to the double excitations from closed shells. These diagrams are described by the effective one-electron radial integrals, designated by a black circle.
Refer to caption
Figure 6: Diagrams contributing to the effective two-electron radial integrals. First diagram accounts for the double excitations to the virtual shells and second diagram accounts for the double excitations from closed shells.

We conclude that in order to account for both valence and core-valence correlations we need to calculate one-electron and two-electron effective radial integrals, which corresponds to the diagrams from Figs. 5 and 6. At the same time, we need to include all single excitations from the core shells and all single excitations to the virtual shells either in the subspace PP or in the subspace QSQ_{S}. After that, we make CI calculation with effective radial integrals possibly followed by the VPT calculation in the QSQ_{S} subspace.

II.3 Sketch of the possible calculation scheme

Let us describe a most general computational scheme.

  • Basis set orbitals are divided into four groups: inner core, outer core, valence, and virtual orbitals. The inner core is kept frozen on all stages of calculation.

  • Effective radial integrals are calculated for the valence orbitals, which account for the double excitations from the outer core and the double excitations from the valence orbitals to the virtual ones.

  • Full CI calculation is done for the valence electrons. The effective radial integrals are added to the conventional radial integrals when the Hamiltonian matrix is formed.

  • Determinant-based PT is used in the complementary subspace QSQ_{S}, which includes single excitations from the outer core and single excitations to the virtual states.

Depending on the number of the valence electrons and the size of the core this scheme can be simplified. If there are only two valence electrons, one can include all virtual basis states into valence space. Single excitation from the core can be also added to the valence space. Double excitations from the core are accounted for through the effective radial integrals, while single excitations are included explicitly in the CI matrix. Formally this means that we substitute PP, QQ decomposition by the PP^{\prime}, QDQ_{D} decomposition:

P+Q=P+QS+QD=P+QD,\displaystyle P+Q=P+Q_{S}+Q_{D}=P^{\prime}+Q_{D}\,, (12)
PP+QS.\displaystyle P^{\prime}\equiv P+Q_{S}\,. (13)

In the new valence space PP^{\prime}, we solve matrix equation with the energy-dependent effective Hamiltonian Dzuba et al. (1996):

Heff(E)=H+Veff(E),\displaystyle H_{\mathrm{eff}}(E)=H+V_{\mathrm{eff}}(E)\,, (14)
P^Heff(Ea)P^Ψa=EaP^Ψa,\displaystyle\hat{P^{\prime}}H_{\mathrm{eff}}(E_{a})\hat{P^{\prime}}\Psi_{a}=E_{a}\hat{P^{\prime}}\Psi_{a}\,, (15)

where P^\hat{P^{\prime}} is the projector on the subspace PP^{\prime}. When the size of the matrix HeffH_{\mathrm{eff}} becomes too large, one can neglect the non-diagonal part of the matrix in the QSQ_{S} space, as in the emu CI method Geddes et al. (2018).

III Energy denominators

Let us discuss the energy denominator ΔE\Delta_{E} in Eq. (11). For simplicity we will consider the Rayleigh-Schrödinger perturbation theory, where the denominator in Eq. (6) would be E¯pE¯q\bar{E}_{p}-\bar{E}_{q}. Here E¯p\bar{E}_{p} and E¯q\bar{E}_{q} are average energies (5) for configurations pp and qq. Note that in order to return to the Brillouin-Wigner perturbation theory we will need to add EaE¯pE_{a}-\bar{E}_{p}, which can be approximately done using the method suggested in Dzuba et al. (1996).

In the conventional MBPT the denominator E¯pE¯q\bar{E}_{p}-\bar{E}_{q} is reduced to the difference of the Hartree-Fock energies of the orbitals εi\varepsilon_{i} which are different in these two configurations. That would give the following energy denominator in Eq. (11):

ΔEΔE(abmn)=εa+εbεmεn,\displaystyle\Delta_{E}\equiv\Delta_{E}(ab\to mn)=\varepsilon_{a}+\varepsilon_{b}-\varepsilon_{m}-\varepsilon_{n}\,, (16)

where we assume that configuration qq differs from pp by excitation of two electrons from shells aa and bb to virtual shells mm and nn respectively. This expression neglects the interaction of the electrons with each other and depends on the choice of the Hartree-Fock potential. In order to improve this approximation, we will consider general expression for the average energy of the relativistic electronic configuration.

III.1 Average energy of the relativistic configuration

The average energy of the relativistic configuration E¯p\bar{E}_{p} Mann (1973); Grant (1970):

E¯p=apqaIa+12apqa(qa1)Uaa+a<b;a,bpqaqbUab,\bar{E}_{p}=\sum_{a\in p}q_{a}\,I_{a}+\tfrac{1}{2}\,\sum_{a\in p}q_{a}\,(q_{a}-1)\,U_{aa}\\ +\sum_{a<b;\,a,b\in p}q_{a}\,q_{b}\,U_{ab}, (17)

where qaq_{a} and qbq_{b} are occupation numbers for the shells aa and bb in configuration pp and matrix elements of the potential UU are given by:

Uab={F0(a,a)+k>02fa,akFk(a,a),a=b,F0(a,b)+kga,bkGk(a,b),ab.\displaystyle U_{ab}=\left\{\begin{array}[]{ll}\displaystyle F^{0}(a,a)+\sum_{k>0}2\,f^{k}_{a,a}\,F^{k}(a,a)\,,&a=b\,,\\[14.22636pt] \displaystyle F^{0}(a,b)+\sum_{k}g^{k}_{a,b}\,G^{k}(a,b)\,,&a\neq b\,.\end{array}\right. (20)

In these equations IaI_{a} is the one-electron radial integral, while Fk(a,b)F^{k}(a,b) and Gk(a,b)G^{k}(a,b) are standard Coulomb and exchange two-electron radial integrals Grant (1970). The angular factors fa,akf^{k}_{a,a} and ga,bkg^{k}_{a,b} are also defined in agreement with Ref. Grant (1970):

fa,ak=122ja+12ja(jajak12120)2,ga,bk=(jajbk12120)2,\displaystyle\begin{array}[]{lll}f^{k}_{a,a}&=\displaystyle-\,\frac{1}{2}\,\frac{2j_{a}+1}{2j_{a}}\,\left(\begin{array}[]{llll}j_{a}&j_{a}&k\\ \frac{1}{2}&-\frac{1}{2}&0\end{array}\right)^{2}\,,\\[14.22636pt] \displaystyle g^{k}_{a,b}&=\displaystyle-\left(\begin{array}[]{llll}j_{a}&j_{b}&k\\ \frac{1}{2}&-\frac{1}{2}&0\end{array}\right)^{2}\,,\end{array} (27)

where jaj_{a} and jbj_{b} are the one-electron total angular momenta.

Let us use Eq. (17) to calculate the energy difference between configurations pp and qq which differ by the excitation of two electrons from shells a,ba,b to shells m,nm,n. In other words we need to calculate how the energy changes when occupation numbers change in the following way: δqa=δqb=1\delta q_{a}=\delta q_{b}=-1 and δqm=δqn=1\delta q_{m}=\delta q_{n}=1. To this end we, can use Taylor expansion of Eq. (17) near the initial configuration pp:

E¯q=E¯p+aE¯pqaδqa+12a,b2E¯pqaqbδqaδqb,\displaystyle\bar{E}_{q}=\bar{E}_{p}+\sum_{a}\frac{\partial\bar{E}_{p}}{\partial q_{a}}\,\delta q_{a}+\frac{1}{2}\sum_{a,b}\frac{\partial^{2}\bar{E}_{p}}{\partial q_{a}\partial q_{b}}\,\delta q_{a}\,\delta q_{b}\,, (28)

where derivatives are given by:

E¯pqa\displaystyle\frac{\partial\bar{E}_{p}}{\partial q_{a}} =Ia+(qa12)Uaa+baqbUab\displaystyle=I_{a}+(q_{a}-\tfrac{1}{2})U_{aa}+\sum_{b\neq a}q_{b}U_{ab}
=Ia12Uaa+bqbUab,\displaystyle=I_{a}-\tfrac{1}{2}U_{aa}+\sum_{b}q_{b}U_{ab}\,, (29)
2E¯pqaqb\displaystyle\frac{\partial^{2}\bar{E}_{p}}{\partial q_{a}\partial q_{b}} =Uab.\displaystyle=U_{ab}\,. (30)

Note that all higher derivatives vanish, so expression (28) is exact. With its help we get:

ΔE(abmn)=Ia+IbImIn+cpqc(Uac+UbcUmcUnc)UaaUbbUabUmn+Uam+Ubn+Uan+Ubm.\Delta_{E}(ab\to mn)=I_{a}+I_{b}-I_{m}-I_{n}\\ +\sum_{c\in p}q_{c}\,(U_{ac}+U_{bc}-U_{mc}-U_{nc})\\ -U_{aa}-U_{bb}-U_{ab}-U_{mn}\\ +U_{am}+U_{bn}+U_{an}+U_{bm}\,. (31)

This expression can be also used for the special cases a=b,δqa=2a=b,\,\delta q_{a}=-2 and/or m=n,δqm=2m=n,\,\delta q_{m}=2.

Equation (31) includes the sum over the occupied shells of the initial configuration pp. Let us introduce one-electron energies in respect to this configuration as:

εa\displaystyle\varepsilon_{a} =Ia+cpqcUac(1δqa,0)Uaa.\displaystyle=I_{a}+\sum_{c\in p}q_{c}\,U_{ac}-(1-\delta_{q_{a},0})\,U_{aa}\,. (32)

Then Eq. (31) is simplified to

ΔE(abmn)=εa+εbεmεnUabUmn+Uam+Ubn+Uan+Ubm.\Delta_{E}(ab\to mn)=\varepsilon_{a}+\varepsilon_{b}-\varepsilon_{m}-\varepsilon_{n}\\ -U_{ab}-U_{mn}+U_{am}+U_{bn}+U_{an}+U_{bm}\,. (33)

The first line here reproduces the conventional MBPT denominator (16), while the second line gives corrections caused by the interactions of the electrons with each other. It is important that in this form we do not have explicit sums over all electrons, which significantly simplifies calculations.

In the relativistic calculations the non-relativistic configurations are typically not used. However, sometimes one may need to find the average energy of the non-relativistic configuration. In the Appendix A we derive the necessary expressions for this case.

IV Numerical tests

We made four test calculations for very different systems. In the first two calculations for He I and B I, there was no core and we tested our method for the valence correlations. Then we applied our method for the highly charged ion Fe XVII, where there is a very strong central field, correlation corrections are rather small, and perturbation theory must be quite accurate. In this system we had core 1s21s^{2}, so we calculated core-valence correlation corrections as well as valence ones. Finally, we made calculations for Sc I, where valence 3d3d electrons have a large overlap with the core shell 3p63p^{6} and core-valence correlation corrections are as important as valence ones.

Table 1: Ground state binding energy of He I (in a.u.). CI calculations are made for three spaces: PP, P+QSP+Q_{S}, and P+QP+Q. ΔP+Q\Delta_{P+Q} is the difference from the CI result in the P+QP+Q space. Three variants of PT calculations are made based on the CI calculation in P+QSP+Q_{S} space: (a) determinant-based PT; (b) effective Hamiltonian with Hartree-Fock denominators (16); (c) effective Hamiltonian with corrected denominators (33). Experimental binding energy is given for comparison in the last column Kramida et al. (2021).
PP P+QSP+Q_{S} P+QP+Q PT NIST
(a) (b) (c) Ref. Kramida et al. (2021)
E(1s2)E(1s^{2}) 2.86262.8626 2.87002.8700 2.90102.9010 2.90212.9021 2.90642.9064 2.90312.9031 2.90342.9034
ΔP+Q\Delta_{P+Q} 0.03840.0384 0.03100.0310 0.00000.0000 0.0011-0.0011 0.0054-0.0054 0.0021-0.0021 0.0024-0.0024
Table 2: Ground state binding energy of B I (in a.u.). CI calculations are made for valence spaces PP and P~\tilde{P}, which included 3 and 4 lower shells respectively. Experimental binding energy is given for comparison in the last column Kramida et al. (2021).
PP P+QSP+Q_{S} P~\tilde{P} P~+Q~S\tilde{P}+\tilde{Q}_{S} NIST
HH HeffH_{\mathrm{eff}} HH HeffH_{\mathrm{eff}} Ref. Kramida et al. (2021)
E(P1/22)E({}^{2}P_{1/2}) 24.568324.5683 24.597624.5976 24.659524.6595 24.572124.5721 24.599924.5999 24.658124.6581 24.658124.6581
ΔNIST\Delta_{\mathrm{NIST}} 0.08980.0898 0.06050.0605 0.0014-0.0014 0.08600.0860 0.05820.0582 0.00000.0000 0.00000.0000
Table 3: Low-lying energy levels of Fe XVII in respect to the ground state (in cm-1). The subspace QSQ_{S} includes single excitations to virtual shells n=517n=5-17. The subspace QSQ_{S}^{\prime} in addition includes single excitations from the 1s1s shell. Effective Hamiltonians account for the respective double excitations. For each calculation we also give relative accuracy in percent.
Config. Level NIST CI(PP) CI(P+QS)emu{}_{\mathrm{emu}}(P+Q_{S}) CI(P+QS)emu{}_{\mathrm{emu}}(P+Q^{\prime}_{S})
Ref. Kramida et al. (2021) HH HeffH_{\mathrm{eff}} HH HeffH_{\mathrm{eff}} HeffH^{\prime}_{\mathrm{eff}}
2p62p^{6} S01{}^{1}S_{0} 0 0 0 0 0 0
2p53p2p^{5}3p S13{}^{3}S_{1} 60934506093450 60763706076370 0.28%-0.28\% 60835406083540 0.16%-0.16\% 60884056088405 0.08%-0.08\% 60956006095600 0.04%0.04\% 60950866095086 0.03%0.03\%
2p53p2p^{5}3p D23{}^{3}D_{2} 61216906121690 61050496105049 0.27%-0.27\% 61119336111933 0.16%-0.16\% 61173076117307 0.07%-0.07\% 61242156124215 0.04%0.04\% 61237096123709 0.03%0.03\%
2p53p2p^{5}3p D33{}^{3}D_{3} 61347306134730 61180106118010 0.27%-0.27\% 61250566125056 0.16%-0.16\% 61300676130067 0.08%-0.08\% 61371376137137 0.04%0.04\% 61366026136602 0.03%0.03\%
2p53p2p^{5}3p P11{}^{1}P_{1} 61438506143850 61272786127278 0.27%-0.27\% 61341936134193 0.16%-0.16\% 61393456139345 0.07%-0.07\% 61462836146283 0.04%0.04\% 61457726145772 0.03%0.03\%
2p53s2p^{5}3s 2o2^{o} 58494905849490 58307785830778 0.32%-0.32\% 58386795838679 0.18%-0.18\% 58429005842900 0.11%-0.11\% 58508235850823 0.02%0.02\% 58503305850330 0.01%0.01\%
2p53s2p^{5}3s 1o1^{o} 58647705864770 58462695846269 0.32%-0.32\% 58541095854109 0.18%-0.18\% 58583975858397 0.11%-0.11\% 58662605866260 0.03%0.03\% 58656785865678 0.02%0.02\%
2p53s2p^{5}3s 1o1^{o} 59608705960870 59421985942198 0.31%-0.31\% 59501035950103 0.18%-0.18\% 59543165954316 0.11%-0.11\% 59622445962244 0.02%0.02\% 59616015961601 0.01%0.01\%
2p53d2p^{5}3d P1o3{}^{3}P_{1}^{o} 64718006471800 64553066455306 0.25%-0.25\% 64620106462010 0.15%-0.15\% 64631496463149 0.13%-0.13\% 64698826469882 0.03%-0.03\% 64689626468962 0.04%-0.04\%
2p53d2p^{5}3d P2o3{}^{3}P_{2}^{o} 64864006486400 64700756470075 0.25%-0.25\% 64767386476738 0.15%-0.15\% 64778396477839 0.13%-0.13\% 64845316484531 0.03%-0.03\% 64836126483612 0.04%-0.04\%
2p53d2p^{5}3d F4o3{}^{3}F_{4}^{o} 64868306486830 64716306471630 0.23%-0.23\% 64785326478532 0.13%-0.13\% 64781296478129 0.13%-0.13\% 64850576485057 0.03%-0.03\% 64841476484147 0.04%-0.04\%
2p53d2p^{5}3d F3o3{}^{3}F_{3}^{o} 64930306493030 64775856477585 0.24%-0.24\% 64843386484338 0.13%-0.13\% 64843196484319 0.13%-0.13\% 64911016491101 0.03%-0.03\% 64901776490177 0.04%-0.04\%
2p53d2p^{5}3d D2o1{}^{1}D_{2}^{o} 65067006506700 64913836491383 0.24%-0.24\% 64980266498026 0.13%-0.13\% 64983606498360 0.13%-0.13\% 65050326505032 0.03%-0.03\% 65041016504101 0.04%-0.04\%
Table 4: Low-lying energy levels of Sc I (in cm-1). For each calculation we also give the differences with NIST Kramida et al. (2021) and the average absolute difference |Δ|av=1ki=1k|Δi||\Delta|_{\mathrm{av}}=\frac{1}{k}\sum_{i=1}^{k}|\Delta_{i}|. For the CI calculations in the P+QSP+Q_{S} space we use the emu CI approach Geddes et al. (2018) where we neglect non-diagonal matrix elements in the QSQ_{S} subspace. On the diagonal we use averaging over relativistic configurations, see Eq. (17).
Config. Level NIST CI(PP) CI(P+QS)emu{}_{\mathrm{emu}}(P+Q_{S})
Ref. Kramida et al. (2021) HH HH HeffH_{\mathrm{eff}}
EE EE Δ\Delta EE Δ\Delta EE Δ\Delta
3d4s23d4s^{2} D3/22{}^{2}D_{3/2} 0 0 0 0 0 0 0
D5/22{}^{2}D_{5/2} 168168 147147 21-21 157157 11-11 155155 13-13
3d24s3d^{2}4s F3/24{}^{4}F_{3/2} 1152011520 1494514945 34253425 73617361 4159-4159 1178611786 266266
F5/24{}^{4}F_{5/2} 1155811558 1496814968 34103410 74227422 4136-4136 1184711847 290290
F7/24{}^{4}F_{7/2} 1161011610 1500115001 33913391 74897489 4121-4121 1191411914 304304
F9/24{}^{4}F_{9/2} 1167711677 1504715047 33703370 75417541 4136-4136 1196311963 285285
3d24s3d^{2}4s F5/22{}^{2}F_{5/2} 1492614926 1736817368 24422442 1133111331 3595-3595 1566115661 735735
F7/22{}^{2}F_{7/2} 1504215042 1745517455 24132413 1145311453 3589-3589 1578115781 739739
3d24s3d^{2}4s D5/22{}^{2}D_{5/2} 1701317013 1997219972 29602960 1457414574 2439-2439 1747517475 462462
D3/22{}^{2}D_{3/2} 1702517025 1998019980 29552955 1460114601 2424-2424 1750017500 475475
3d24s3d^{2}4s P1/24{}^{4}P_{1/2} 1722617226 2032920329 31033103 1460614606 2620-2620 1747217472 246246
P3/24{}^{4}P_{3/2} 1725517255 2033920339 30843084 1467914679 2576-2576 1755217552 297297
P5/24{}^{4}P_{5/2} 1730717307 20380\quad 20380 30733073 14739\quad 14739 2568-2568 17606\quad 17606 299299
3d4s4p3d4s4p F3/2o4{}^{4}F_{3/2}^{o} 1567315673 1392113921 1751-1751 1601916019 346346 1587215872 200200
F5/2o4{}^{4}F_{5/2}^{o} 1575715757 1400214002 1754-1754 1609916099 342342 1595315953 197197
F7/2o4{}^{4}F_{7/2}^{o} 1588215882 1413914139 1743-1743 1621116211 330330 1606416064 183183
F9/2o4{}^{4}F_{9/2}^{o} 1602716027 1429014290 1737-1737 1634016340 314314 1619416194 168168
3d4s4p3d4s4p D1/2o4{}^{4}D_{1/2}^{o} 1601016010 1426514265 1745-1745 1631816318 308308 1644816448 438438
D3/2o4{}^{4}D_{3/2}^{o} 1602216022 1431114311 1711-1711 1635116351 329329 1651716517 495495
D5/2o4{}^{4}D_{5/2}^{o} 1614116141 1437514375 1766-1766 1640316403 262262 1655916559 418418
D7/2o4{}^{4}D_{7/2}^{o} 1621116211 1445814458 1753-1753 1650316503 292292 1662116621 410410
3d4s4p3d4s4p D3/2o2{}^{2}D_{3/2}^{o} 1602316023 1417214172 1851-1851 1651616516 493493 1644216442 419419
D5/2o2{}^{2}D_{5/2}^{o} 1609716097 1418914189 1907-1907 1652516525 428428 1644916449 352352
3d4s4p3d4s4p P1/2o4{}^{4}P_{1/2}^{o} 1850418504 1685416854 1650-1650 1852818528 2424 1852918529 2525
P3/2o4{}^{4}P_{3/2}^{o} 1851618516 1693016930 1586-1586 1853818538 2323 1854318543 2727
P5/2o4{}^{4}P_{5/2}^{o} 1857118571 1700717007 1565-1565 1857718577 66 1857218572 11
|Δ|av|\Delta|_{\mathrm{av}} 2247 1595 310

IV.1 Ground state of He I

Helium is the simplest system where correlation effects can be tested. We calculate the ground state energy, where correlation corrections are the largest. We choose the space PP to include shells n=13n=1\dots 3. The space QQ includes virtual shells s,p,ds,p,d with 4n204\leq n\leq 20. For this model problem, we can easily do CI in the whole space P+QP+Q thus producing the “exact” solution and compare these results with different variants of the perturbation theory discussed above. Results are listed in Table 1.

One can see that the valence CI provides accuracy on the order of 1%. The accuracy does not improve when we account for the single excitations to the virtual shells. However, when we include double excitations the agreement with the “exact” answer is significantly better. The determinant-based PT gives the best result. The results obtained with the effective Hamiltonian are less accurate, but corrections to the denominators reduce the discrepancy. Even the uncorrected variant of the MBPT is closer to the “exact” answer by an order of magnitude compared to the valence CI.

IV.2 Ground state of B I

B is a five electron system. The full CI calculation here is already very expensive. The determinant-based PT is also rather lengthy, so we made calculations only with the effective Hamiltonian and compared our results with the experiment Kramida et al. (2021). The effective radial integrals were calculated using the Hartree-Fock denominators. We tested two variants of the valence space: the first one, PP, included shells n=13n=1\dots 3 and the second one, P~\tilde{P}, included also the shell n=4n=4. Corresponding QQ and Q~\tilde{Q} spaces included s,p,d,f,gs,p,d,f,g shells up to n=20n=20. Results of these calculations for the ground state P1/22{}^{2}P_{1/2} are given in Table 2. We see that the accuracy of the CI calculation does not change much when we include an extra shell in the subspace PP. The accuracy of the CI calculation in the subspaces P+QSP+Q_{S} and P~+Q~S\tilde{P}+\tilde{Q}_{S} is only slightly better than similar calculation in the subspaces PP and P~\tilde{P}. Only including double excitations by means of the MBPT improves the agreement with the experiment by more than an order of magnitude.

IV.3 Spectrum of Fe XVII

Ten-electron ion Fe XVII plays an important role in astrophysics and plasma physics, see Ref. Kühn et al. (2020a) and references therein. The spectrum of this ion was calculated within several different approaches Kühn et al. (2020b) with relative accuracy of about 0.03%. Here we repeat these calculations using the new method. We use basis set [17spdfg][17spdfg]. Virtual orbitals starting from 4s4s and up are formed from B-splines using the method from Ref. Kozlov and Tupitsyn (2019). Valence subspace PP includes shells 2s,2p,3s,3p,3d,4s,4p,4d2s,2p,3s,3p,3d,4s,4p,4d, and 4f4f, while the 1s1s shell is frozen. Single excitations to all higher orbitals are included in the subspace QSQ_{S} and the subspace QSQ_{S}^{\prime} in addition includes single excitations from the 1s1s shell. We make two CI calculations in the spaces PP and P+QSP+Q_{S} respectively. Then we repeat these calculations using the effective Hamiltonian, which accounts for the excitations to the subspace QDQ_{D}. Finally, we make CI calculation in the P+QSP+Q_{S}^{\prime} for the effective Hamiltonian HeffH_{\mathrm{eff}}^{\prime} which accounts for the double excitations from 1s1s shell as well as for the double excitations to the virtual shells with n5n\geq 5. Results of all these calculations are given in Table 3.

One can see that already the CI calculation in the subspace PP is quite accurate here, the relative errors being about 0.3%. This is not surprising for such a strong central field. When we increase the size of the configuration space by adding single excitations to the virtual shells n=517n=5\dots 17 the errors substantially decrease but remain of the same order of magnitude. The same happens when we do CI for the effective Hamiltonian in the subspace PP. Only when we include both single and double excitations to the virtual shells by doing CI for the effective Hamiltonian in the subspace P+QSP+Q_{S} we increase the accuracy by an order of magnitude, the errors being 0.04% or less. Adding S and D excitations from the 1s1s shell leads to corrections to the transition energies within 0.01%0.01\%. Our final accuracy is similar to the accuracy obtained in Ref. Kühn et al. (2020b), where CI space included all double and some triple excitations to all virtual shells (the basis set there was different, but of the same length). In our present calculation, the size of the space P+QSP+Q_{S} is about 1.4 million determinants, and the size of the space P+QSP+Q_{S}^{\prime} is close to 2 million determinants, which is significantly less than the CI space of Ref. Kühn et al. (2020a).

IV.4 Spectrum of Sc I

The ground state configuration for Sc I is [Ar]3d14s2\mathrm{[Ar]}3d^{1}4s^{2} and lowest excited states belong to the configurations 3d24s3d^{2}4s and 3d4s4p3d4s4p. The 3d3d shell has a large overlap with the core shells 3s3s and 3p3p. Because of that frozen core approximation can not reproduce even the lowest part of the spectrum. Including 3s3s and 3p3p shells into the valence space makes its size extremely large. Therefore, this is a good system to apply our method.

We use a short basis set [9spdfgh][9spdfgh], which is constructed as described in Ref. Kozlov and Tupitsyn (2019). In the valence space PP, the shells n3n\leq 3 are closed and the virtual shells n8n\geq 8 and all hh orbitals are empty. The space QSQ_{S} includes single excitations from the upper core shells n=3n=3 and single excitations to the virtual shells. We keep core shells up to n2n\leq 2 frozen on all stages. Results of the calculation of the spectrum are presented in Table 4, where excitation energies from the ground state in cm -1 are shown for approximately 10 lower levels of each parity. The sizes of the valence space PP and P+QSP+Q_{S} are about 6×1046\times 10^{4} and 1×1061\times 10^{6} determinants respectively. We list the results of three calculations: the full CI in the valence space PP and emu CI Geddes et al. (2018) in the space P+QSP+Q_{S} for the bare and the effective Hamiltonians. The effective radial integrals were calculated with the Hartree-Fock denominators. For each of these calculations we also give differences from the experimental values Kramida et al. (2021) and the averaged absolute difference.

One can see that all the levels in the CI calculation are shifted from their experimental energies: the levels of the configuration 3d24s3d^{2}4s lie higher by 3 thousand inverse centimeters, while the levels of the configuration 3d4s4p3d4s4p lie lower by 2 thousand inverse centimeters. The picture changes drastically when we add single excitations and solve the problem in the space P+QSP+Q_{S}. Now the levels of the configuration 3d24s3d^{2}4s lie lower by 3 thousand inverse centimeters, while the levels of the configuration 3d4s4p3d4s4p are almost in place. Finally, when we use the effective Hamiltonian, which accounts for the double excitations, the levels get closer to their places with the average deviation about 300 cm-1, or 7 times smaller, than for the CI calculation.

In this test calculation, we used a rather short basis set and were probably rather far from saturation. Therefore we can not reliably estimate the ultimate accuracy of the method for scandium. Looking at the results we see that the size of the PT corrections is very large and there is also large cancellation between contributions of the single and double excitations. Therefore it is unlikely that converged results would be significantly better than what we got here. On the other hand, we see systematic improvement in our final results compared to the pure valence calculation. It is also worth mentioning that if one would try to include all double excitations in CI calculation the size of the configuration space would be much above 1×1081\times 10^{8} even for the basis set as short as this one.

V Conclusions

We suggest a new version of the CI+MBPT method Dzuba et al. (1996) with the different division of the many-electron space into parts where non-perturbative and perturbative methods are used. This new division may be more practical for the atoms with many valence electrons, where the size of the valence space may be too big for solving the matrix eigenvalue problem. This method can be used in the all-electron calculations for light atoms as well as for the calculations with the frozen core. In the latter case, the single and double excitations from (some of) the core shells can be treated perturbatively. We ran four rather different tests which showed systematic one-order-of-magnitude improvement of the results when we added MBPT corrections to the CI calculations.

Acknowledgements

We thank Marianna Safronova, Charles Cheung, and Sergey Porsev for their constant interest in this work and very useful discussions. This work was supported by the Russian Science Foundation (Grant No. 19-12-00157). I.I.T. acknowledges the support from the Resource Center “Computer Center of SPbU”, St. Petersburg, Russia.

Appendix A Average energy of the non-relativistic configuration

In the average over non-relativistic configuration (LS-average) Tupitsyn et al. (2018); Lindgren and Rosén (1975), the occupation numbers for the relativistic orbitals qaq_{a} may be non-integer, while occupation numbers for non-relativistic orbitals qAq_{A} are still integer (we use capital letters A,B,M,NA,B,M,N to designate non-relativistic orbitals). Below we show that properly defining one-electron integrals IAI_{A} and two-electron matrix elements UABU_{AB} we obtain expressions similar to Eqs. (3233).

The average energy of the non-relativistic configuration RR can be written as:

E¯R=aq~aIa+12aq~a(q~awa)F0(a,a)+a<bq~aq~bwabF0(a,b)+a,k>0q~a(q~awa)faakFk(a,a)+a<b,kq~aq~bwabgabkGk(a,b),\bar{E}_{R}=\sum_{a}\tilde{q}_{a}\,I_{a}+\frac{1}{2}\sum_{a}\tilde{q}_{a}\,(\tilde{q}_{a}-w_{a})\,F^{0}(a,a)+\sum_{a<b}\tilde{q}_{a}\,\tilde{q}_{b}\,w_{ab}\,F^{0}(a,b)\\ +\sum_{a,k>0}\tilde{q}_{a}(\tilde{q}_{a}-w_{a})\,f^{k}_{aa}\,F^{k}(a,a)+\sum_{a<b,k}\tilde{q}_{a}\,\tilde{q}_{b}\,w_{ab}\,g^{k}_{ab}\,G^{k}(a,b)\,, (34)

where

q~a=2ja+14la+2qA,wa=qAq~a+2ja4la+1,wab={4la+24la+1qA1qA,A=B,1,AB.\displaystyle\tilde{q}_{a}=\displaystyle\frac{2j_{a}+1}{4l_{a}+2}\,q_{A}\,,\qquad w_{a}=\displaystyle\frac{q_{A}-\tilde{q}_{a}+2j_{a}}{4l_{a}+1}\,,\qquad w_{ab}=\left\{\begin{array}[]{cll}\frac{4l_{a}+2}{4l_{a}+1}\,\frac{q_{A}-1}{q_{A}}\,,&\,A=B\,,\\ 1\,,&\,A\neq B\,.\end{array}\right. (37)

Using expressions:

q~a(q~awa)=2ja+14la+22ja4la+1qA(qA1),\displaystyle\tilde{q}_{a}(\tilde{q}_{a}-w_{a})=\frac{2j_{a}+1}{4l_{a}+2}\,\frac{2j_{a}}{4l_{a}+1}\,q_{A}(q_{A}-1)\,, wabq~aq~b=2ja+14la+22ja+14la+1qA(qA1),A=B,jaja,\displaystyle w_{ab}\,\tilde{q}_{a}\tilde{q}_{b}=\frac{2j_{a}+1}{4l_{a}+2}\,\frac{2j_{a^{\prime}}+1}{4l_{a}+1}\,q_{A}(q_{A}-1)\,,\quad A=B,\,\,j_{a}\neq j_{a^{\prime}}\,, (38)

we rewrite the equation (34) in the form

E¯R=AqAja2ja+14la+2Ia+12AqA(qA1)a,aA(2ja+1)(2ja+1δa,a)(4la+2)(4la+1)F0(a,a)++12ABqAqBaA,bB(2ja+1)(2jb+1)(4la+2)(4lb+2)[F0(a,b)+kgabkGk(a,b)]+12AqA(qA1)a,aAk>02ja+14la+22ja+14la+1gaakGk(a,a).\bar{E}_{R}=\sum_{A}q_{A}\,\sum_{j_{a}}\frac{2j_{a}+1}{4l_{a}+2}I_{a}+\frac{1}{2}\sum_{A}q_{A}\,(q_{A}-1)\,\sum_{a,a^{\prime}\in A}\frac{(2j_{a}+1)(2j_{a^{\prime}}+1-\delta_{a,a^{\prime}})}{(4l_{a}+2)(4l_{a}+1)}\,F^{0}(a,a^{\prime})+\\ +\frac{1}{2}\sum_{A\neq B}q_{A}\,q_{B}\sum_{a\in A,b\in B}\frac{(2j_{a}+1)(2j_{b}+1)}{(4l_{a}+2)(4l_{b}+2)}\,\left[F^{0}(a,b)+\sum_{k}g^{k}_{ab}\,G^{k}(a,b)\right]\\ +\frac{1}{2}\sum_{A}\,q_{A}(q_{A}-1)\sum_{a,a^{\prime}\in A}\sum_{k>0}\,\frac{2j_{a}+1}{4l_{a}+2}\,\frac{2j_{a^{\prime}}+1}{4l_{a}+1}\,g^{k}_{aa^{\prime}}\,G^{k}(a,a^{\prime})\,. (39)

In the last sum, the term k=0k=0 is absent since jajaj_{a}\neq j_{a^{\prime}} and k|jaja|k\geq|j_{a}-j_{a^{\prime}}|. Now we can introduce non-relativistic analogues of the integrals IaI_{a} and matrix elements UabU_{ab} and rewrite Eq. (34) alike Eq. (17):

E¯R\displaystyle\bar{E}_{R} =AqAIA+12AqA(qA1)UAA+12ABqAqBUAB,\displaystyle=\sum_{A}q_{A}\,I_{A}+\frac{1}{2}\,\sum_{A}q_{A}\,(q_{A}-1)\,U_{AA}+\frac{1}{2}\,\sum_{A\neq B}\,q_{A}\,q_{B}\,U_{AB}\,, (40)
IA\displaystyle I_{A} =aA2ja+14la+2Ia,\displaystyle=\sum_{a\in A}\frac{2j_{a}+1}{4l_{a}+2}\,I_{a}\,, (41)
UAA\displaystyle U_{AA} =a,aA(2ja+1)(2ja+1)(4la+2)(4la+1)[F0(a,a)+k>0gaakGk(a,a)]aA(2ja+1)(4la+2)(4la+1)F0(a,a),\displaystyle=\sum_{a,a^{\prime}\in A}\frac{(2j_{a}+1)(2j_{a^{\prime}}+1)}{(4l_{a}+2)(4l_{a}+1)}\,\left[F^{0}(a,a^{\prime})+\sum_{k>0}g^{k}_{aa^{\prime}}\,G^{k}(a,a^{\prime})\right]-\sum_{a\in A}\frac{(2j_{a}+1)}{(4l_{a}+2)(4l_{a}+1)}\,F^{0}(a,a)\,, (42)
UAB\displaystyle U_{AB} =aA,bB(2ja+1)(2jb+1)(4la+2)(4lb+2)[F0(a,b)+kgabkGk(a,b)].\displaystyle=\sum_{a\in A,b\in B}\frac{(2j_{a}+1)(2j_{b}+1)}{(4l_{a}+2)(4l_{b}+2)}\,\left[F^{0}(a,b)+\sum_{k}g^{k}_{ab}\,G^{k}(a,b)\right]\,. (43)

Using Eq. (40) we get the following derivatives by analogy with Eqs. (2930):

E¯RqA=IA+BqBUAB12UAA,\displaystyle\frac{\partial\bar{E}_{R}}{\partial q_{A}}=I_{A}+\sum_{B}q_{B}\,U_{AB}-\frac{1}{2}\,U_{AA}\,, 2E¯RqAqB\displaystyle\frac{\partial^{2}\bar{E}_{R}}{\partial q_{A}\partial q_{B}} =UAB.\displaystyle=\displaystyle U_{AB}\,. (44)

The difference in energy between two configurations is

ΔE¯\displaystyle\Delta\bar{E} =AIAδqA+A(qA12)UAAδqA+BAqBUABδqB+12A,BUABδqAδqB.\displaystyle=\sum_{A}I_{A}\,\delta q_{A}+\sum_{A}\left(q_{A}-\frac{1}{2}\right)\,U_{AA}\,\delta q_{A}+\sum_{B\neq A}q_{B}\,U_{AB}\,\delta q_{B}+\frac{1}{2}\sum_{A,B}U_{AB}\,\delta q_{A}\,\delta q_{B}\,. (45)

This equation allow us to find the energy of a double excitation δqA=1,δqB=1,δqN=1,δqM=1\delta q_{A}=-1,\,\delta q_{B}=-1,\,\delta q_{N}=1,\,\delta q_{M}=1:

ΔE¯(ABNM)=IA+IBIMIN+CqC(UAC+UBCUMCUNC)UAAUBBUABUNM+UAN+UBN+UAM+UBM.\Delta_{\bar{E}}(AB\to NM)=I_{A}+I_{B}-I_{M}-I_{N}+\sum_{C}q_{C}\,(U_{AC}+U_{BC}-U_{MC}-U_{NC})\\ -U_{AA}-U_{BB}-U_{AB}-U_{NM}+U_{AN}+U_{BN}+U_{AM}+U_{BM}\,. (46)

If we introduce an averaged one-electron energy by analogy with (32) we can rewrite (46) as

εA\displaystyle\varepsilon_{A} =IA+BqBUAB(1δqA,0)UAA,\displaystyle=I_{A}+\sum_{B}q_{B}\,U_{AB}-(1-\delta_{q_{A},0})\,U_{AA}\,, (47)
ΔE¯(ABNM)\displaystyle\Delta_{\bar{E}}(AB\to NM) =εA+εBεMεNUABUNM+UAN+UBN+UAM+UBM.\displaystyle=\varepsilon_{A}+\varepsilon_{B}-\varepsilon_{M}-\varepsilon_{N}-U_{AB}-U_{NM}+U_{AN}+U_{BN}+U_{AM}+U_{BM}\,. (48)

We obtained corrections to the standard MBPT energy denominator (16) using two approximations. Averaging over relativistic configurations gives expression (33) and averaging over non-relativistic configurations leads to expression (48). These expressions differ only by the definitions of the one-electron energies and two-electron matrix elements.

References

  • Desclaux (1975) J. P. Desclaux, Computer Physics Communications 9, 31 (1975).
  • Jönsson et al. (2013) P. Jönsson, G. Gaigalas, J. Bieroń, C. F. Fischer, and I. P. Grant, 184, 2197 (2013).
  • Froese Fischer et al. (2016) C. Froese Fischer, M. Godefroid, T. Brage, P. Jönsson, and G. Gaigalas, Journal of Physics B Atomic Molecular Physics 49, 182004 (2016).
  • Gu (2008) M. F. Gu, Canadian Journal of Physics 86, 675 (2008).
  • Fritzsche et al. (2002) S. Fritzsche, C. F. Fischer, and G. Gaigalas, Computer Physics Communications 148, 103 (2002).
  • Jiang et al. (2016) J. Jiang, J. Mitroy, Y. Cheng, and M. W. J. Bromley, Phys. Rev. A 94, 062514 (2016).
  • Tupitsyn et al. (2003) I. I. Tupitsyn, V. M. Shabaev, J. R. Crespo López-Urrutia, I. Draganić, R. S. Orts, and J. Ullrich, Phys. Rev. A 68, 022511 (2003).
  • Tupitsyn et al. (2005) I. I. Tupitsyn, A. V. Volotka, D. A. Glazov, V. M. Shabaev, G. Plunien, J. R. C. López-Urrutia, A. Lapierre, and J. Ullrich, Phys. Rev. A 72, 062503 (2005).
  • Dzuba et al. (1987) V. A. Dzuba, V. V. Flambaum, P. G. Silvestrov, and O. P. Sushkov, J. Phys. B 20, 1399 (1987).
  • Blundell et al. (1987) S. A. Blundell, D. S. Guo, W. R. Johnson, and J. Sapirstein, Atomic Data and Nuclear Data Tables 37, 103 (1987).
  • Sapirstein (1998) J. Sapirstein, Reviews of Modern Physics 70, 55 (1998).
  • Dzuba et al. (1996) V. A. Dzuba, V. V. Flambaum, and M. G. Kozlov, Phys. Rev. A 54, 3948 (1996).
  • Kahl and Berengut (2019) E. Kahl and J. Berengut, Computer Physics Communications 238, 232 (2019), eprint arXiv:1805.11265.
  • Kozlov et al. (2015) M. G. Kozlov, S. G. Porsev, M. S. Safronova, and I. I. Tupitsyn, Comput. Phys. Commun. 195, 199 (2015), ISSN 0010-4655.
  • Eliav and Kaldor (2010) E. Eliav and U. Kaldor, Relativistic Four-Component Multireference Coupled Cluster Methods: Towards A Covariant Approach (Springer Netherlands, Dordrecht, 2010), pp. 113–144, ISBN 978-90-481-2885-3.
  • Saue et al. (2020) T. Saue, R. Bast, A. S. P. Gomes, H. J. A. Jensen, L. Visscher, I. A. Aucar, R. Di Remigio, K. G. Dyall, E. Eliav, E. Fasshauer, et al., J. Chem. Phys.  152, 204104 (2020), eprint arXiv:2002.06121.
  • Blundell et al. (1989) S. A. Blundell, W. R. Johnson, Z. W. Liu, and J. Sapirstein, Phys. Rev. A 40, 2233 (1989).
  • Safronova and Johnson (2008) M. S. Safronova and W. R. Johnson, Advances in Atomic Molecular and Optical Physics 55, 191 (2008).
  • Oleynichenko et al. (2020) A. V. Oleynichenko, A. Zaitsevskii, L. V. Skripnikov, and E. Eliav, Symmetry 12 (2020).
  • Johnson (2007) W. R. Johnson, Atomic Structure Theory. Lectures on Atomic Physics (Springer Berlin Heidelberg, 2007).
  • Flambaum and Ginges (2005) V. V. Flambaum and J. S. M. Ginges, Phys. Rev. A 72, 052115 (2005).
  • Ginges and Berengut (2016) J. S. M. Ginges and J. C. Berengut, Phys. Rev. A 93, 052509 (2016).
  • Shabaev et al. (2013) V. M. Shabaev, I. I. Tupitsyn, and V. A. Yerokhin, Phys. Rev. A 88, 012513 (2013).
  • Tupitsyn et al. (2016) I. I. Tupitsyn, M. G. Kozlov, M. S. Safronova, V. M. Shabaev, and V. A. Dzuba, Phys. Rev. Lett. 117, 253001 (2016).
  • Dzuba et al. (2017) V. A. Dzuba, J. C. Berengut, C. Harabati, and V. V. Flambaum, Phys. Rev. A 95, 012503 (2017).
  • Geddes et al. (2018) A. J. Geddes, D. A. Czapski, E. V. Kahl, and J. C. Berengut, Physical Review A 98, 042508 (2018), eprint arXiv:1805.06615.
  • Dzuba et al. (2019) V. A. Dzuba, V. V. Flambaum, and M. G. Kozlov, Phys. Rev. A 99, 032501 (2019), eprint ArXiv:1812.09826.
  • Imanbaeva and Kozlov (2019) R. T. Imanbaeva and M. G. Kozlov, Annalen der Physik 531, 1800253 (2019).
  • Cheung et al. (2020) C. Cheung, M. S. Safronova, S. G. Porsev, M. G. Kozlov, I. I. Tupitsyn, and A. I. Bondarev, Physical Review Letters 124, 163001 (2020), eprint arXiv:1912.08714.
  • Li and Dzuba (2020) J. Li and V. Dzuba, Journal of Quantitative Spectroscopy and Radiative Transfer 247, 106943 (2020), eprint arXiv:2002.08033.
  • Mann (1973) J. B. Mann, Atomic Data and Nuclear Data Tables 12, 1 (1973).
  • Grant (1970) I. P. Grant, Advances in Physics 19, 747 (1970).
  • Kramida et al. (2021) A. Kramida, Y. Ralchenko, J. Reader, and NIST ASD Team, Nist atomic spectra database (version 5.9) (2021), URL https://physics.nist.gov/asd.
  • Kühn et al. (2020a) S. Kühn, C. Shah, J. R. C. López-Urrutia, K. Fujii, R. Steinbrügge, J. Stierhof, M. Togawa, Z. Harman, N. S. Oreshkina, C. Cheung, et al., Physical Review Letters 124, 225001 (2020a), eprint arXiv:1911.09707.
  • Kühn et al. (2020b) S. Kühn et al., High resolution photoexcitation measurements exacerbate the long-standing Fe XVII oscillator strength problem: Supplemental material (2020b), URL 10.1103/physrevlett.124.225001.
  • Kozlov and Tupitsyn (2019) M. Kozlov and I. Tupitsyn, Atoms 7, 92 (2019).
  • Tupitsyn et al. (2018) I. I. Tupitsyn, N. A. Zubova, V. M. Shabaev, G. Plunien, and T. Stöhlker, Physical Review A 98, 022517 (2018), eprint arXiv:1804.09986.
  • Lindgren and Rosén (1975) I. Lindgren and A. Rosén, in Case Studies in Atomic Physics, edited by E. McDaniel and M. McDowell (Elsevier, 1975), pp. 93 – 196, ISBN 978-0-7204-0331-2.