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

Parity-mixed coupled-cluster formalism for computing parity-violating amplitudes

H. B. Tran Tan    Di Xiao    A. Derevianko Department of Physics, University of Nevada, Reno, Nevada 89557, USA
Abstract

We formulate a parity-mixed coupled-cluster (PM-CC) approach for high-precision calculations of parity non-conserving amplitudes in mono-valent atoms. Compared to the conventional formalism which uses parity-proper (PP) one-electron orbitals, the PM-CC method is built using parity-mixed (PM) orbitals. The PM orbitals are obtained by solving the Dirac-Hartree-Fock equation with the electron-nucleus electroweak interaction included (PM-DHF). There are several advantages to such a PM-CC formulation: (i) reduced role of correlations, as for the most experimentally-accurate to date Cs133{}^{133}\mathrm{Cs} 6S1/27S1/26S_{1/2}-7S_{1/2} transition, the PM-DHF result is only 3% away from the accurate many-body value, while the conventional DHF result is off by 18%; (ii) avoidance of directly summing over intermediate states in expressions for parity non-conserving amplitudes which reduces theoretical uncertainties associated with highly-excited and core-excited intermediate states, and (iii) relatively straightforward upgrade of existing and well-tested large-scale PP-CC codes. We reformulate the CC method in terms of the PM-DHF basis and demonstrate that the cluster amplitudes are complex numbers with opposite parity real and imaginary parts. We then use this fact to map out a strategy through which the new PM-CC scheme may be implemented.

I Introduction

The field of parity violation started with the seminal paper by Lee and Yang [1] and the discovery of parity non-conservation (PNC) in nuclear β\beta-decay [2]. Shortly after, the possibility of measuring atomic parity violation (APV) as a low-energy test for the Standard Model (SM) was investigated by Zel’dovich [3], whose consideration for hydrogen suggested that the effects were too small to be observable. The situation changed when the Bouchiats [4, 5, 6] demonstrated that APV effects scale as Z3Z^{3}, where ZZ is the nuclear charge, thus reopening the case for observing them in heavy neutral atoms. Following a proposal by Khriplovich [7], APV effects were first observed in bismuth by Barkov and Zolotorev [8]. Following this discovery, several APV experiments were performed for cesium [9, 10, 11, 12, 13], bismuth [14], lead [15, 16], thallium [17, 18] and ytterbium [19]. New APV experiments are underway or in the planning stage [20, 21, 22, 23, 24, 25, 26, 27] (see also the review [28] and references therein), with the aim of attaining a 0.1%\sim 0.1\% accuracy in 133Cs.

The APV measurements are usually interpreted in terms of the nuclear weak charge QWQ_{W}, which is related to the measured PNC amplitude, EPVE_{\rm PV}, via EPV=kPVQWE_{\rm PV}=k_{\rm PV}Q_{W}, where kPVk_{\rm PV} is an atomic-structure factor. One wishes to compare the experimentally obtained value of QWQ_{W} with the value predicted by the SM. For this purpose, the quantity kPVk_{\rm PV} should be known with a better accuracy than that of the amplitude EPVE_{\rm PV}, thus yielding an accurate estimate of QWQ_{W}. This approach has been so far most successful in 133Cs, due to its large nuclear charge, Z=55Z=55, and its relatively simple atomic structure with one valence electron above a closed Xe-like core [29]. Part of the success was also due to the fact that 133Cs is used in the primary frequency standard, providing a wealth of information on its basic atomic properties.

In Cs133{}^{133}\mathrm{Cs}, the experimental uncertainty for the 6S1/27S1/26S_{1/2}-7S_{1/2} PNC amplitude eventually reached 0.35% [10]. The most accurate theoretical computations for the atomic-structure factor kPVk_{\rm PV} was built upon the relativistic Many-Body Perturbation Theory (MBPT), a systematic order-by-order approach which includes electron correlations. Certain classes of MBPT diagrams can be summed to all orders, taking advantage of the underlying topology of the diagrams. In the late 1980s and early 1990s, the accuracy of MBPT calculations for kPVk_{\rm PV} was estimated to be at the level of 1% [30, 31, 32, 33].

A later re-analysis [34], based on an improved theory-experiment agreement with the new measurements of atomic properties, reduced the theoretical uncertainty of Refs. [30, 31, 32, 33] to the level of 0.4%. The deduced value for the Cs133{}^{133}\mathrm{Cs} weak charge differed by 2.5σ2.5\sigma from the SM prediction, thus suggesting new physics beyond the SM [35, 36, 37, 38]. However, the inclusion of Breit [39, 40, 41] and QED radiative corrections [42, 43, 44, 45] brought the Cs133{}^{133}\mathrm{Cs} result back into the essential 1σ1\sigma agreement with the SM. Clearly, the answer to whether the Cs133{}^{133}\mathrm{Cs} PNC result confirms the SM or hints at new physics very much depends on the quality of theoretical atomic calculation for the atomic-structure factor kPVk_{\rm PV}. In the works [39, 40, 41, 42, 43, 45], the theoretical uncertainty stood at 0.5%0.5\%, still larger than the 0.35%0.35\% experimental error bar.

Since the early 2000s, the theoretical error bar has been (and still is) dominated by the uncertainty of solving the basic many-body problem of atomic structure. Further progress in improving the theoretical accuracy was reported in the late 2000s [46, 47]. These calculations built upon the ab initio relativistic coupled-cluster (CC) scheme [32]. While the calculations of Refs. [31, 32, 33] were complete through the third order of MBPT for matrix elements, the scheme in Refs. [46, 47] was complete through the fourth order of MBPT. References [46, 47] have reduced the theoretical uncertainty in the Cs133{}^{133}\mathrm{Cs} atomic-structure factor kPVk_{\rm PV} to 0.27%. The final value of the Cs133{}^{133}\mathrm{Cs} weak charge extracted from this calculation was in an agreement with the SM prediction, placing strong constraints on a variety of new physics scenarios.

The works [32, 33, 46, 47] used a sum-over-states approach to calculate the PNC 6S1/27S1/26S_{1/2}-7S_{1/2} transition amplitude in Cs133{}^{133}\mathrm{Cs}

EPV\displaystyle E_{\rm PV} =n[6S1/2|HW|nP1/2nP1/2|Dz|7S1/2E6S1/2EnP1/2\displaystyle=\sum_{n}\left[\frac{\bra{6S_{1/2}}H_{W}\ket{nP_{1/2}}\bra{nP_{1/2}}{D}_{z}\ket{7S_{1/2}}}{E_{6S_{1/2}}-E_{nP_{1/2}}}\right. (1)
+6S1/2|Dz|nP1/2nP1/2|HW|7S1/2E7S1/2EnP1/2].\displaystyle\left.+\frac{\bra{6S_{1/2}}{D}_{z}\ket{nP_{1/2}}\bra{nP_{1/2}}H_{W}\ket{7S_{1/2}}}{E_{7S_{1/2}}-E_{nP_{1/2}}}\right]\,.

In this second-order expression, |nLJ\ket{nL_{J}} stands for various states of the Cs133{}^{133}\mathrm{Cs} atom, with nn being the principal quantum number, LL the orbital angular momentum, and JJ the total angular momentum. These are the true many-body eigen-states of the parity-proper (PP) atomic Hamiltonian. Further, DzD_{z} is the zz-component of the electric dipole operator 𝐃i𝐝i=ie𝐫i{\bf D}\equiv\sum_{i}\mathbf{d}_{i}=-\sum_{i}e{\bf r}_{i} and HW=ihW(i)H_{W}=\sum_{i}h_{W}(i) is the PP-odd electron-nucleus weak interaction with the single-electron operator hWh_{W} having the form

hW(i)GF22QWγ5ρ(ri).h_{W}(i)\equiv-\frac{G_{F}}{2\sqrt{2}}Q_{W}\gamma_{5}\rho(r_{i})\,. (2)

Here, GF=2.2225×1014G_{F}=2.2225\times 10^{-14} a.u. is the Fermi constant of the weak interaction, QWQ_{W} is the weak nuclear charge, ρ(r)\rho(r) is the nuclear neutron density (see Ref. [48] for a discussion of neutron skin effects) and γ5\gamma_{5} is the conventional Dirac matrix.

The largest contributions to EPVE_{\rm PV} in the sum-over-states expression (1) come from terms with n=6,7,8,9n=6,7,8,9 (the “main” contribution). In Refs. [46, 47], the required many-body states were computed using the CC approximation including singles, doubles and valence triples (CCSDvT). The computed CCSDvT wave functions were subsequently used to compute the dipole and weak interaction matrix elements entering Eq. (1). Residual contributions to Eq. (1) come from intermediate states with n10n\geq 10 (the “tail” contribution) and core-excited states. These residual contributions are sub-dominant and were evaluated using less accurate methods, having an estimated uncertainty of 10%10\%.

In a later work [49], the value of the residual contributions was reevaluated and Ref. [49] claimed a contribution of core-excited states to EPVE_{\rm PV} having an opposite sign as compared to the analyses of both Refs. [32, 33] and Refs. [46, 47]. The Cs133{}^{133}\mathrm{Cs} weak charge extracted from the revised atomic-structure factor is 1.5σ1.5\sigma away from the SM value, thus relaxing Refs. [46, 47] constraints on new physics. In addition, Ref. [49] raised the theoretical uncertainty in the atomic structure factor kPVk_{\rm PV} back to 0.5%, above the experimental error bar on EPVE_{\rm PV}.

The latest Dalgarno-Lewis-type coupled-cluster computations [50] support both the sign and the value of the core-excited state contributions of Refs. [46, 47]. However, as of now, a clear understanding of why the two approaches, Refs. [49] and [46, 47], lead to core-excited state contribution of opposite signs is still lacking. Furthermore, objections [51] have been raised with regards to the error estimates of Ref. [50].

It may be observed that the disagreement between Refs. [46, 47] and [49] arose due to the artificial separation into the “main” and “tail” contributions characteristic of the sum-over-states method [29, 28]. In this paper, we seek to directly include the weak interaction into the single-particle atomic Hamiltonian, thus avoiding this artificial separation, and treat all the intermediate states on equal high-precision footing. In this approach, the single-electron eigen-states of the modified Hamiltonian will already have a parity-mixed (PM) character. The MBPT calculations of the PM many-body wave functions |6S1/2\ket{6S_{1/2}^{\prime}} and |7S1/2\ket{7S_{1/2}^{\prime}} can be carried out in a conventional fashion using this PM singe-electron basis. This PM approach was first suggested in Ref. [52] and carried through to all second-order MBPT corrections in Ref. [53]. In this paper, we extend it to a more-complete CC method. Once the PM many-body states are computed, the PNC amplitude can be expressed simply as,

EPV=6S1/2|Dz|7S1/2,E_{\rm PV}=\bra{6S_{1/2}^{\prime}}D_{z}\ket{7S_{1/2}^{\prime}}\,, (3)

avoiding the summation over intermediate states altogether.

In addition, the lowest-order Dirac-Hartree-Fock (DHF) result in this PM approach is only 3% away from the more accurate CCSDvT value. This is to be compared with the traditional parity-proper (PP) DHF result which is off by 18%. This indicates that the correlation corrections in the PM approach are substantially smaller than in the conventional PP method. Depending on the MBPT convergence pattern, one can generically anticipate an improved theoretical accuracy.

Another important point is that in the sum-over-states approach employed in Refs. [46, 47], the theoretical uncertainty budget of EPVE_{\mathrm{PV}} included comparable contributions of the accurately computed low-lying states (in the CCSDvT approach) and of the less-accurate highly-excited and core-excited states. Our method would allow us to treat all of these contributions on the same high-accuracy CCSDvT footing, thus improving the overall theoretical uncertainty even without the potentially reduced role of the correlation corrections.

To follow this program through in the context of the CC method, one requires a numerically complete set of PM orbitals (single-particle states) {ψi}\{\psi^{\prime}_{i}\}. Generating such PM basis sets and quantifying their numerical accuracy is one of the goals of this paper. A PM basis set has to be obtained in the modified DHF potential of the Xe-like core which includes the weak interaction (2). Considering the increased numerical accuracy demanded of the quality of basis sets, we employ the dual-kinetic-balance B-splines basis sets [54, 55] which are more numerically robust and have the correct behaviour inside the finite nucleus compared to the B-spline basis sets originally used by the Notre Dame group [56].

Once a PM basis set is obtained, one may proceed to computing matrix elements of various operators such as the one-body dipole operator zijz^{\prime}_{ij} and the two-body inter-electron Coulomb interaction gijklg^{\prime}_{ijkl} in the new PM-DHF basis. With these computed matrix elements, the MBPT and CCSDvT expressions can be evaluated. As we will show, all the matrix elements in the PM basis can be decomposed into real and imaginary parts with opposite parities (with the conventional choice of radial wave functions being real-valued). Then all the information about opposite-parity admixtures is contained in the imaginary parts of various MBPT expressions. This greatly simplifies the formalism and only requires only minor modifications to already developed and tested MBPT codes. We demonstrate the utility of this technique for the random-phase approximation (RPA) subset of MBPT diagrams and discuss a strategy for applying these ideas in the more-complete CC calculations.

The paper is organized as follows. In Sec. II, we briefly present the set-up of our problem. Although this section does not contain new results, it serves as a starting point for our main discussion and a mean to define our notations. We will also derive, in Sec. III, the second-quantized form of the parity operator which will be useful in deriving selection rules for our PM-CC method. Section III is followed by Sec. IV, where we present several methods through which a basis of PM single-electron orbitals may be obtained. In Sec. V, we present the PM matrix elements of one- and two-body operators computed using the obtained PM single-electron orbitals. In Sec. VI, we illustrate how these PM matrix elements can be used in an RPA calculation of the PNC amplitude. The generalization to the CC method is discussed in Sec. VII. Finally, Sec. VIII draws conclusions and presents an outlook for our future work. The paper contains several appendixes which provide further technical details. Unless specified otherwise, the atomic units, |e|=me==1|e|=m_{e}=\hbar=1, are used.

II Theory

In this section, we lay out the theoretical framework for computing the PNC amplitude in an atom. The material presented here is not new but serves as a starting point and a mean to define our notations.

Let us begin by considering the Hamiltonian of the atomic electrons propagating in the combined PP Coulomb potential iVnuc(ri)\sum_{i}V_{\rm nuc}(r_{i}) and the PP-odd electron-nucleus interaction HW=ihW(i)H_{W}=\sum_{i}h_{W}(i). Here, ii labels all the atomic electrons. The full electronic Hamiltonian HH^{\prime} may be decomposed into

H\displaystyle H^{\prime} =ih0(i)+Vc,\displaystyle=\sum_{i}h^{\prime}_{0}(i)+V^{\prime}_{c}\,, (4)
h0(i)\displaystyle h^{\prime}_{0}(i) =c𝜶i𝐩i+mec2βi\displaystyle=c\boldsymbol{\alpha}_{i}\cdot{\bf p}_{i}+m_{e}c^{2}\beta_{i}
+Vnuc(ri)+hW(ri)+U(ri),\displaystyle+V_{\rm nuc}(r_{i})+h_{W}(r_{i})+U^{\prime}(r_{i})\,,
Vc\displaystyle V^{\prime}_{c} =12ije2|𝐫i𝐫j|iU(ri),\displaystyle=\frac{1}{2}\sum_{i\neq j}\frac{e^{2}}{|{\bf r}_{i}-{\bf r}_{j}|}-\sum_{i}U^{\prime}(r_{i})\,,

where U(ri)U^{\prime}(r_{i}) is some single-electron potential to be specified later. We use the prime on h0h^{\prime}_{0} and VcV^{\prime}_{c} to distinguish them from the PP Hamiltonian h0=c𝜶i𝐩i+mec2βi+Vnuc(ri)+U(ri)h_{0}=c\boldsymbol{\alpha}_{i}\cdot{\bf p}_{i}+m_{e}c^{2}\beta_{i}+V_{\rm nuc}(r_{i})+U(r_{i}) and the PP eee-e interaction Vc=ije2/(2|𝐫i𝐫j|)iU(ri)V_{c}=\sum_{i\neq j}e^{2}/(2|{\bf r}_{i}-{\bf r}_{j}|)-\sum_{i}U(r_{i}). For brevity, we suppressed the positive-energy projection operators for the two-electron interactions (no-pair approximation).

As usual, we assume that the energies εi\varepsilon^{\prime}_{i} and orbitals ψi\psi^{\prime}_{i} of the unperturbed single-electron Hamiltonian h0h^{\prime}_{0} are known. Note that since the weak interaction is a pseudo-scalar, the total angular momentum jij_{i} and its projection mim_{i} remain good quantum numbers, while the parity is no longer conserved. For example, p1/2p_{1/2} and s1/2s_{1/2} orbitals or d5/2d_{5/2} and f5/2f_{5/2} orbitals of h0h_{0} are mixed to form eigen-states of h0h^{\prime}_{0}. The many-body eigen-states Ψ\Psi^{\prime} of HH^{\prime} are then expanded over antisymmetrized products of the PM one-particle orbitals ψi\psi^{\prime}_{i}. In MBPT, one obtains these eigen-states by treating the residual eee-e interaction VcV^{\prime}_{c} as a perturbation.

As the next step, we express the terms in Eq. (4) in second quantization. Let us denote by aia^{\prime\dagger}_{i} and aia^{\prime}_{i} the creation and annihilation operators associated with the one-particle eigen-state ψi\psi^{\prime}_{i} of h0h^{\prime}_{0}. We will follow the indexing convention that core electron orbitals are denoted by the letters at the beginning of alphabet a,b,c,a,b,c,\dots, while valence electron orbitals are denoted by v,w,v,w,\dots, and the indices i,j,k,i,j,k,\dots refer to an arbitrary orbital, core or excited (including valence states). The letters m,n,p,m,n,p,\dots are reserved for those orbitals unoccupied in the core (these could be valence orbitals).

The operators H0ih0(i)H^{\prime}_{0}\equiv\sum_{i}h^{\prime}_{0}(i) and VcV^{\prime}_{c} may then be written as

H0\displaystyle H^{\prime}_{0} =iεiN[aiai]\displaystyle=\sum_{i}\varepsilon^{\prime}_{i}N[a^{\prime\dagger}_{i}a^{\prime}_{i}] (5)
Vc\displaystyle V^{\prime}_{c} =ij(VHFU)ijN[aiaj]\displaystyle=\sum_{ij}\left(V^{\prime}_{\rm HF}-U^{\prime}\right)_{ij}N[a^{\prime\dagger}_{i}a^{\prime}_{j}]
+12ijklgijklN[aiajakal].\displaystyle+\frac{1}{2}\sum_{ijkl}g^{\prime}_{ijkl}N[a^{\prime\dagger}_{i}a^{\prime\dagger}_{j}a^{\prime}_{k}a^{\prime}_{l}]\,.

where NN denotes normal ordering and VHFV^{\prime}_{\rm HF} is the PM-DHF potential, whose matrix elements are defined by

(VHF)ijag~iaja,\left(V^{\prime}_{\rm HF}\right)_{ij}\equiv\sum_{a}\tilde{g}^{\prime}_{iaja}\,, (6)

with g~ijklgijklgijlk\tilde{g}^{\prime}_{ijkl}\equiv g^{\prime}_{ijkl}-g^{\prime}_{ijlk} being the anti-symmetrized combination of the Coulomb matrix elements,

gijkld3r1d3r2|𝐫1𝐫2|ψi(𝐫1)ψj(𝐫2)ψk(𝐫1)ψl(𝐫2).g^{\prime}_{ijkl}\equiv\int\frac{d^{3}{r}_{1}d^{3}{r}_{2}}{\left|{\bf r}_{1}-{\bf r}_{2}\right|}\psi^{\prime\dagger}_{i}({\bf r}_{1})\psi^{\prime\dagger}_{j}({\bf r}_{2})\psi^{\prime}_{k}({\bf r}_{1})\psi^{\prime}_{l}({\bf r}_{2})\,. (7)

The irrelevant constant offset energy term a(VHF/2U)aa\sum_{a}\left(V^{\prime}_{\rm HF}/2-U^{\prime}\right)_{aa} has been omitted in Eq. (5).

Notice that the choice U=VHFU^{\prime}=V^{\prime}_{\rm HF} causes the first term in VcV^{\prime}_{c} in Eq. (5) to vanish, significantly reducing the number of MBPT contributions. In addition, since our final goal is to implement the CCSDvT scheme that has been originally built on DHF potential, we fix U=VHFU^{\prime}=V^{\prime}_{\rm HF}.

With U=VHFU^{\prime}=V^{\prime}_{\rm HF} fixed, we now consider the correlation corrections to the independent-particle wave functions. Consider a univalent atom, e.g, Cs133{}^{133}\mathrm{Cs}, with a single valence electron above the closed-shell core. The zeroth-order wave function may be expressed as |Ψv(0)=av|0c\ket{\Psi^{\prime(0)}_{v}}=a^{\prime\dagger}_{v}\ket{0^{\prime}_{c}}, where |0c\ket{0^{\prime}_{c}} represents the filled Fermi sea of the atomic core (again, the prime indicates that the single-particle orbitals are of PM character).

To the first order in the residual interaction VcV^{\prime}_{c}, the many-body correction δΨv\delta\Psi^{\prime}_{v} to Ψv(0)\Psi^{\prime(0)}_{v} has the form

|δΨv\displaystyle\ket{\delta\Psi^{\prime}_{v}} =amngnmvaεavεnmanamaa|0c\displaystyle=\sum_{amn}\frac{g^{\prime}_{nmva}}{\varepsilon^{\prime}_{av}-\varepsilon^{\prime}_{nm}}a^{\prime\dagger}_{n}a^{\prime\dagger}_{m}a^{\prime}_{a}\ket{0^{\prime}_{c}}
+12abmngnmabεabεnmabaaanamav|0c,\displaystyle+\frac{1}{2}\sum_{abmn}\frac{g^{\prime}_{nmab}}{\varepsilon^{\prime}_{ab}-\varepsilon^{\prime}_{nm}}a^{\prime}_{b}a^{\prime}_{a}a^{\prime\dagger}_{n}a^{\prime\dagger}_{m}a^{\prime\dagger}_{v}\ket{0^{\prime}_{c}}\,, (8)

where we used the notation εijεi+εj\varepsilon^{\prime}_{ij}\equiv\varepsilon^{\prime}_{i}+\varepsilon^{\prime}_{j}. Here, the first term describes a valence electron being promoted to an excited state orbital mm with a simultaneous particle-hole excitation of the core (this is so-called valence double excitation, DvD_{v}, in the language of the CC method). The second contribution is a double particle-hole excitation from the core with valence electron being a spectator (core double excitation, DcD_{c}). The anti-commutation relations for creation and annihilation operators assure that the electrons in the second term do not get excited into the valence orbital, i.e. the Pauli exclusion principle is built into the formalism automatically.

With |δΨv\ket{\delta\Psi^{\prime}_{v}}, one can compute the second-order correction to the matrix element of a one-electron operator T=ijtijaiajT=\sum_{ij}t^{\prime}_{ij}a^{\prime\dagger}_{i}a^{\prime}_{j}. Once again, the primed quantities refer to the PM orbitals used in computing the matrix elements tiji|t|jt^{\prime}_{ij}\equiv\langle i^{\prime}|t|j^{\prime}\rangle. The operator TT can be, for example, the electric dipole operator 𝐃\bf D. The correction to the matrix element between two valence many-body states |Ψw\ket{\Psi^{\prime}_{w}} and |Ψv\ket{\Psi^{\prime}_{v}}, wvw\neq v, has the form

δTwv\displaystyle\delta T_{wv} =δΨw|T|Ψv+Ψw|T|δΨv\displaystyle=\bra{\delta\Psi^{\prime}_{w}}T\ket{\Psi^{\prime}_{v}}+\bra{\Psi^{\prime}_{w}}T\ket{\delta\Psi^{\prime}_{v}}
=antang~wnvaεaεnω+ang~wavntnaεaεn+ω,\displaystyle=\sum_{an}\frac{t^{\prime}_{an}\tilde{g}^{\prime}_{wnva}}{\varepsilon^{\prime}_{a}-\varepsilon^{\prime}_{n}-\omega}+\sum_{an}\frac{\tilde{g}^{\prime}_{wavn}t^{\prime}_{na}}{\varepsilon^{\prime}_{a}-\varepsilon^{\prime}_{n}+\omega}\,, (9)

where ωεwεv\omega\equiv\varepsilon^{\prime}_{w}-\varepsilon^{\prime}_{v} and g~ijklgijklgijlk\tilde{g}^{\prime}_{ijkl}\equiv g^{\prime}_{ijkl}-g^{\prime}_{ijlk}. Equations (II) and (II) are, of course, well-known results [57] with the only difference of using the mixed-parity basis instead of the conventional PP basis.

In Eq. (II), we sum over the core orbitals aa and the excited orbitals nn. Each orbital ψi\psi^{\prime}_{i} is characterized by a principal quantum number nin_{i}, a total angular momentum jij_{i}, and its projection mim_{i}. The sums over the magnetic quantum numbers mim_{i} can be carried out analytically using the rules of Racah algebra. Although the sums over jij_{i} are infinite, they are restricted by angular momentum selection rules which radically reduce the number of surviving terms. Moreover, the sums over total angular momenta converge well and in practice, it suffices to sum over a few lowest values of jij_{i}. The sums over the principal quantum numbers nin_{i} involve, on the other hand, summing over the infinite discrete spectrum and integrating over the continuum. In the basis set method, these infinite summations are replaced by summations over a finite-size pseudo-spectrum [58, 59, 54, 55].

The basis orbitals in the pseudo-spectrum are obtained by placing the atom in a sufficiently large cavity and imposing boundary conditions at the cavity wall and at the origin (see Ref. [55] for further details on a dual-kinetic-basis B-spline sets used in our paper). For each value of jij_{i}, one then finds a discrete set of 2N2N orbitals, NN from the Dirac sea and the remaining NN with energies above the Dirac sea threshold (conventionally referred to as “negative” and “positive” energy parts of the spectrum in analogy with free-fermion solutions).

If the size of the cavity is large enough, typically about 40a0/Z40a_{0}/Z where a0a_{0} is the Bohr radius, the low-lying orbitals with positive energies map with a good accuracy to the discrete orbitals of the exact DHF spectrum. Higher-energy orbitals do not closely match their physical counterparts. Nevertheless, since the pseudo-spectrum is complete, it forms a basis set for the function space spanning the cavity and thus can be used instead of the real spectrum to evaluate correlation corrections to states confined to the cavity. From now on, all single-particle orbitals ψi\psi^{\prime}_{i} are understood to be members of the B-spline basis set.

To reiterate, the parity-mixed (PM) formalism presented so far is essentially the same as in the conventional parity-proper (PP) MBPT. The only difference is that all the quantities are defined with respect to the PM orbitals ψi\psi^{\prime}_{i} instead of the PP ones. Since PM orbitals are eigen-states of total angular momentum, one can directly use the results of angular reduction for various MBPT expressions, and the existing MBPT codes require minor changes, mostly related to modifying parity selection rules and the use of Coulomb integrals in the PM basis. In Sec. III, we derive the second-quantized form of the parity operator which will be useful for deriving the PM selection rules and in Sec. IV, we present several methods through which the PM orbitals may be generated in practice.

III Parity operator in second quantization

Since the MBPT derivations are built on the second-quantization formalism, in this section we derive the second-quantized form of the parity operator Π\Pi to be used in deriving parity selection rules (see Appendix B). Parity transformation is defined by 𝐫i𝐫i\mathbf{r}_{i}\rightarrow-\mathbf{r}_{i} for all the NeN_{e} electrons in the system. Consider a PP state (Slater determinant) |Ψa1aμm1mμ\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}} composed of orbitals of definite parity. This many-body state is obtained by removing μ=0,,Ne\mu=0,\ldots,N_{e} electrons a1,,aμa_{1},\ldots,a_{\mu} from the reference state |Ψv(0)\ket{\Psi^{(0)}_{v}} while adding the same number of excited electrons m1,,mμm_{1},\ldots,m_{\mu}. Notice that in this notation the valence orbital is treated as initially occupied and, thereby, vv can be one of the labels a1,,aμa_{1},\ldots,a_{\mu}. In the second quantization,

|Ψa1aμm1mμam1amμaa1aaμ|Ψv(0).\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\equiv a^{\dagger}_{m_{1}}\ldots a^{\dagger}_{m_{\mu}}a_{a_{1}}\ldots a_{a_{\mu}}\ket{\Psi^{(0)}_{v}}\,. (10)

We emphasize that the creation and annihilation operators in Eq. (10) are the PP ones.

Since the PP Hamiltonian H0ih0(i)H_{0}\equiv\sum_{i}h_{0}(i) is invariant under spatial reflection, it commutes with the parity operator,[H0,Π]=0[H_{0},\Pi]=0. As a result, the states |Ψv(0)\ket{\Psi_{v}^{(0)}} and |Ψa1aμm1mμ\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}, being eigen-states of H0H_{0}, are also an eigen-states of the parity operator Π\Pi. Furthermore, since |Ψv(0)\ket{\Psi_{v}^{(0)}} and |Ψa1aμm1mμ\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}} are antisymmetrized products of single-electron orbitals, their eigen-values with respect to Π\Pi equal the products of the parities of their constituents

Π|Ψv(0)\displaystyle\Pi\ket{\Psi_{v}^{(0)}} =(1)v|Ψv(0),\displaystyle=(-1)^{\ell_{v}}\ket{\Psi_{v}^{(0)}}\,, (11a)
Π|Ψa1aμm1mμ\displaystyle\Pi\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}} =(1)v+i=1μai+mi|Ψa1aμm1mμ,\displaystyle=(-1)^{\ell_{v}+\sum_{i=1}^{\mu}\ell_{a_{i}}+\ell_{m_{i}}}\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\,, (11b)

where we have used the fact that the closed-shell core has even parity.

To transform an operator into the second-quantized form, we recall the conventional formula

Π=α,β|αα|Π|ββ|.\Pi=\sum_{\alpha,\beta}\ket{\alpha}\langle\alpha|\Pi|\beta\rangle\bra{\beta}\,. (12)

Its proof relies on the identity resolution (closure relation) for a complete orthonormal basis I=α|αα|=β|ββ|I=\sum_{\alpha}\ket{\alpha}\bra{\alpha}=\sum_{\beta}\ket{\beta}\bra{\beta}. For a system of identical particles, however, one needs to proceed with caution due to the possibility of permutations of orbitals in |Ψa1aμm1mμ\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}. Indeed, in the many-fermion case, the orthonormality condition reads

Ψa1aμm1mμ|Ψb1bνn1nν=δμνδa1aμb1bνδm1mμn1nν,\displaystyle\braket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}{\Psi^{n_{1}\ldots n_{\nu}}_{b_{1}\ldots b_{\nu}}}=\delta_{\mu\nu}\delta^{b_{1}\ldots b_{\nu}}_{a_{1}\ldots a_{\mu}}\delta^{n_{1}\ldots n_{\nu}}_{m_{1}\ldots m_{\mu}}\,, (13)

where the generalized Kronecker delta is defined as

δl1lμk1kμ={+1k1,,kμareanevenpermutationofl1,,lμ1k1,,kμareanoddpermutationofl1,,lμ0otherwise.\delta^{k_{1}\ldots k_{\mu}}_{l_{1}\ldots l_{\mu}}=\left\{\begin{matrix}+1&\begin{matrix}k_{1},\ldots,k_{\mu}{\rm\,\,are\,\,an\,\,even}\\ {\rm\quad permutation\,\,of\,\,}l_{1},\ldots,l_{\mu}\end{matrix}\\ -1&\begin{matrix}k_{1},\ldots,k_{\mu}{\rm\,\,are\,\,an\,\,odd}\\ {\rm\quad permutation\,\,of\,\,}l_{1},\ldots,l_{\mu}\end{matrix}\\ 0&{\rm otherwise}\end{matrix}\right.\,. (14)

For many-fermion systems, the general closure relation is given in Ref. [60]. Since Π\Pi is a diagonal operator, the general identity resolution [60] simplifies to

I=μ=0Ne1(μ!)2{a}{m}|Ψa1aμm1mμΨa1aμm1mμ|,\displaystyle I=\sum_{\mu=0}^{N_{e}}\frac{1}{(\mu!)^{2}}\sum_{\{a\}\{m\}}\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\bra{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\,, (15)

where {a}\{a\} and {m}\{m\} denote strings of orbital labels in |Ψa1aμm1mμ\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}.

Sandwiching Π\Pi in between two identity operators and using the closure relation (15), we find

Π\displaystyle\Pi =μ,ν=0Ne1(μ!ν!)2{a}{m}{b}{n}\displaystyle=\sum_{\mu,\nu=0}^{N_{e}}\frac{1}{(\mu!\nu!)^{2}}\sum_{\{a\}\{m\}}\sum_{\{b\}\{n\}}
×Ψa1aμm1mμ|Π|Ψb1bνn1nν|Ψa1aμm1mμΨb1bνn1nν|.\displaystyle\times\bra{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\Pi\ket{\Psi^{n_{1}\ldots n_{\nu}}_{b_{1}\ldots b_{\nu}}}\ket{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\bra{\Psi^{n_{1}\ldots n_{\nu}}_{b_{1}\ldots b_{\nu}}}\,. (16)

Using the eigenvalue equation (11b), we obtain the following result for the matrix element of Π\Pi

Ψa1aμm1mμ|Π|Ψb1bνn1nν\displaystyle\bra{\Psi^{m_{1}\ldots m_{\mu}}_{a_{1}\ldots a_{\mu}}}\Pi\ket{\Psi^{n_{1}\ldots n_{\nu}}_{b_{1}\ldots b_{\nu}}} =(1)v+i=1νbi+ni\displaystyle=(-1)^{\ell_{v}+\sum_{i=1}^{\nu}\ell_{b_{i}}+\ell_{n_{i}}}
×δμνδa1aμb1bνδm1mμn1nν,\displaystyle\times\delta_{\mu\nu}\delta^{b_{1}\ldots b_{\nu}}_{a_{1}\ldots a_{\mu}}\delta^{n_{1}\ldots n_{\nu}}_{m_{1}\ldots m_{\mu}}\,, (17)

where we used the orthonormality relation (13). Substituting Eq. (14) into Eq. (III), we finally obtain

Π\displaystyle\Pi =μ=0Ne1(μ!)2{a}{m}(1)v+i=1μai+mi\displaystyle=\sum_{\mu=0}^{N_{e}}\frac{1}{(\mu!)^{2}}\sum_{\{a\}\{m\}}(-1)^{\ell_{v}+\sum_{i=1}^{\mu}\ell_{a_{i}}+\ell_{m_{i}}}
×am1amμaa1aaμ|Ψv(0)Ψv(0)|\displaystyle\times a^{\dagger}_{m_{1}}\ldots a^{\dagger}_{m_{\mu}}a_{a_{1}}\ldots a_{a_{\mu}}\ket{\Psi^{(0)}_{v}}\bra{\Psi^{(0)}_{v}}
×aaμaa1amμam1.\displaystyle\times a^{\dagger}_{a_{\mu}}\ldots a^{\dagger}_{a_{1}}a_{m_{\mu}}\ldots a_{m_{1}}\,. (18)

IV Parity-mixed single-electron basis orbitals

In this section, we demonstrate how the PM single-particle wave functions ψiψnijimi\psi^{\prime}_{i}\equiv\psi^{\prime}_{n_{i}j_{i}m_{i}} may be obtained. They are the solutions to the PM-DHF equation

h0ψnijimi=εnijiψnijimi,\displaystyle h_{0}^{\prime}\psi^{\prime}_{n_{i}j_{i}m_{i}}=\varepsilon^{\prime}_{n_{i}j_{i}}\psi^{\prime}_{n_{i}j_{i}m_{i}}\,, (19)
h0=c𝜶𝐩+mec2β+Vnuc+hW+VHF.\displaystyle h_{0}^{\prime}=c\boldsymbol{\alpha}\cdot{\bf p}+m_{e}c^{2}\beta+V_{\rm nuc}+h_{W}+V^{\prime}_{\mathrm{HF}}\,.

Here, nin_{i} is the principal quantum number, jij_{i} is the total angular momentum and mim_{i} is the projection of jij_{i} on a quantization axis.

Our goal is to expand ψi\psi^{\prime}_{i} in terms of the PP orbitals ψiPψniijimiP\psi^{P}_{i}\equiv\psi^{P}_{n_{i}\ell_{i}j_{i}m_{i}}, which are solutions to the conventional DHF equation

h0ψniijimiP=εniijiψniijimiP,\displaystyle h_{0}\psi^{P}_{n_{i}\ell_{i}j_{i}m_{i}}=\varepsilon_{n_{i}\ell_{i}j_{i}}\psi^{P}_{n_{i}\ell_{i}j_{i}m_{i}}\,, (20)
h0=c𝜶𝐩+mec2β+Vnuc+VHF.\displaystyle h_{0}=c\boldsymbol{\alpha}\cdot{\bf p}+m_{e}c^{2}\beta+V_{\rm nuc}+V_{\mathrm{HF}}\,.

Note that besides the principal quantum number nin_{i}, the total angular momentum jij_{i}, and the magnetic quantum number mim_{i}, we have characterized the PP orbital ψiP\psi^{P}_{i} with an extra quantum number, the orbital angular momentum i\ell_{i}, which indicates that ψiP\psi^{P}_{i} has a definite parity (1)i(-1)^{\ell_{i}}.

The two DHF potentials VHFV^{\prime}_{\mathrm{HF}} and VHFV_{\mathrm{HF}} depend on the core orbitals. Since core orbitals are self-consistent solutions of these DHF equations, the PM and PP core orbitals differ. Therefore, the effects of the weak interaction on the single-electron orbitals are contained in the difference of the two Hamiltonians,

Δh=h0h0=hW+VHFVHF,\Delta h=h_{0}^{\prime}-h_{0}=h_{W}+V^{\prime}_{\mathrm{HF}}-V_{\mathrm{HF}}\,, (21)

which is a pseudo-scalar interaction, preserving rotational symmetry but spoiling mirror symmetry (this is why we have used the quantum numbers nin_{i}, jij_{i} and mim_{i} but not i\ell_{i} to index the PM orbital ψi\psi^{\prime}_{i}). This suggests the following parameterization [61] for the solutions to the PM-DHF equations (19),

ψi(𝐫)\displaystyle\psi^{\prime}_{i}(\mathbf{r}) =ψi(𝐫)+iηψ¯i(𝐫),\displaystyle=\psi_{i}(\mathbf{r})+{\rm i}\eta\bar{\psi}_{i}(\mathbf{r})\,, (22a)
ψi(𝐫)\displaystyle\psi_{i}(\mathbf{r}) =1r(iPniκi(r)Ωκimi(𝐫^)Qniκi(r)Ωκimi(𝐫^)),\displaystyle=\frac{1}{r}\left(\begin{array}[]{c}{\rm i}P_{n_{i}\kappa_{i}}(r)\Omega_{\kappa_{i}m_{i}}(\hat{\mathbf{r}})\\ Q_{n_{i}\kappa_{i}}(r)\Omega_{-\kappa_{i}m_{i}}(\hat{\mathbf{r}})\end{array}\right)\,, (22d)
ψ¯i(𝐫)\displaystyle\bar{\psi}_{i}(\mathbf{r}) =1r(iP¯niκi(r)Ωκimi(𝐫^)Q¯niκi(r)Ωκimi(𝐫^)),\displaystyle=\frac{1}{r}\left(\begin{array}[]{c}{\rm i}\bar{P}_{n_{i}\kappa_{i}}(r)\Omega_{-\kappa_{i}m_{i}}(\hat{\mathbf{r}})\\ \bar{Q}_{n_{i}\kappa_{i}}(r)\Omega_{\kappa_{i}m_{i}}(\hat{\mathbf{r}})\end{array}\right)\,, (22g)

where

ηGFQW22a021015\eta\equiv\frac{G_{F}Q_{W}}{2\sqrt{2}a_{0}^{2}}\sim 10^{-15} (23)

is a dimensionless factor characteristic of the strength of the weak interaction and its numerical value is given for Cs133{}^{133}\mathrm{Cs}. In Eqs. (22), κi=(ji+12)(1)ji+i+1/2\kappa_{i}=\left(j_{i}+\frac{1}{2}\right)(-1)^{j_{i}+\ell_{i}+1/2} is a relativistic angular quantum number that encodes the values of both the total and orbital angular momenta jij_{i} and i\ell_{i}. From the definition of the relativistic angular quantum number, flipping the parity (1)i(-1)^{\ell_{i}} of the orbital while preserving the total angular momentum jij_{i} is equivalent to changing κiκi\kappa_{i}\rightarrow-\kappa_{i}, as presented in the parameterization of ψ¯i\bar{\psi}_{i}, Eq. (22g).

It is appropriate to pause here and introduce a point of semantics. Although the orbital ψi\psi^{\prime}_{i} does not have a definite parity, one can nevertheless speak of its “nominal parity”, defined as that of the component ψi\psi_{i}, which is not suppressed by the factor η\eta. In the light of Eq. (22a), we shall refer to ψi\psi_{i} as the “real” component and ψ¯i\bar{\psi}_{i} as the “imaginary” component of ψi\psi^{\prime}_{i}. In what follows, in particular when discussing MBPT and the CC formalism, we shall refer to the nominal parity of a PM single-electron orbital, meaning that of its real component. The nominal parity of ψi\psi^{\prime}_{i} is thus (1)i(-1)^{\ell_{i}}.

The combination (22a) clearly demonstrates the admixing of the opposite-parity orbital ψ¯i\bar{\psi}_{i} with κiκi\kappa_{i}\rightarrow-\kappa_{i} to the reference orbital ψi\psi_{i}. With the imaginary unity factored out in the admixture component ψ¯i\bar{\psi}_{i} and with the conventional definition of the spherical spinors Ωκm\Omega_{\kappa m} [62], all the radial wave functions PniκiP_{n_{i}\kappa_{i}}, QniκiQ_{n_{i}\kappa_{i}}, P¯niκi\bar{P}_{n_{i}\kappa_{i}}, and Q¯niκi\bar{Q}_{n_{i}\kappa_{i}} can be chosen to be real-valued. The rest of this section will be devoted to finding these radial components. In what follows, we will assume a parameteriziton for the PP solutions to Eq. (LABEL:Normal_DHF) similar to that presented in Eqs. (22d), namely

ψiP(𝐫)=1r(iPniκiP(r)Ωκimi(𝐫^)QniκiP(r)Ωκimi(𝐫^)).\psi^{P}_{i}(\mathbf{r})=\frac{1}{r}\left(\begin{array}[]{c}{\rm i}P^{P}_{n_{i}\kappa_{i}}(r)\Omega_{\kappa_{i}m_{i}}(\hat{\mathbf{r}})\\ Q^{P}_{n_{i}\kappa_{i}}(r)\Omega_{-\kappa_{i}m_{i}}(\hat{\mathbf{r}})\end{array}\right)\,. (24)

Where there is no risk of confusion, we will abbreviate PniκiP_{n_{i}\kappa_{i}} (QniκiQ_{n_{i}\kappa_{i}}) to PiP_{i} (QiQ_{i}) and PniκiPP^{P}_{n_{i}\kappa_{i}} (QniκiPQ^{P}_{n_{i}\kappa_{i}}) to PiPP^{P}_{i} (QiPQ^{P}_{i}). The energy eigenvalues ϵniji\epsilon^{\prime}_{n_{i}j_{i}} and ϵniliji\epsilon_{n_{i}l_{i}j_{i}} in Eqs. (19) and (LABEL:Normal_DHF) will also be abbreviated to ϵi\epsilon^{\prime}_{i} and ϵi\epsilon_{i}, respectively.

IV.1 Parity-mixed basis set construction: finite-difference method

We start our computation of the radial functions PiP_{i}, QiQ_{i}, P¯i\bar{P}_{i} and Q¯i\bar{Q}_{i} with a discussion of the finite-difference method, where we integrate the PM-DHF equations directly, without using the basis set technique. While the finite-difference method does not produce the finite basis set for MBPT-type calculations, it generates the PM core orbitals entering the PM-DHF potential that can be used in constructing the basis set. In addition, the finite-difference method provides reference results that are used to gauge the fidelity of the basis set representation of core and low-energy orbitals.

Due to the smallness of the dimensionless coupling constant η\eta, we may set ψi=ψiP\psi_{i}=\psi^{P}_{i} which is accurate up to O(η2)O(\eta^{2}). As a result, to the first order in η\eta, Eq. (19) yields a pair of integro-differential equations for the radial functions P¯i\bar{P}_{i} and Q¯i\bar{Q}_{i}. For a core orbital, i=ai=a, these equations read

c(ddrκar)P¯a(Veffεac2)Q¯a\displaystyle c\left(\frac{d}{dr}-\frac{\kappa_{a}}{r}\right)\bar{P}_{a}-\left(V_{\rm eff}-\varepsilon_{a}-c^{2}\right)\bar{Q}_{a}
=ρnucPaPbV¯baQbPbVbaQ¯b,\displaystyle=-\rho_{\rm nuc}P^{P}_{a}-\sum_{b}\bar{V}_{ba}Q^{P}_{b}-\sum_{b}V_{ba}\bar{Q}_{b}\,, (25a)
c(ddr+κar)Q¯a+(Veffεa+c2)P¯a\displaystyle c\left(\frac{d}{dr}+\frac{\kappa_{a}}{r}\right)\bar{Q}_{a}+\left(V_{\rm eff}-\varepsilon_{a}+c^{2}\right)\bar{P}_{a}
=ρnucQaP+bV¯baPbP+bVbaP¯b,\displaystyle=-\rho_{\rm nuc}Q^{P}_{a}+\sum_{b}\bar{V}_{ba}P^{P}_{b}+\sum_{b}V_{ba}\bar{P}_{b}\,, (25b)

where VeffV_{\rm eff} is an effective potential comprising of the electron-nucleus Coulomb potential and the direct part of the conventionally-defined DHF potential ([j]2j+1[j]\equiv 2j+1)

Veff(r)Vnuc(r)+b[jb]v0(b,b,r),V_{\rm eff}(r)\equiv V_{\rm nuc}(r)+\sum_{b}[j_{b}]v_{0}(b,b,r)\,, (26)

while VbaV_{ba} is the DHF exchange potential

Vba(r)1[ja]kκaCkκb2vk(b,a,r),V_{ba}(r)\equiv\frac{1}{[j_{a}]}\sum_{k}\langle\kappa_{a}||C_{k}||\kappa_{b}\rangle^{2}v_{k}(b,a,r)\,, (27)

and V¯ba\bar{V}_{ba} is the PNC-DHF exchange potential

V¯ba(r)1[ja]kκaCkκb2(vk(b,a¯,r)vk(b¯,a,r)).\bar{V}_{ba}(r)\equiv\frac{1}{[j_{a}]}\sum_{k}\langle-\kappa_{a}||C_{k}||\kappa_{b}\rangle^{2}\left(v_{k}(b,\bar{a},r)-v_{k}(\bar{b},a,r)\right)\,. (28)

In Eqs. (26) and (27), the multipolar potential vk(b,a,r)v_{k}(b,a,r) is defined as

vk(b,a,r)\displaystyle v_{k}(b,a,r) r<kr>k1𝑑r\displaystyle\equiv\int r_{<}^{k}r_{>}^{-k-1}dr^{\prime} (29)
×(PbP(r)PaP(r)+QbP(r)QaP(r)),\displaystyle\times\left(P^{P}_{b}(r^{\prime})P^{P}_{a}(r^{\prime})+Q^{P}_{b}(r^{\prime})Q^{P}_{a}(r^{\prime})\right)\,,

whereas the quantity vk(b,a¯,r)v_{k}(b,\bar{a},r) in Eq. (28) is defined as

vk(b,a¯,r)\displaystyle v_{k}(b,\bar{a},r) r<kr>k1𝑑r\displaystyle\equiv\int r_{<}^{k}r_{>}^{-k-1}dr^{\prime} (30)
×(PbP(r)P¯a(r)+QbP(r)Q¯a(r)),\displaystyle\times\left(P^{P}_{b}(r^{\prime})\bar{P}_{a}(r^{\prime})+Q^{P}_{b}(r^{\prime})\bar{Q}_{a}(r^{\prime})\right)\,,

and similarly for vk(b¯,a,r)v_{k}(\bar{b},a,r). In these equations, r<=min(r,r)r_{<}=\min(r,r^{\prime}) and r>=max(r,r)r_{>}=\max(r,r^{\prime}).

Equations (25) may be solved using an iterative scheme (nn is the iteration number)

c(ddrκar)P¯a(n+1)\displaystyle c\left(\frac{d}{dr}-\frac{\kappa_{a}}{r}\right)\bar{P}^{(n+1)}_{a} (Veffεac2)Q¯a(n+1)\displaystyle-\left(V_{\rm eff}-\varepsilon_{a}-c^{2}\right)\bar{Q}^{(n+1)}_{a}
=ρnucPaPXa(n+1),\displaystyle=-\rho_{\rm nuc}P^{P}_{a}-X_{a}^{(n+1)}\,, (31a)
c(ddr+κar)Q¯a(n+1)\displaystyle c\left(\frac{d}{dr}+\frac{\kappa_{a}}{r}\right)\bar{Q}^{(n+1)}_{a} +(Veffεa+c2)P¯a(n+1)\displaystyle+\left(V_{\rm eff}-\varepsilon_{a}+c^{2}\right)\bar{P}^{(n+1)}_{a}
=ρnucQaP+Ya(n+1),\displaystyle=-\rho_{\rm nuc}Q^{P}_{a}+Y_{a}^{(n+1)}\,, (31b)

where we have defined

Xa(n+1)bV¯ba(n+1)QbPbVbaQ¯b(n+1),\displaystyle X_{a}^{(n+1)}\equiv\sum_{b}\bar{V}^{(n+1)}_{ba}Q^{P}_{b}-\sum_{b}V_{ba}\bar{Q}^{(n+1)}_{b}\,, (32a)
Ya(n+1)bV¯ba(n+1)PbP+bVbaP¯b(n+1).\displaystyle Y_{a}^{(n+1)}\equiv\sum_{b}\bar{V}^{(n+1)}_{ba}P^{P}_{b}+\sum_{b}V_{ba}\bar{P}^{(n+1)}_{b}\,. (32b)

Note that XaX_{a} and YaY_{a} are themselves functions of P¯a\bar{P}_{a} and Q¯a\bar{Q}_{a}, which appear explicitly in the second terms of Eqs. (32) and implicitly via the PNC-DHF exchange potential V¯ba\bar{V}_{ba} in the first terms of Eqs. (32).

Equations (31) are inhomogeneous second-order differential equations which may be solved using the conventional technique of variation of parameters. In this method, one first finds the solution to the homogeneous version of Eqs. (31). Since the operators acting on P¯a\bar{P}_{a} and Q¯a\bar{Q}_{a} on the left hand side of Eqs. (31) do not change from iteration to iteration, neither will the homogeneous solutions. As a result, they only need to be computed once. The inhomogeneous solutions P¯a(n+1)\bar{P}^{(n+1)}_{a} and Q¯a(n+1)\bar{Q}^{(n+1)}_{a} are then obtained by convoluting the corresponding homogeneous solutions with the right hand sides of Eqs. (31) (see, e.g., Ref. [62] for further details on the technique of variation of parameters for DHF equation).

Once the radial functions P¯a\bar{P}_{a} and Q¯a\bar{Q}_{a} are obtained, we may proceed to solving for the radial functions P¯m\bar{P}_{m} and Q¯m\bar{Q}_{m} of the unoccupied orbitals. The equations for P¯m\bar{P}_{m} and Q¯m\bar{Q}_{m} are obtained by replacing ama\rightarrow m in Eqs. (25) and we may set up a similar iteration scheme for valence orbitals as in Eqs. (31). Note that in this case, the driving terms XmX_{m} and YmY_{m} depend on P¯m\bar{P}_{m} and Q¯m\bar{Q}_{m} via the PNC-DHF potential V¯bm\bar{V}_{bm} only. Other than this, the procedure for solving the PNC-DHF for unoccupied orbitals is the same as for core orbitals.

We note, however, that in general, the iteration scheme (31) and also its counterpart for unoccupied orbitals do not converge but oscillate. Such behavior can be removed if the driving terms XX and YY are changed slowly between iterations. This is accomplished by setting

Xi(n+1)\displaystyle X_{i}^{(n+1)} =λXi(n+1)+(1λ)Xi(n),\displaystyle=\lambda X_{i}^{(n+1)}+(1-\lambda)X_{i}^{(n)}\,, (33)
Yi(n+1)\displaystyle Y_{i}^{(n+1)} =λYi(n+1)+(1λ)Yi(n).\displaystyle=\lambda Y_{i}^{(n+1)}+(1-\lambda)Y_{i}^{(n)}\,.

We find that choosing λ=0.010.1\lambda=0.01\sim 0.1 generally assures iteration convergence for all the orbitals, core and unoccupied. In the rest of this section, we shall discuss two matrix methods which allow us to avoid altogether this issue of convergence.

As a check for our numerical procedure for the finite-difference method, we recovered the previous literature results [33, 42] for the lowest order 6S1/27S1/26S_{1/2}-7S_{1/2} PNC transition amplitude in Cs133{}^{133}\mathrm{Cs}. We calculated the amplitudes in both the frozen-core (fc) approximation, which involves neglecting the PNC effects on core orbitals, obtaining

EPVfc=0.73946×1011i|e|a0(QW/N),E^{\rm fc}_{\rm PV}=0.73946\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,, (34)

and the full core-perturbed (cp) case, where the PNC perturbation to core orbitals is fully taken into account, obtaining

EPVcp=0.92700×1011i|e|a0(QW/N).E^{\rm cp}_{\rm PV}=0.92700\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,. (35)

In all our numerical examples, the nuclear charge distribution is approximated by a Fermi distribution ρnuc(r)=ρ0/(1+exp[(rc)/a])\rho_{\mathrm{nuc}}(r)=\rho_{0}/(1+\exp[(r-c)/a]), where ρ0\rho_{0} is a normalization constant. For Cs133{}^{133}\mathrm{Cs}, we use c=5.6748fmc=5.6748\,\mathrm{fm} and a=0.52338fma=0.52338\,\mathrm{fm}. We also use the same nuclear distribution in computations of weak interaction (2), ρ(r)ρnuc(r)\rho(r)\equiv\rho_{\mathrm{nuc}}(r).

IV.2 Parity-mixed basis set construction: exact matrix diagonalization methods

The goal of this section is to construct a PM-DHF basis set {ψi}\{\psi^{\prime}_{i}\} by transforming a numerically complete PP-DHF basis set {ψiP}\{\psi_{i}^{P}\}: {ψiP}{ψi}\{\psi_{i}^{P}\}\rightarrow\{\psi^{\prime}_{i}\} (basis rotation). The PP-DHF basis sets based on the solution of the conventional PP-DHF equations are widely used both in atomic structure and quantum chemistry calculations and we assume that the set {ψiP}\{\psi_{i}^{P}\} was pre-computed.

The two DHF equations, PM- and PP-DHF, differ by Δh\Delta h, Eq. (21), which includes the weak interaction and the difference between the two DHF potentials. While the weak interaction is a small perturbation, Δhη1015\Delta h\sim\eta\sim 10^{-15} for Cs133{}^{133}\mathrm{Cs}, one may encounter accidental degeneracies between basis orbitals of opposite parities (especially in the high-energy part of the pseudo-spectra), making application of perturbative approaches error-prone. In this subsection, we discuss two exact methods based on the diagonalization of the PM-DHF Hamiltonian, and in the next subsection, we explore the perturbative approach.

The two approaches considered in this subsection involve transforming the PP-DHF basis {ψiP}\{\psi_{i}^{P}\} into the desired PM-DHF basis {ψi}\{\psi^{\prime}_{i}\}: (i) without requiring the prior computation of the PM-DHF core orbitals and (ii) with the PM-DHF potential pre-computed using, say, the finite-difference method of the previous section.

Let us consider the first method. Suppose we do not know the PM-DHF core orbitals and thus can not immediately construct the PM-DHF potential beforehand. Recall that the PM-DHF orbitals are represented as ψi=ψi+iηψ¯i\psi^{\prime}_{i}=\psi_{i}+i\eta\bar{\psi}_{i}, Eq. (22), where ψi\psi_{i} is the nominal parity contribution and ψ¯i\bar{\psi}_{i} is the opposite-parity admixture. Since the PP-DHF set {ψiP}\{\psi_{i}^{P}\} forms a numerically-complete basis, the nominal parity contribution ψi\psi_{i} can be expanded in terms of the 2N2N orbitals ψjP\psi^{P}_{j} of the same total angular momentum and parity as ψi\psi_{i}, i.e. κj=κi\kappa_{j}=\kappa_{i} (recall that 2N2N is the number of basis functions for a given κ\kappa). Similarly, the opposite-parity admixtures ψ¯i\bar{\psi}_{i} may be expanded over the 2N2N PP-DHF orbitals ψj¯P\psi^{P}_{\bar{j}} which have the same total angular momentum but opposite parity to ψi\psi_{i}, i.e. κj¯=κi\kappa_{\bar{j}}=-\kappa_{i}.

As a result, the PM wave function ψi\psi^{\prime}_{i} may be written as

ψi=jχijψjP+ij¯χij¯ψj¯P,\psi^{\prime}_{i}=\sum_{j}\chi_{ij}\psi^{P}_{j}+{\rm i}\sum_{\bar{j}}\chi_{i\bar{j}}\psi^{P}_{\bar{j}}\,, (36)

where the factor η\eta has been absorbed into the opposite-parity admixture coefficients χij¯\chi_{i\bar{j}}, i.e. χij¯O(η)\chi_{i\bar{j}}\sim O(\eta). More explicitly, Eq. (36) reads

ψnijimi=njχijψnjijimiP+inj¯χij¯ψnj¯¯ijimiP,\psi^{\prime}_{n_{i}j_{i}m_{i}}=\sum_{n_{j}}\chi_{ij}\psi^{P}_{n_{j}\ell_{i}j_{i}m_{i}}+{\rm i}\sum_{n_{\bar{j}}}\chi_{i\bar{j}}\psi^{P}_{n_{\bar{j}}\bar{\ell}_{i}j_{i}m_{i}}\,, (37)

where the index ¯i=i±1\bar{\ell}_{i}=\ell_{i}\pm 1 indicates that terms in the second summation in Eq. (37) have opposite parities to those in the first summation.

In terms of the radial wave functions PiP_{i} (QiQ_{i}) and P¯i\bar{P}_{i} (Q¯i\bar{Q}_{i}), Eqs. (36) and (37) are equivalent to

Pi\displaystyle P_{i} =njχijPnjκiP,Gi=njχijQnjκiP,\displaystyle=\sum_{n_{j}}\chi_{ij}P^{P}_{n_{j}\kappa_{i}}\,,\quad G_{i}=\sum_{n_{j}}\chi_{ij}Q^{P}_{n_{j}\kappa_{i}}\,, (38)
P¯i\displaystyle\bar{P}_{i} =njχij¯PnjκiP,Q¯i=njχij¯QnjκiP,\displaystyle=\sum_{n_{j}}\chi_{i\bar{j}}P^{P}_{n_{j}-\kappa_{i}}\,,\quad\bar{Q}_{i}=\sum_{n_{j}}\chi_{i\bar{j}}Q^{P}_{n_{j}-\kappa_{i}}\,,

where we have fixed the relativistic angular numbers κj=±κi\kappa_{j}=\pm\kappa_{i} to reflect parities.

Substituting the expansion (36) into Eq. (19), multiplying with ψjP{\psi^{P}_{j}}^{\dagger} and ψj¯P{\psi^{P}_{\bar{j}}}^{\dagger}, and then integrating, we obtain

εkχij+ij¯j|Δh|j¯χij¯=εiχij,\displaystyle\varepsilon_{k}\chi_{ij}+{\rm i}\sum_{\bar{j}}\bra{j}\Delta h\ket{\bar{j}}\chi_{i\bar{j}}=\varepsilon^{\prime}_{i}\chi_{ij}\,, (39a)
jj¯|Δh|jχij+iεj¯χij¯=iεiχij¯,\displaystyle\sum_{j}\bra{\bar{j}}\Delta h\ket{j}\chi_{ij}+{\rm i}\varepsilon_{\bar{j}}\chi_{i{\bar{j}}}={\rm i}\varepsilon^{\prime}_{i}\chi_{i{\bar{j}}}\,, (39b)

where we have used the fact that Δh\Delta h, Eq. (21), can only connect orbitals of opposite parities.

Equations (39) may be put in the form of an eigenvalue matrix equation,

𝑴𝝌i=εi𝝌i,\boldsymbol{M}\boldsymbol{\chi}_{i}=\varepsilon^{\prime}_{i}\boldsymbol{\chi}_{i}\,, (40)

where 𝝌i(χij,χij¯)\boldsymbol{\chi}_{i}\equiv(\chi_{ij},\chi_{i{\bar{j}}}) and 𝑴\boldsymbol{M} is a 4N×4N4N\times 4N matrix defined by

Mjj\displaystyle M_{jj} =εj,Mjj¯=ij|Δh|j¯,\displaystyle=\varepsilon_{j}\,,\quad M_{j\bar{j}}={\rm i}\bra{j}\Delta h\ket{\bar{j}}\,, (41)
Mj¯j¯\displaystyle M_{{\bar{j}}{\bar{j}}} =εj¯,Mj¯j=ij¯|Δh|j.\displaystyle=\varepsilon_{\bar{j}}\,,\quad M_{\bar{j}j}=-{\rm i}\bra{\bar{j}}\Delta h\ket{j}\,.

The matrix 𝑴\boldsymbol{M}, Eq. (41), is a real symmetric matrix. As a result, its eigenvalues εi\varepsilon^{\prime}_{i} and eigenvectors 𝝌i\boldsymbol{\chi}_{i} are real. Further, we may express the off-diagonal elements of 𝑴\boldsymbol{M} in a more explicit form:

ij¯|Δh|j\displaystyle-{\rm i}\bra{\bar{j}}\Delta h\ket{j} =ij¯|hW+VHFVHF|j\displaystyle=-{\rm i}\bra{\bar{j}}h_{W}+V^{\prime}_{\rm HF}-V_{\rm HF}\ket{j}
=ηSj¯jakk¯χakχak¯(gj¯kk¯jgj¯k¯kj),\displaystyle=\eta S_{\bar{j}j}-\sum_{ak\bar{k}}\chi_{ak}\chi_{a\bar{k}}(g_{\bar{j}k\bar{k}j}-g_{\bar{j}\bar{k}kj})\,, (42)

where

Siji|iργ5|j=(PiPQjPPjPQiP)ρ(r)𝑑r.S_{ij}\equiv\bra{i}{\rm i}\rho\gamma_{5}\ket{j}=\int\left(P^{P}_{i}Q^{P}_{j}-P^{P}_{j}Q^{P}_{i}\right)\rho(r)dr\,. (43)

Note that here, the Coulomb matrix elements gj¯kk¯jg_{\bar{j}k\bar{k}j} and gj¯k¯kjg_{\bar{j}\bar{k}kj} are defined with respect to the PP basis orbitals ψjP\psi^{P}_{j}, ψkP\psi^{P}_{k}, ψk¯P\psi^{P}_{\bar{k}} and ψj¯P\psi^{P}_{\bar{j}}. The orbital ψkP\psi^{P}_{k} has the same total angular momentum and parity as the core orbital ψaP\psi^{P}_{a}, i.e. κk=κa\kappa_{k}=\kappa_{a}, whereas ψk¯P\psi^{P}_{\bar{k}} has the same total angular momentum but opposite parity to ψaP\psi^{P}_{a}, i.e. κk¯=κa\kappa_{\bar{k}}=-\kappa_{a}. Note that the orbitals ψkP\psi^{P}_{k} and ψk¯P\psi^{P}_{\bar{k}} are not limited to the core and do not necessarily have the same principal quantum numbers. The quantities SijS_{ij} defined in Eq. (43) are real and anti-symmetric.

Due to the second term in Eq. (IV.2), the matrix element of Δh\Delta h depends on the PM-DHF potential and thereby on the yet to be determined PM-DHF core orbitals. Therefore, Eq. (40) is nonlinear and needs to be iterated until convergence. The iteration of Eq. (40) generally does not suffer from the oscillating convergence behaviour as the finite-difference method. The change in the results from one iteration to another oscillates for the first few iterations but quickly decreases in a monotonous fashion. The price to be paid for this well-behaved convergence pattern is the need to pre-compute a large number of matrix elements of the form gj¯kk¯jgj¯k¯kjg_{\bar{j}k\bar{k}j}-g_{\bar{j}\bar{k}kj} required in forming the VHFV^{\prime}_{\rm HF} term in Eq. (40).

Note also that since the 𝑴\boldsymbol{M} matrices corresponding to ψi¯\psi^{\prime}_{\bar{i}} and ψi\psi^{\prime}_{i} are related by swapping jj¯j\leftrightarrow\bar{j} in Eq. (41), there is no need to diagonalize them separately. Instead, we form the 𝑴\boldsymbol{M} matrix only for negative values of κi=1,2,\kappa_{i}=-1,-2,\dots. Each such matrix then has 4N4N eigenvectors, 2N2N of which correspond to the negative and positive energy orbitals ψi\psi^{\prime}_{i} while the other 2N2N give the expansion for the orbitals ψi¯\psi^{\prime}_{\bar{i}}. We ensure the correct assignment of eigenvectors to orbitals by exploiting the fact that χiiO(1)\chi_{ii}\sim O(1), χijO(η2)\chi_{ij}\sim O(\eta^{2}) for jij\neq i, and χij¯O(η)\chi_{i\bar{j}}\sim O(\eta), in accordance with the results from perturbation theory.

We now discuss the second method where, to avoid iterations in determining the PM-DHF core orbitals, one can also pre-compute them using the finite-difference solution of PM-DHF equations, see Sec. IV.1. This is the strategy used earlier for basis set generation in the context of Breit interaction [48]. Then the required matrix elements of Δh\Delta h, Eq. (21), can be computed immediately and the diagonalization proceeds in a single step. Comparing the PM-DHF core and low-lying excited orbitals from the finite-difference and basis-set solutions provides a valuable test of the accuracy.

In both approaches, one has to be mindful of the smallness of the parameter η1015\eta\sim 10^{-15}, which is comparable to the accuracy of double precision operations. Care should be taken when diagonalizing the matrix 𝑴\boldsymbol{M} to avoid numerical truncation errors. This issue may be effectively dealt with by using a multiple-precision diagonalization algorithm. In our numerical computations, we modified the routines tred2 and tqli presented in Ref. [63] to perform quadruple (128 bits) precision diagonalization and used these upgraded routines to diagonalize the matrices 𝑴\boldsymbol{M}.

An alternative to matrix diagonalization is a perturbative approach that uses the smallness of parameter η\eta, see Sec. IV.3. However, the non-perturbative method described in this subsection is more general and is more accurate in the case of accidental degeneracies in the pseudo-spectra of h0h_{0} between orbitals with the same total angular momentum but of opposite parities (see Sec. IV.4 below for further discussions).

We used the matrix diagonalization method discussed in this subsection to generate for Cs133{}^{133}\mathrm{Cs} a PM basis of total angular momenta ranging from 1/21/2 to 13/213/2 (one set for each method). The PP set used to expand the PM orbitals are B-splines obtained using the dual-kinetic-balance method [55]. Each set of the PP partial waves with κi{±1,,±7}\kappa_{i}\in\{\pm 1,\dots,\pm 7\} contains N=40N=40 positive-energy orbitals. The cavity radius is chosen to be 50 a.u. and computations were performed on a nonuniform grid of 500 points with 40 points inside the nucleus.

The PM core orbitals are read in from the finite-difference calculation and the PNC-DHF potential VHFVHFV^{\prime}_{\rm HF}-V_{\rm HF} is computed with these core orbitals. The rest of the PM basis is obtained by diagonalizing the matrices MM corresponding to κi=1,2,,7\kappa_{i}=-1,-2,\dots,-7. The lowest order 6S1/27S1/26S_{1/2}-7S_{1/2} PNC frozen-core and core-perturbed amplitudes for Cs computed using the so-obtained PM-DHF valence orbitals ψ6s1/2\psi^{\prime}_{6s_{1/2}} and ψ6s1/2\psi^{\prime}_{6s_{1/2}} are, respectively

EPVfc\displaystyle E^{\rm fc}_{\rm PV} =0.73949×1011i|e|a0(QW/N),\displaystyle=0.73949\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,, (44)
EPVcp\displaystyle E^{\rm cp}_{\rm PV} =0.92701×1011i|e|a0(QW/N).\displaystyle=0.92701\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,.

The differences between these basis-set values and the finite-difference results (34) and (35) are at the level of 0.001%. This numerical error is adequate for our goals.

IV.3 Parity-mixed basis set construction: Perturbative matrix method

The need for an iterative scheme and the numerical difficulty associated with the smallness of the PNC matrix elements may be avoided entirely if we adopt ab initio a form of expansion for the PM orbitals ψi\psi^{\prime}_{i}, Eq. (22a), in accordance with perturbation theory. To the first order in η\eta, perturbation theory tells us that

ψi=ψiP+j¯j¯|Δh|iεiεj¯ψj¯P,\psi^{\prime}_{i}=\psi^{P}_{i}+\sum_{\bar{j}}\frac{\bra{\bar{j}}\Delta h\ket{i}}{\varepsilon_{i}-\varepsilon_{\bar{j}}}\psi^{P}_{\bar{j}}\,, (45)

where the sum runs over all PP orbitals ψj¯P\psi^{P}_{\bar{j}} with the same total angular momentum but opposite parity to ψiP\psi^{P}_{i}. More explicitly, Eq. (45) has the form

ψnijimi\displaystyle\psi^{\prime}_{n_{i}j_{i}m_{i}} =ψniijimiP\displaystyle=\psi^{P}_{n_{i}\ell_{i}j_{i}m_{i}} (46)
+nj¯nj¯¯ijimi|Δh|niijimiεniijiεnj¯¯ijiψnj¯¯ijimiP,\displaystyle+\sum_{n_{\bar{j}}}\frac{\bra{n_{\bar{j}}\bar{\ell}_{i}j_{i}m_{i}}\Delta h\ket{n_{i}\ell_{i}j_{i}m_{i}}}{\varepsilon_{n_{i}\ell_{i}j_{i}}-\varepsilon_{n_{\bar{j}}\bar{\ell}_{i}j_{i}}}\psi^{P}_{n_{\bar{j}}\bar{\ell}_{i}j_{i}m_{i}}\,,

where, again, the index ¯i=i±1\bar{\ell}_{i}=\ell_{i}\pm 1 indicates that terms in the sum over nj¯n_{\bar{j}} have opposite parities to ψniijimiP\psi^{P}_{n_{i}\ell_{i}j_{i}m_{i}}.

If the PM-DHF potential VHFV^{\prime}_{\rm HF} has been constructed beforehand, e.g., by solving the finite difference Eqs. (25) for the PM core orbitals then Eq. (45) may be used to directly compute the opposite-parity admixtures (the sum) for all PM excited orbitals. In contrast, if the PM core orbitals and the PM-DHF potential VHFV^{\prime}_{\rm HF} are not known beforehand, the matrix method developed in Sec. IV.2 may be used to solve for these orbitals as follows.

It is clear from Eq. (45) that in a perturbative approach, the expansion coefficients χij\chi_{ij} and χij¯\chi_{i\bar{j}} in Eq. (36) have the form

χij=δij,χij¯=ηγij¯,\begin{matrix}\chi_{ij}=\delta_{ij}\,,&\chi_{i\bar{j}}=\eta\gamma_{i\bar{j}}\,,\end{matrix} (47)

which makes it explicit that in the limit where η0\eta\rightarrow 0, ψiψiP\psi^{\prime}_{i}\rightarrow\psi^{P}_{i}. Setting χii=1\chi_{ii}=1 guarantees that ψi\psi^{\prime}_{i} is normalized up to O(η2)O(\eta^{2}). Factoring out the imaginary unit from the PNC corrections also makes sure that γij¯\gamma_{i\bar{j}} are real and of order 1.

Substituting the coefficients χij\chi_{ij} and χij¯\chi_{i\bar{j}} in Eqs. (47) into Eq. (39b), one obtains

j¯|Δh|i=iη(εiεj¯)γij¯,\bra{\bar{j}}\Delta h\ket{i}={\rm i}\eta(\varepsilon_{i}-\varepsilon_{\bar{j}})\gamma_{i\bar{j}}\,, (48)

which is the matrix equivalence of Eq. (45). We now need to solve Eq. (48) for the unknown coefficients γij¯\gamma_{i\bar{j}}. For this purpose, we need to express the matrix element j¯|Δh|i\bra{\bar{j}}\Delta h\ket{i} in terms of the coefficients γij¯\gamma_{i\bar{j}}. Substituting Eqs. (47) into Eq. (IV.2) and replacing jj with ii therein, we find

j¯|Δh|i\displaystyle\bra{\bar{j}}\Delta h\ket{i} =iη[Sj¯iak¯γak¯(gj¯ak¯igj¯k¯ai)],\displaystyle={\rm i}\eta\left[S_{\bar{j}i}-\sum_{a\bar{k}}\gamma_{a\bar{k}}(g_{\bar{j}a\bar{k}i}-g_{\bar{j}\bar{k}ai})\right]\,, (49)

where the summation runs over all PP core orbitals ψaP\psi^{P}_{a} and all PP orbitals ψk¯P\psi^{P}_{\bar{k}} which have the same total angular momentum but opposite parity to ψaP\psi^{P}_{a}

Substituting Eq. (49) into Eq. (48), one obtains

Sj¯iak¯γak¯(gj¯ak¯igj¯k¯ai)=(εiεj¯)γij¯.S_{\bar{j}i}-\sum_{a\bar{k}}\gamma_{a\bar{k}}(g_{\bar{j}a\bar{k}i}-g_{\bar{j}\bar{k}ai})=(\varepsilon_{i}-\varepsilon_{\bar{j}})\gamma_{i\bar{j}}\,. (50)

Remember that in Eq. (50), the orbitals j¯\bar{j} have the same total angular momentum but opposite parity to the orbital ii whereas the orbitals k¯\bar{k} have the same total angular momentum but opposite parity to the orbital aa. Equation (50) allows us to solve for the PNC mixing coefficients γij¯\gamma_{i\bar{j}}. It is the matrix version of the finite-difference equations (25). In contrast with Eq. (40), it is independent of the small parameter η\eta so is not subject to the issue with numerical inaccuracy as was the method described in Sec. IV.2.

Let us consider the case where i=bi=b, i.e., a core orbital. Denote by NcN_{c} the number of core orbitals. We may then arrange all the coefficients γbj¯\gamma_{b\bar{j}} into a vector 𝜸b\boldsymbol{\gamma}_{b} of length 2NNc2NN_{c}, all the quantities Sj¯bS_{\bar{j}b} into a vector 𝐒𝐛\bf S_{b} of length 2NNc2NN_{c}, all the quantities εbεj¯\varepsilon_{b}-\varepsilon_{\bar{j}} into a diagonal matrix Δεb\Delta\varepsilon_{b} of size 2NNc×2NNc2NN_{c}\times 2NN_{c} and all the quantities gj¯ak¯bgj¯k¯abg_{\bar{j}a\bar{k}b}-g_{\bar{j}\bar{k}ab} into a matrix 𝑮b\boldsymbol{G}_{b} of size 2NNc×2NNc2NN_{c}\times 2NN_{c}. As a result, Eq. (50) may be written in a more suggestive form as

𝐒b𝑮b𝜸b=Δεb𝜸b,{\bf S}_{b}-\boldsymbol{G}_{b}\boldsymbol{\gamma}_{b}=\Delta\varepsilon_{b}\boldsymbol{\gamma}_{b}\,, (51)

whose solution reads

𝜸b=(Δεb+𝑮b)1𝐒b.\boldsymbol{\gamma}_{b}=(\Delta\varepsilon_{b}+\boldsymbol{G}_{b})^{-1}{\bf S}_{b}\,. (52)

Equation (52) allows us to obtain the mixing coefficients γbj¯\gamma_{b\bar{j}} for all core orbitals. We point out that Eq. (52) is linear so there is no need for an iterative scheme as with the methods discussed in Secs. IV.1 and IV.2.

After solving for the PNC mixing coefficients γbj¯\gamma_{b\bar{j}} of all NcN_{c} core orbitals, we again use Eq. (50) to solve for the mixing coefficients of all unoccupied orbitals ψm\psi^{\prime}_{m}, obtaining

γmj¯=Sj¯maj¯γak¯(gj¯ak¯mgj¯k¯am)εmεj¯.\gamma_{m\bar{j}}=\frac{S_{\bar{j}m}-\sum_{a\bar{j}}\gamma_{a\bar{k}}(g_{\bar{j}a\bar{k}m}-g_{\bar{j}\bar{k}am})}{\varepsilon_{m}-\varepsilon_{\bar{j}}}\,. (53)

In this form, Eq. (53) clearly demonstrates the perturbative nature of the current approach. As a result, during computation, one should check that accidental degeneracy does not happen, or in other words, that the coefficients |ηγmj¯|1|\eta\gamma_{m\bar{j}}|\ll 1. If such event does occur, the more general method described in Sec. IV.2 should be used instead.

We used the perturbative matrix method discussed in this subsection to generate for Cs133{}^{133}\mathrm{Cs} a PM basis of total angular momenta ranging from 1/21/2 to 13/213/2. The PP set used to expand the PM orbitals are the same as that used in Sec. IV.2. The lowest order 6S1/27S1/26S_{1/2}-7S_{1/2} PNC frozen-core and core-perturbed amplitudes for Cs computed using the so-obtained PM-DHF valence orbitals ψ6s1/2\psi^{\prime}_{6s_{1/2}} and ψ6s1/2\psi^{\prime}_{6s_{1/2}} are, respectively

EPVfc\displaystyle E^{\rm fc}_{\rm PV} =0.73947×1011i|e|a0(QW/N),\displaystyle=0.73947\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,, (54)
EPVcp\displaystyle E^{\rm cp}_{\rm PV} =0.92697×1011i|e|a0(QW/N).\displaystyle=0.92697\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,.

The small differences between the results (44) and (54) of the two matrix methods may be attributed to nonlinear O(η2)O(\eta^{2}) terms, which, although small, may propagate through the computation. At the level of 0.004%, these numerical differences are acceptable for our goals as we ultimately aim at 0.2% overall accuracy in the PNC amplitude.

IV.4 Numerical stability of parity-mixed basis sets

In the previous sections, we have presented different methods through which basis sets of PM single-electron orbitals may be obtained. Before discussing the application of these basis sets in MBPT and CC calculations, we pause here to make a few remarks regarding their numerical stability, specifically with respect to the small parameter η\eta.

In the finite difference and perturbative matrix methods, a PM single-electron orbital is expanded into two components of opposite parities: a real part being independent of η\eta and an imaginary part having a linear dependence on η\eta. Furthermore, as was shown in Secs. IV.1 and IV.3, the factor η\eta may be completely separated from the imaginary part, allowing one to reliably compute this component. At the DHF level, the PNC transition amplitude EPVE_{\rm PV} obtained using the resulting PM orbitals reads

EPV\displaystyle E_{\rm PV} =ψ6s1/2|Dz|ψ7s1/2\displaystyle=\bra{\psi^{\prime}_{6s_{1/2}}}D_{z}\ket{\psi^{\prime}_{7s_{1/2}}}
=iη(ψ6s1/2|Dz|ψ¯7s1/2ψ¯6s1/2|Dz|ψ7s1/2),\displaystyle={\rm i}\eta(\bra{\psi_{6s_{1/2}}}D_{z}\ket{\bar{\psi}_{7s_{1/2}}}-\bra{\bar{\psi}_{6s_{1/2}}}D_{z}\ket{\psi_{7s_{1/2}}})\,, (55)

which shows that EPVE_{\rm PV} depends linearly on η\eta.

In contrast, if the exact matrix diagonalization method is used, the resulting PM single-electron orbitals contain, in principle, nonlinear dependence on η\eta. As remarked in Sec. IV.2, this is due to the need of solving the nonlinear eigenvalue Eq. (40). As a result, the PNC transition amplitude EPVE_{\rm PV}, computed as in Eq. (IV.4) will also contain contributions nonlinear in η\eta. However, these nonlinear contributions are not manifest at the level of accuracy we are interested in, as may be observed from Fig. 1, which shows the linear dependence on η\eta of the PNC transition amplitudes EPVfcE^{\rm fc}_{\rm PV} and EPVcpE^{\rm cp}_{\rm PV} calculated using the exact matrix diagonalization method.

Refer to caption
Figure 1: (Color online) The dependence on the dimensionless parameter ηGFQW/(22a02)\eta\equiv G_{F}Q_{W}/(2\sqrt{2}a_{0}^{2}) of the 6S1/27S1/26S_{1/2}-7S_{1/2} PNC transition amplitudes (in both frozen core and core perturbed approximations) in Cs133{}^{133}\mathrm{Cs} calculated using the exact matrix diagonalization method described in Sec. IV.2. The lines being straight demonstrate that the effects that are nonlinear in η\eta do not show when computing EPVE_{\rm PV}.

Similarly, it may be argued that when PM single-electron orbitals are used in the MBPT and CC computations, terms that are O(η2)O(\eta^{2}) or higher do not contribute numerically. This justifies our direct upgrade of the conventional PP-MBPT and PP-CC formalism to the PM ones without having first to linearize their equations in terms of η\eta. At the desired level of numerical accuracy 0.2%\ll 0.2\%, contributions that are O(η2)O(\eta^{2}) or higher simply do not show up.

We end this section by elaborating on the advantage of the exact matrix diagonalization method in the case of accidental (near) degeneracy between states with the same angular momentum but opposite parities. We stress that this degeneracy can appear as an artefact of using finite basis set of orbitals (pseudo-spectrum). Two orbitals ψ1\psi_{1} and ψ2\psi_{2} are considered to be nearly degenerate if the perturbation theory convergence criterion, |ψ1|hW|ψ2/(ε1ε2)|1|\bra{\psi_{1}}h_{W}\ket{\psi_{2}}/(\varepsilon_{1}-\varepsilon_{2})|\ll 1, fails. This problem may be avoided by varying the parameters of the basis set, such as the radius of the cavity, so as to make all quantities of the |ψ1|hW|ψ2/(ε1ε2)||\bra{\psi_{1}}h_{W}\ket{\psi_{2}}/(\varepsilon_{1}-\varepsilon_{2})| form to be much smaller than 1. We test our numerical sets for these accidental degeneracies before applying the perturbative approach.

Alternatively, this tuning of the basis set parameters may be avoided by using matrix diagonalization: quantities of the form |ψ1|hW|ψ2/(ε1ε2)||\bra{\psi_{1}}h_{W}\ket{\psi_{2}}/(\varepsilon_{1}-\varepsilon_{2})| do not arise in this method. It is worth noting also that in this case, the lifting of degeneracy by hWh_{W} is O(η)O(\eta). For example, consider again two states ψ1\psi_{1} and ψ2\psi_{2} of the same total angular momentum, opposite parities, and energies ε1ε2=ε\varepsilon_{1}\approx\varepsilon_{2}=\varepsilon. To find the energy corrections due to the perturbation hWh_{W}, one solves the secular equation for the perturbed energy ε\varepsilon^{\prime}

det(εεψ1|hW|ψ2ψ2|hW|ψ2εε)=0,{\rm det}\begin{pmatrix}\varepsilon-\varepsilon^{\prime}&&\bra{\psi_{1}}h_{W}\ket{\psi_{2}}\\ \bra{\psi_{2}}h_{W}\ket{\psi_{2}}&&\varepsilon-\varepsilon^{\prime}\end{pmatrix}=0\,, (56)

obtaining

ε=ε±|ψ1|hW|ψ2|,\varepsilon^{\prime}=\varepsilon\pm|\bra{\psi_{1}}h_{W}\ket{\psi_{2}}|\,, (57)

which shows that the energy corrections are O(η)O(\eta) for degenerate states. Note that in this case, strictly speaking, one also needs to include the natural decay widths Γi\Gamma_{i} to the energy levels, εiεiiΓi/2\varepsilon_{i}\rightarrow\varepsilon_{i}-{\rm i}\Gamma_{i}/2 which can lift the degeneracy and requires further modifications to the code.

V Matrix elements in the parity-mixed basis

Now with the PM basis constructed, we go back to the MBPT formalism of Sec. II. The basic building blocks of MBPT expressions are the matrix elements of one-body (e.g., the electric dipole) and two-body (e.g., Coulomb interaction) operators in the PM basis, ψi=ψi+iηψ¯i\psi^{\prime}_{i}=\psi_{i}+{\rm i}\eta\bar{\psi}_{i}. Due to the smallness of the parameter η\eta, we may linearize the resulting expressions in η\eta. Then any matrix element of an operator of definite parity splits into a part involving only the PP orbitals ψi\psi_{i} and a correction that involves opposite parity admixtures ψ¯i\bar{\psi}_{i} (PNC correction). The former is already implemented in traditional MBPT codes. The latter may be readily added to these codes by modifying parity selection rules and using the radial components of ψ¯i\bar{\psi}_{i}.

Furthermore, we show that the matrix elements of any operator of definite parity in the PM basis is either purely real- or imaginary-valued. With our phase convention for the PM radial components (22), PP parts are always real, while the PNC corrections are always imaginary. We will exploit this fact to derive useful symmetries of the reduced matrix elements of the one- and two-body operators. These symmetries will help significantly reduce the amount of computation and storage needed in MBPT calculations.

V.1 Angular reduction of matrix elements

Let us begin by discussing the angular reduction of the PM matrix elements of one- and two-body operators. The particular operators of interest here are of course the electric dipole operator and the inter-electron Coulomb interaction.

Without loss of generality, we assume that the operators in question are Hermitian and can be represented as components of irreducible tensor operators. We also observe that the PM orbitals are eigen-states of the total angular momentum operators JzJ_{z} and 𝐉2\mathbf{J}^{2}. As a result, the Wigner-Eckart theorem applies to the matrix elements of the one- and two-body operators with respect to these PM orbitals. Moreover, since the weak interaction is a pseudo-scalar, a PM orbital has the very same total angular momentum as its PP counterpart.

As a result, the angular reduction of a one-body matrix element tijt^{\prime}_{ij} has the same form as that of the PP tijt_{ij}. More explicitly, for the case where t=zt=z, the Wigner-Eckart theorem states

zkl=(1)jkmk(jk1jlmk0ml)kzl,\displaystyle z^{\prime}_{kl}=(-1)^{j_{k}-m_{k}}\begin{pmatrix}j_{k}&1&j_{l}\\ -m_{k}&0&m_{l}\end{pmatrix}\langle k||z||l\rangle^{\prime}\,, (58)

where all the information about mixing parities is contained in the reduced matrix element kzl\langle k||z||l\rangle^{\prime}.

Similarly, the angular reduction of the PM Coulomb integrals gijklg^{\prime}_{ijkl} is the same as that for the PP gijklg_{ijkl}, namely

gijkl\displaystyle g^{\prime}_{ijkl} =LJL(ijkl)XL(ijkl),\displaystyle=\sum_{L}J_{L}(ijkl)X^{\prime}_{L}(ijkl)\,, (59)
JL(ijkl)\displaystyle J_{L}(ijkl) M(1)jimi+jjmj\displaystyle\equiv\sum_{M}(-1)^{j_{i}-m_{i}+j_{j}-m_{j}}
×(jiLjkmiMmk)(jjLjlmjMml),\displaystyle\times\begin{pmatrix}j_{i}&L&j_{k}\\ -m_{i}&-M&m_{k}\end{pmatrix}\begin{pmatrix}j_{j}&L&j_{l}\\ -m_{j}&M&m_{l}\end{pmatrix}\,,

where all the information about mixing parities is contained in the reduced matrix element XL(ijkl)X^{\prime}_{L}(ijkl).

We may write the PM reduced matrix elements of a one-body operator tt as (since ψi=ψi+iηψ¯i\psi^{\prime}_{i}=\psi_{i}+i\eta\bar{\psi}_{i})

ktl\displaystyle\langle k||t||l\rangle^{\prime} =ktl+iηktl′′,\displaystyle=\langle k||t||l\rangle+{\rm i}\eta\langle k||t||l\rangle^{\prime\prime}\,, (60a)
ktl′′\displaystyle\langle k||t||l\rangle^{\prime\prime} ktl¯k¯tl,\displaystyle\equiv\langle k||t||\bar{l}\rangle-\langle\bar{k}||t||l\rangle\,, (60b)

where have we dropped O(η2)O(\eta^{2}) terms. Explicitly, for an electric-dipole operator zz, relevant to computing PNC amplitudes, the three reduced matrix elements read

kzl\displaystyle\langle k||z||l\rangle =κkC1κl(PkPl+QkQl)r𝑑r,\displaystyle=\langle\kappa_{k}||C_{1}||\kappa_{l}\rangle\int\left(P_{k}P_{l}+Q_{k}Q_{l}\right)rdr\,, (61)
kzl¯\displaystyle\langle k||z||\bar{l}\rangle =κkC1κl¯(PkP¯l+QkQ¯l)r𝑑r,\displaystyle=\langle\kappa_{k}||C_{1}||\kappa_{\bar{l}}\rangle\int\left(P_{k}\bar{P}_{l}+Q_{k}\bar{Q}_{l}\right)rdr\,,
k¯zl\displaystyle\langle\bar{k}||z||l\rangle =κk¯C1κl(P¯kPl+Q¯kQl)r𝑑r.\displaystyle=\langle\kappa_{\bar{k}}||C_{1}||\kappa_{l}\rangle\int\left(\bar{P}_{k}P_{l}+\bar{Q}_{k}Q_{l}\right)rdr\,.

Here kzl\langle k||z||l\rangle is the PP contribution, and the kzl¯\langle k||z||\bar{l}\rangle and k¯zl\langle\bar{k}||z||l\rangle contributions are due to the opposite parity admixtures as indicated by the large and small radial components with overhead bars, c.f. Eq. (22).

The parity selection rules are encoded into the reduced matrix elements of the normalized spherical harmonic via κkC1κlmod2(k+l)\langle\kappa_{k}||C_{1}||\kappa_{l}\rangle\propto{\rm mod}_{2}(\ell_{k}+\ell_{l}), i.e. k+l\ell_{k}+\ell_{l} must be odd. Similar selection rules apply to the P-odd corrections, e.g., κk¯C1κlmod2(k+l+1)\langle\kappa_{\bar{k}}||C_{1}||\kappa_{l}\rangle\propto{\rm mod}_{2}(\ell_{k}+\ell_{l}+1), since k¯=k±1\ell_{\bar{k}}=\ell_{k}\pm 1, i.e. k+l\ell_{k}+\ell_{l} must be even. Note that for two fixed PM orbitals, these selection rules cannot be satisfied simultaneously, thereby, the matrix element is either pure real- or imaginary-valued. With our phase convention for the PM radial components, the PP part is always real, while the PNC correction is always imaginary. The above statements can be easily generalized to any irreducible tensor operator of definite parity.

Similar considerations apply to the reduced Coulomb matrix element:

XL(ijkl)\displaystyle X^{\prime}_{L}(ijkl) =XL(ijkl)+iηXL′′(ijkl),\displaystyle=X_{L}(ijkl)+{\rm i}\eta X^{\prime\prime}_{L}(ijkl)\,, (62a)
XL′′(ijkl)\displaystyle X^{\prime\prime}_{L}(ijkl) XL(ijk¯l)+XL(ijkl¯)\displaystyle\equiv X_{L}(ij\bar{k}l)+X_{L}(ijk\bar{l})
XL(i¯jkl)XL(ij¯kl).\displaystyle-X_{L}(\bar{i}jkl)-X_{L}(i\bar{j}kl)\,. (62b)

Here, the quantity XL(ijkl)X_{L}(ijkl) is expressed in terms of the reduced matrix element of the normalized spherical harmonic CL(𝐫^)C_{L}(\hat{\bf r}) and the Slater integral RL(ijkl)R_{L}(ijkl):

XL(ijkl)=(1)LκiCLκkκjCLκlRL(ijkl).X_{L}(ijkl)=(-1)^{L}\langle\kappa_{i}||C_{L}||\kappa_{k}\rangle\langle\kappa_{j}||C_{L}||\kappa_{l}\rangle R_{L}(ijkl)\,. (63)

The parity selection rules for XL(ijkl)X_{L}(ijkl) are (1)i+L+k=+1(-1)^{\ell_{i}+L+\ell_{k}}=+1 and (1)j+L+l=+1(-1)^{\ell_{j}+L+\ell_{l}}=+1.

The quantities in Eq. (62b) are defined similarly. For example,

XL(i¯jkl)=(1)Lκi¯CLκkκjCLκlRL(i¯jkl),X_{L}(\bar{i}jkl)=(-1)^{L}\langle\kappa_{\bar{i}}||C_{L}||\kappa_{k}\rangle\langle\kappa_{j}||C_{L}||\kappa_{l}\rangle R_{L}(\bar{i}jkl)\,, (64)

where the index i¯\bar{i} in RL(i¯jkl)R_{L}(\bar{i}jkl) means that we use the radial functions P¯i\bar{P}_{i} and Q¯i\bar{Q}_{i} as defined in Eq. (22g).

The parity selection rules for the various terms in Eqs. (62a) and (62b) are also clear. If i+L+k\ell_{i}+L+\ell_{k} and j+L+l\ell_{j}+L+\ell_{l} are both even then XL(ijkl)=XL(ijkl)X^{\prime}_{L}(ijkl)=X_{L}(ijkl) which is purely real whereas if they are both odd then XL(ijkl)=0X^{\prime}_{L}(ijkl)=0. If i+L+k\ell_{i}+L+\ell_{k} is odd but j+L+l\ell_{j}+L+\ell_{l} is even then XL(ijkl)=iηXL′′(ijkl)X^{\prime}_{L}(ijkl)={\rm i}\eta X^{\prime\prime}_{L}(ijkl) is purely imaginary and XL′′(ijkl)X^{\prime\prime}_{L}(ijkl) is given by the first two terms in Eq. (62b). On the other hand, if i+L+k\ell_{i}+L+\ell_{k} is even but j+L+l\ell_{j}+L+\ell_{l} is odd then XL′′(ijkl)X^{\prime\prime}_{L}(ijkl) is given by the last two terms in Eq. (62b). Translating to gijklg^{\prime}_{ijkl}, these rules mean that if i+j+k+l\ell_{i}+\ell_{j}+\ell_{k}+\ell_{l} is even then gijklg^{\prime}_{ijkl} is real and equals its PP counterpart gijklg_{ijkl}, whereas if i+j+k+l\ell_{i}+\ell_{j}+\ell_{k}+\ell_{l} is odd then gijklg^{\prime}_{ijkl} is purely imaginary. These observations will prove useful in the formulation of the PM-CC formalism, Sec VII.

Another frequently occurring matrix element is that of the anti-symmetrized Coulomb interaction, g~ijklgijklgijlk\tilde{g}^{\prime}_{ijkl}\equiv{g}^{\prime}_{ijkl}-{g}^{\prime}_{ijlk}, which can be brought in the angular-diagram form identical to that of gijkl{g}^{\prime}_{ijkl}, c.f. Eq. (59),

g~ijklgijklgijlk\displaystyle\tilde{g}^{\prime}_{ijkl}\equiv{g}^{\prime}_{ijkl}-{g}^{\prime}_{ijlk} =LJL(ijkl)ZL(ijkl),\displaystyle=\sum_{L}J_{L}(ijkl)Z^{\prime}_{L}(ijkl)\,, (65)

where the reduced matrix element is given by

ZL(ijkl)\displaystyle Z^{\prime}_{L}(ijkl) =ZL(ijkl)+iηZL′′(ijkl)\displaystyle=Z_{L}(ijkl)+{\rm i}\eta Z^{\prime\prime}_{L}(ijkl) (66a)
ZL′′(ijkl)\displaystyle Z^{\prime\prime}_{L}(ijkl) ZL(ijk¯l)+ZL(ijkl¯)\displaystyle\equiv Z_{L}(ij\bar{k}l)+Z_{L}(ijk\bar{l})
ZL(i¯jkl)ZL(ij¯kl).\displaystyle-Z_{L}(\bar{i}jkl)-Z_{L}(i\bar{j}kl)\,. (66b)

In these equations, ZL(ijkl)Z_{L}(ijkl) may be expressed in terms of XL(ijkl)X_{L}(ijkl) via

ZL(ijkl)\displaystyle Z_{L}(ijkl) =XL(ijkl)\displaystyle=X_{L}(ijkl)
+L(2L+1){jkjiLjljjL}XL(ijlk),\displaystyle+\sum_{L^{\prime}}(2L+1)\begin{Bmatrix}j_{k}&j_{i}&L\\ j_{l}&j_{j}&L^{\prime}\end{Bmatrix}X_{L^{\prime}}(ijlk)\,, (67)

and the other quantities are defined similarly. Again, the overhead bars in Eq. (66b) signify the use of the PP-odd radial functions P¯i\bar{P}_{i} and Q¯i\bar{Q}_{i} as defined in Eq. (22g). The parity selection rules for gijklg^{\prime}_{ijkl} also apply to g~ijkl\tilde{g}^{\prime}_{ijkl}, namely, g~ijkl\tilde{g}^{\prime}_{ijkl} is real and equals its PP counterpart if i+j+k+l\ell_{i}+\ell_{j}+\ell_{k}+\ell_{l} is even whereas g~ijkl\tilde{g}^{\prime}_{ijkl} is purely imaginary if i+j+k+l\ell_{i}+\ell_{j}+\ell_{k}+\ell_{l} is odd.

To reiterate, we observe that the PM matrix elements, Eqs. (58) and (65), split into real PP parts and purely imaginary PNC parts. Due to the small coefficient η\eta, the imaginary parts are many orders of magnitude smaller than the real parts. This, however, does not give rise to a problem with numerical accuracy due to truncation as mentioned in Sec. IV.2. In fact, if the MBPT code is modified to use complex instead of real numbers and the PP and PNC parts are stored separately then algebraic operations will always involve adding terms of the same order of magnitude. In Sec. VI, we shall present how such a procedure is carried out with the example of the random-phase approximation (RPA).

V.2 Symmetries of reduced matrix elements

Before starting with the discussion of RPA, however, let us present the symmetries of the reduced matrix elements kzl\langle k||z||l\rangle^{\prime}, XL(ijkl)X^{\prime}_{L}(ijkl) and ZL(ijkl)Z^{\prime}_{L}(ijkl) with respect to the exchange of their indices. In a conventional MBPT formalism which uses PP single-electron orbitals, the corresponding symmetries of the PP reduced matrix elements are exploited to a great extent to significantly reduce the amount of computation and storage needed. We will show that similar symmetries are also available for a MBPT scheme using PM orbitals so the same economy may be achieved.

We begin with the matrix elements of a one-body operator. For our purpose, we concentrate on the electric dipole operator zz. Using the definitions (61), (63), and (V.1) and the following property of the reduced matrix elements of the normalized spherical harmonics

κkCLκl=(1)jkjlκlCLκk,\langle\kappa_{k}||C_{L}||\kappa_{l}\rangle=(-1)^{j_{k}-j_{l}}\langle\kappa_{l}||C_{L}||\kappa_{k}\rangle\,, (68)

it may be verified that the PP reduced matrix elements of zz satisfy

kzl=(1)jkjllzk.\langle k||z||l\rangle=(-1)^{j_{k}-j_{l}}\langle l||z||k\rangle\,. (69)

Next, by using Eqs. (60a) and (69), we find the following symmetry for the PM reduced matrix elements of the electric dipole moment

kzl=(1)jkjl[lzk],\langle k||z||l\rangle^{\prime}=(-1)^{j_{k}-j_{l}}[\langle l||z||k\rangle^{\prime}]^{*}\,, (70)

where * denotes complex conjugation.

We note that although we only considered the electric dipole operator zz, the result presented above applies to any single-electron irreducible tensor operator of rank kk, T(k)T^{(k)}, namely

kT(k)l=(1)jkjl[lT(k)k].\langle k||T^{(k)}||l\rangle^{\prime}=(-1)^{j_{k}-j_{l}}[\langle l||T^{(k)}||k\rangle^{\prime}]^{*}\,. (71)

We now turn to the reduced matrix elements of the inter-electron Coulomb interaction. We begin by presenting the familiar symmetries of the PP reduced matrix elements XL(ijkl)X_{L}(ijkl) and ZL(ijkl)Z_{L}(ijkl). Although these results are not new, they serve as a convenient reference point for our discussion of the PM matrix elements. Using the definitions (63) and (V.1) and the property (68), one easily finds the following relations

XL(ijkl)\displaystyle X_{L}(ijkl) =XL(jilk),\displaystyle=X_{L}(jilk)\,, (72a)
XL(ijkl)\displaystyle X_{L}(ijkl) =(1)jijkXL(kjil),\displaystyle=(-1)^{j_{i}-j_{k}}X_{L}(kjil)\,, (72b)
XL(ijkl)\displaystyle X_{L}(ijkl) =(1)ji+jj+jk+jlXL(klij),\displaystyle=(-1)^{j_{i}+j_{j}+j_{k}+j_{l}}X_{L}(klij)\,, (72c)
ZL(ijkl)\displaystyle Z_{L}(ijkl) =ZL(jilk),\displaystyle=Z_{L}(jilk)\,, (72d)
ZL(ijkl)\displaystyle Z_{L}(ijkl) =(1)ji+jj+jk+jlZL(klij),\displaystyle=(-1)^{j_{i}+j_{j}+j_{k}+j_{l}}Z_{L}(klij)\,, (72e)
ZL(ijkl)\displaystyle Z_{L}(ijkl) =[L]L{jjjlLjijkL}ZL(jikl),\displaystyle=[L]\sum_{L^{\prime}}\left\{\begin{matrix}j_{j}&&j_{l}&&L\\ j_{i}&&j_{k}&&L^{\prime}\end{matrix}\right\}Z_{L^{\prime}}(jikl)\,, (72f)

where [L]2L+1[L]\equiv 2L+1 and {jjjlLjijkL}\left\{\begin{matrix}j_{j}&&j_{l}&&L\\ j_{i}&&j_{k}&&L^{\prime}\end{matrix}\right\} is the 6j6j-symbol.

From the expansions (62a) and (66a) for XL(ijkl)X^{\prime}_{L}(ijkl) and ZL(ijkl)Z^{\prime}_{L}(ijkl) and the properties (72a) and (72d), one sees that simultaneously swapping iji\leftrightarrow j and klk\leftrightarrow l has no effect on the Coulomb reduced matrix elements, i.e.,

XL(ijkl)\displaystyle X^{\prime}_{L}(ijkl) =XL(jilk),\displaystyle=X^{\prime}_{L}(jilk)\,, (73a)
ZL(ijkl)\displaystyle Z^{\prime}_{L}(ijkl) =ZL(jilk).\displaystyle=Z^{\prime}_{L}(jilk)\,. (73b)

It may also be observed from Eqs. (72c) and (72e) that swapping the pair ijklij\leftrightarrow kl is equivalent to introducing the phase factor (1)ji+jj+jk+jl(-1)^{j_{i}+j_{j}+j_{k}+j_{l}} to XL(ijkl)X^{\prime}_{L}(ijkl) and ZL(ijkl)Z^{\prime}_{L}(ijkl) as well as switching the sign of XL′′(ijkl)X^{\prime\prime}_{L}(ijkl) and ZL′′(ijkl)Z^{\prime\prime}_{L}(ijkl). As a result, we have

XL(ijkl)\displaystyle X^{\prime}_{L}(ijkl) =(1)ji+jj+jk+jl[XL(klij)],\displaystyle=(-1)^{j_{i}+j_{j}+j_{k}+j_{l}}[X^{\prime}_{L}(klij)]^{*}\,, (74a)
ZL(ijkl)\displaystyle Z^{\prime}_{L}(ijkl) =(1)ji+jj+jk+jl[ZL(klij)].\displaystyle=(-1)^{j_{i}+j_{j}+j_{k}+j_{l}}[Z^{\prime}_{L}(klij)]^{*}\,. (74b)

Next, we present the PM equivalence of Eq. (72b). For this purpose, it is convenient to consider two separate cases. First, let us assume that the nominal parities of the orbitals ψi\psi^{\prime}_{i} and ψk\psi^{\prime}_{k} satisfy the condition (1)i+L+k=1(-1)^{\ell_{i}+L+\ell_{k}}=1. In this case, Eq. (62a) simplifies to

XL(ijkl)=XL(ijkl)+iη[XL(ijkl¯)XL(ij¯kl)],\displaystyle X^{\prime}_{L}(ijkl)=X_{L}(ijkl)+{\rm i}\eta\left[X_{L}(ijk\bar{l})-X_{L}(i\bar{j}kl)\right]\,, (75)

which, when combined with Eq. (72b), gives

XL(ijkl)=(1)jijkXL(kjil),X^{\prime}_{L}(ijkl)=(-1)^{j_{i}-j_{k}}X^{\prime}_{L}(kjil)\,, (76)

if (1)i+L+k=1(-1)^{\ell_{i}+L+\ell_{k}}=1. On the other hand, if (1)i+L+k=1(-1)^{\ell_{i}+L+\ell_{k}}=-1 then Eq. (62a) simplifies to

XL(ijkl)=iη[XL(ijk¯l)XL(i¯jkl)].X^{\prime}_{L}(ijkl)={\rm i}\eta\left[X_{L}(ij\bar{k}l)-X_{L}(\bar{i}jkl)\right]\,. (77)

It is clear from Eq. (77) that in this case, swapping iki\leftrightarrow k introduces the factor (1)jijk(-1)^{j_{i}-j_{k}} as well as a minus sign. Thus, we have

XL(ijkl)=(1)jijkXL(kjil).X^{\prime}_{L}(ijkl)=-(-1)^{j_{i}-j_{k}}X^{\prime}_{L}(kjil)\,. (78)

if (1)i+L+k=1(-1)^{\ell_{i}+L+\ell_{k}}=-1. We may combine Eqs. (76) and (78) into a single formula, writing

XL(ijkl)=(1)i+L+k(1)jijkXL(kjil),\displaystyle X^{\prime}_{L}(ijkl)=(-1)^{\ell_{i}+L+\ell_{k}}(-1)^{j_{i}-j_{k}}X^{\prime}_{L}(kjil)\,, (79)

which is the PM equivalence of Eq. (72b).

Finally, since the recoupling rule Eq. (72f) involves only total angular momenta and no sign change, its PM version has the same form, i.e.,

ZL(ijkl)=[L]L{jjjlLjijkL}ZL(jikl).Z^{\prime}_{L}(ijkl)=[L]\sum_{L^{\prime}}\left\{\begin{matrix}j_{j}&&j_{l}&&L\\ j_{i}&&j_{k}&&L^{\prime}\end{matrix}\right\}Z^{\prime}_{L^{\prime}}(jikl)\,. (80)

Equations (70), (73), (74), (79) and (80) represent the symmetries of the PM reduced matrix elements of the electric dipole and inter-electron Coulomb interaction operators with respect to permutations of the PM orbitals. They will be used extensively in the PM-MBPT as well as PM-CC calculations.

VI Random-phase approximation for the parity non-conserving amplitude

In Sec. IV we presented several methods through which basis sets of PM orbitals may be constructed. The numerical accuracy of these basis sets was tested by computing the PNC amplitude between the PM-DHF valence states. Strictly speaking, this test only involves two single-electron PM-DHF valence orbitals, |6s1/2|6s_{1/2}^{\prime}\rangle and |7s1/2|7s_{1/2}^{\prime}\rangle. In Sec. V we discussed formulas for the matrix elements of one- and two-body operators, in particular the electric dipole and Coulomb operators, in terms of the PM-DHF bases. These matrix elements are needed in the MBPT paradigm to take into account the effects of inter-electron correlation on the PNC amplitude. In this section, we shall use these formulas to compute the second-order and RPA all-order correlation corrections to the matrix elements of the electric dipole operator.

The relevant second-order formula for PNC amplitude is given in Eq. (II) and it involves summations over the entire PM-DHF basis set. Here, using this formula, we test the accuracy of our generated PM-DHF basis sets by computing PNC amplitude in Cs133{}^{133}\mathrm{Cs} in the well-established random-phase approximation (RPA) [57, 30, 31, 64]. RPA sums diagrams topologically similar to second-order Eq. (II) to all orders of MBPT. This not only tests the quality of PM-DHF basis sets, but importantly builds the foundation for the formulation of parity-mixed coupled-cluster (PM-CC) method, which systematically enables all-order summations of substantially larger classes of diagrams, see Sec. VII.

For now, we focus on the RPA method. In this approximation, one first takes into account the second-order correction to the “core-to-excited” matrix elements tant^{\prime}_{an} and tnat^{\prime}_{na} present in Eq. (II). Denoted by tanRPAt^{\prime\rm RPA}_{an} and tnaRPAt^{\prime\rm RPA}_{na} the RPA vertices, one finds that these quantities satisfy equations similar to Eq. (II), namely

tanRPA\displaystyle t^{\prime\rm RPA}_{an} =tan+bmtbmRPAg~amnbεbεmω+bmg~abnmtmbRPAεbεm+ω,\displaystyle=t^{\prime}_{an}+\sum_{bm}\frac{t^{\prime\rm RPA}_{bm}\tilde{g}^{\prime}_{amnb}}{\varepsilon^{\prime}_{b}-\varepsilon^{\prime}_{m}-\omega}+\sum_{bm}\frac{\tilde{g}^{\prime}_{abnm}t^{\prime\rm RPA}_{mb}}{\varepsilon^{\prime}_{b}-\varepsilon^{\prime}_{m}+\omega}\,, (81)
tnaRPA\displaystyle t^{\prime\rm RPA}_{na} =tna+bmtbmRPAg~nmabεbεmω+bmg~nbamtmbRPAεbεm+ω,\displaystyle=t^{\prime}_{na}+\sum_{bm}\frac{t^{\prime\rm RPA}_{bm}\tilde{g}^{\prime}_{nmab}}{\varepsilon^{\prime}_{b}-\varepsilon^{\prime}_{m}-\omega}+\sum_{bm}\frac{\tilde{g}^{\prime}_{nbam}t^{\prime\rm RPA}_{mb}}{\varepsilon^{\prime}_{b}-\varepsilon^{\prime}_{m}+\omega}\,,

which will be solved by iteration to convergence. Once the RPA vertices are obtained, the matrix elements between two valence orbitals ψv\psi^{\prime}_{v} and ψw\psi^{\prime}_{w} are given by

Twv=twv+antanRPAg~wnvaεaεnω+ang~wvnatnaRPAεaεn+ω.\displaystyle T^{\prime}_{wv}=t^{\prime}_{wv}+\sum_{an}\frac{t^{\prime\rm RPA}_{an}\tilde{g}^{\prime}_{wnva}}{\varepsilon^{\prime}_{a}-\varepsilon^{\prime}_{n}-\omega}+\sum_{an}\frac{\tilde{g}^{\prime}_{wvna}t^{\prime\rm RPA}_{na}}{\varepsilon^{\prime}_{a}-\varepsilon^{\prime}_{n}+\omega}\,. (82)

For computations of Cs133{}^{133}\mathrm{Cs} PNC amplitudes, tt is the electric-dipole operator, and v=6sv=6s, w=7sw=7s.

We used the PM-DHF basis sets of Sec. IV to compute the RPA correction to the 6S1/27S1/26S_{1/2}-7S_{1/2} PNC transition amplitude in Cs. The forms of the dipole matrix elements zwvz^{\prime}_{wv} and the Coulomb matrix elements g~wnva\tilde{g}^{\prime}_{wnva} and g~wvna\tilde{g}^{\prime}_{wvna} needed for this computation were presented in Eqs. (58) and (65). The resulting value of the amplitude as a function of the number of RPA iteration is shown in Fig. 2 where the oscillatory behavior typical of an RPA calculation is clearly visible. The final value for the 6S1/27S1/26S_{1/2}-7S_{1/2} PNC amplitude is at

EPVRPA=0.89034×1011i|e|a0(QW/N).E^{\rm RPA}_{\rm PV}=0.89034\times 10^{-11}{\rm i}|e|a_{0}(Q_{W}/N)\,. (83)

This value is 0.04% away from the RPA result in Ref. [65]. It is worth noting that the RPA value is only 1% away from the more accurate CCSDvT result [46, 47].

Refer to caption
Figure 2: (Color online) The value of the 6S1/27S1/26S_{1/2}-7S_{1/2} PNC transition amplitude in Cs133{}^{133}\mathrm{Cs} as a function of the number of RPA iteration. Convergence occurs after 20 iterations at the level of fractional accuracy of 10610^{-6}.

VII Parity-mixed coupled-cluster method

In the previous section, we demonstrated the utility of the PM-DHF basis sets in relativistic many-body calculations by computing the Cs133{}^{133}\mathrm{Cs} 6S1/27S1/26S_{1/2}-7S_{1/2} PNC transition amplitude in the all-order RPA method. The RPA result (83) includes all second-order MBPT corrections to matrix elements, but omits important third-order effects including the so-called Brueckner-orbital diagrams whose contributions are numerically as important as the RPA ones [66, 33, 64]. The task of accounting for these higher-order MBPT corrections can be systematically carried out by means of the coupled-cluster (CC) method [67, 68]. For example, it is well-known that a CC formalism which includes singles, doubles and valence triples (CCSDvT) particle-hole excitations from the lowest-order state is complete through the fourth order of MBPT for energies and through the fifth order for matrix elements [69, 70].

The goal of this section is to outline a PM generalization to the PP-CCSDvT method used in Refs. [46, 47], where the conventional PP-DHF basis sets were employed. A labor-intensive numerical implementation of the method discussed here will be the subject of our future work. Since there are multiple implementations of relativistic PP-CC methods, especially in the quantum chemistry community, our theoretical formulation may be useful in the work of other groups as well.

There are several advantages to the PM-CC formulation. First of all, the PP-CC codes are already available, and we outline the strategy for a relatively straightforward generalization of these codes. For example, the CCSDvT method reproduces the relevant atomic properties at a few 0.1% accuracy level, therefore the PM-CCSDvT method (barring implementation errors) should at least be as accurate. Moreover, as mentioned in Sec. I, since the lowest order PM-DHF result is only 3% away from the more accurate CCSDvT value [46, 47], the correlation corrections in the PM approach are substantially smaller than in CCSDvT and, hence, a greater accuracy can be expected. In addition, the PM-CC formulation avoids directly summing over intermediate states in expressions for parity non-conserving amplitudes, as in the original PP-CCSDvT method. This reduces theoretical uncertainties associated with highly-excited and core-excited intermediate states, a subject of controversy [46, 47, 49, 50].

We begin our discussion by going back to the second-quantized form of the full electronic Hamiltonian HH^{\prime}, Eq. (5),

H\displaystyle H^{\prime} =H0+G\displaystyle=H^{\prime}_{0}+G^{\prime}
=iεiN[aiai]+12ijklgijklN[aiajalak],\displaystyle=\sum_{i}\varepsilon^{\prime}_{i}N[a^{\prime\dagger}_{i}a^{\prime}_{i}]+\frac{1}{2}\sum_{ijkl}g^{\prime}_{ijkl}N[a^{\prime\dagger}_{i}a^{\prime\dagger}_{j}a^{\prime}_{l}a^{\prime}_{k}]\,, (84)

where we have dropped the one-particle term ij(VHFU)ijN[aiaj]\sum_{ij}\left(V^{\prime}_{\rm HF}-U^{\prime}\right)_{ij}N[a^{\prime\dagger}_{i}a^{\prime}_{j}] which vanishes due to our choice of the potential U=VHFU^{\prime}=V^{\prime}_{\rm HF}.

In the CC formalism, the exact many-body eigen-state |Ψv|\Psi^{\prime}_{v}\rangle of the Hamiltonian HH^{\prime} can be represented as

|Ψv\displaystyle|\Psi^{\prime}_{v}\rangle =Ω|Ψv(0)=N[exp(K)]|Ψv(0)\displaystyle=\Omega^{\prime}|\Psi_{v}^{(0)}\rangle=N[\exp(K^{\prime})]\,|\Psi^{\prime(0)}_{v}\rangle
=(1+K+12!N[K2]+)|Ψv(0),\displaystyle=\left(1+K^{\prime}+\frac{1}{2!}N[K^{\prime 2}]+\ldots\right)\,|\Psi^{\prime(0)}_{v}\rangle\,, (85)

where Ω\Omega^{\prime} is the wave operator, |Ψv(0)|\Psi^{\prime(0)}_{v}\rangle is again the lowest-order PM-DHF state and the cluster operator KK^{\prime} is expressed in terms of connected diagrams of the wave operator [71]. In the CCSDvT approach, the cluster operator KK^{\prime} is approximated by

K\displaystyle K^{\prime} =n(Kc)n+n(Kv)n\displaystyle=\sum_{n}(K^{\prime}_{c})_{n}+\sum_{n}(K^{\prime}_{v})_{n}
Sc+Dc+Sv+Dv+Tv\displaystyle\approx S^{\prime}_{c}+D^{\prime}_{c}+S^{\prime}_{v}+D^{\prime}_{v}+T^{\prime}_{v}
=maρmaamaa+12!mnabρmnabamanabaa\displaystyle=\sum_{ma}\rho^{\prime}_{ma}a^{\prime\dagger}_{m}a^{\prime}_{a}+\frac{1}{2!}\sum_{mnab}\rho^{\prime}_{mnab}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{b}a^{\prime}_{a}
+mvρmvamav+12!mnaρmnvaamanaaav\displaystyle+\sum_{m\neq v}\rho^{\prime}_{mv}a^{\prime\dagger}_{m}a^{\prime}_{v}+\frac{1}{2!}\sum_{mna}\rho^{\prime}_{mnva}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{a}a^{\prime}_{v}
+13!mnrabρmnrvabamanarabaaav\displaystyle+\frac{1}{3!}\sum_{mnrab}\rho^{\prime}_{mnrvab}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime\dagger}_{r}a^{\prime}_{b}a^{\prime}_{a}a^{\prime}_{v}
=[Uncaptioned image],\displaystyle=\begin{array}[]{l}\includegraphics*[scale={0.4}]{clusterCCSDvT-plus.eps}\end{array}\,, (87)

with the double-headed arrow representing the valence state. Here Sv(Kv)1S^{\prime}_{v}\equiv(K^{\prime}_{v})_{1} and Dv(Kv)2D^{\prime}_{v}\equiv(K^{\prime}_{v})_{2} are the PM valence singles and doubles, Sc(Kc)1S^{\prime}_{c}\equiv(K^{\prime}_{c})_{1} and Dc(Kc)2D^{\prime}_{c}\equiv(K^{\prime}_{c})_{2} are the PM core singles and doubles and Tv(Kv)3T^{\prime}_{v}\equiv(K^{\prime}_{v})_{3} is the PM valence triples. Note that here, they are expressed in terms of the PM creation and annihilation operators aia^{\prime\dagger}_{i} and aia^{\prime}_{i}, in contrast to the conventional CC approach where the cluster operators are expressed in terms of the PP creation and annihilation operators.

The goal of a CC calculation is to compute the cluster amplitudes ρ\rho^{\prime} in Eq. (VII). These amplitudes may be found from the Bloch equation [71] specialized for univalent systems [72]

(εvH0)(Kc)n\displaystyle\left(\varepsilon^{\prime}_{v}-H^{\prime}_{0}\right)(K_{c})_{n} ={QGΩ}connected,n,\displaystyle=\left\{Q^{\prime}G^{\prime}\Omega^{\prime}\right\}_{{\rm connected},n}\,, (88)
(εv+δEvH0)(Kv)n\displaystyle\left(\varepsilon^{\prime}_{v}+\delta E_{v}-H^{\prime}_{0}\right)(K_{v})_{n} ={QGΩ}connected,n,\displaystyle=\left\{Q^{\prime}G^{\prime}\Omega^{\prime}\right\}_{{\rm connected},n}\,,

where the valence correlation energy is given by

δEv=Ψv(0)|GΩ|Ψv(0),\delta E_{v}=\bra{\Psi^{(0)}_{v}}G\Omega\ket{\Psi^{(0)}_{v}}\,, (89)

and Q1|Ψv(0)Ψv(0)|Q^{\prime}\equiv 1-\ket{\Psi^{\prime(0)}_{v}}\bra{\Psi^{\prime(0)}_{v}} is a projection operator onto the space spanned by the PM excited states. The subscript “connected” means that only connected diagrams are retained on the right-hand sides of Eqs. (88).

It is worth stressing that we have used for the energy correction δEv\delta E_{v}, Eq. (89) the formula in the conventional PP-CC scheme. This is justified since the effects of the weak interaction on energies are O(η2)O(\eta^{2}). Although this is intuitively clear, we shall provide a rigorous proof once we have presented the parity decomposition of the PM cluster amplitudes, c.f. Eqs. (104).

Since the commutation and contraction relations among the PM operators aia^{\prime\dagger}_{i} and aia^{\prime}_{i} are identical to those for the PP operators aia^{\dagger}_{i} and aia_{i}, the structure of Eqs. (88) for the PM cluster amplitudes is the same as for the PP amplitudes. [69, 73, 46, 47]. In appendix A, we collect these equations and list them in their explicit form.

Note that in presenting the PM-CC equations, we have used anti-symmetrized combinations for doubles

ρ~mnab\displaystyle\tilde{\rho}^{\prime}_{mnab} ρmnabρmnba=ρmnabρnmab\displaystyle\equiv\rho^{\prime}_{mnab}-\rho^{\prime}_{mnba}=\rho^{\prime}_{mnab}-\rho^{\prime}_{nmab}
=12(ρmnab+ρnmbaρmnbaρnmab),\displaystyle=\frac{1}{2}\left(\rho^{\prime}_{mnab}+\rho^{\prime}_{nmba}-\rho^{\prime}_{mnba}-\rho^{\prime}_{nmab}\right)\,, (90a)
ρ~mnva\displaystyle\tilde{\rho}^{\prime}_{mnva} ρmnvaρnmva,\displaystyle\equiv\rho^{\prime}_{mnva}-\rho^{\prime}_{nmva}\,, (90b)

which have the following symmetry properties

ρ~mnab\displaystyle\tilde{\rho}^{\prime}_{mnab} =ρ~nmab=ρ~mnba,\displaystyle=-\tilde{\rho}^{\prime}_{nmab}=-\tilde{\rho}^{\prime}_{mnba}\,, (91a)
ρ~mnva\displaystyle\tilde{\rho}^{\prime}_{mnva} =ρ~nmva,\displaystyle=-\tilde{\rho}^{\prime}_{nmva}\,, (91b)

and the fully anti-symmetrized valence triples amplitude ρ~mnrvab\tilde{\rho}^{\prime}_{mnrvab}, which is anti-symmetric with respect to any permutation of the indices mnrmnr or abab, e.g.,

ρ~mnrvab\displaystyle\tilde{\rho}^{\prime}_{mnrvab} =ρ~nmrvab=ρ~mrnvab\displaystyle=-\tilde{\rho}^{\prime}_{nmrvab}=-\tilde{\rho}^{\prime}_{mrnvab}
=ρ~mnrvba=ρ~mrnvba=.\displaystyle=-\tilde{\rho}^{\prime}_{mnrvba}=\tilde{\rho}^{\prime}_{mrnvba}=\ldots\,. (92)

The symmetry properties (91) and (VII) are useful for simplifying the CC codes.

Let us now discuss the structure of the PM-CC amplitudes. A general knowledge of this structure will prove useful for the implementation of the PM-CC equations. We begin with the PM single amplitudes ρma\rho^{\prime}_{ma} and ρmv\rho^{\prime}_{mv}.

In the conventional CC approach where a PP single-particle basis is used, the single amplitudes have the angular decomposition

ρma=δκmκaδmmmaS(ma),\displaystyle\rho_{ma}=\delta_{\kappa_{m}\kappa_{a}}\delta_{m_{m}m_{a}}S(ma)\,, (93)
ρmv=δκmκvδmmmvS(mv),\displaystyle\rho_{mv}=\delta_{\kappa_{m}\kappa_{v}}\delta_{m_{m}m_{v}}S(mv)\,,

where S(ma)S(ma) and S(mv)S(mv) are PP reduced single amplitudes. In the current formalism with PM single-particle orbitals, Eqs. (93) may be generalized by appending to ρma\rho_{ma} and ρmv\rho_{mv} PP-odd imaginary components as

ρma=δmmma[δκmκaS(ma)+iηδκm,κaS′′(ma)],\displaystyle\rho^{\prime}_{ma}=\delta_{m_{m}m_{a}}\left[\delta_{\kappa_{m}\kappa_{a}}S(ma)+{\rm i}\eta\delta_{\kappa_{m},-\kappa_{a}}S^{\prime\prime}(ma)\right]\,, (94)
ρmv=δmmmv[δκmκvS(mv)+iηδκm,κvS′′(mv)],\displaystyle\rho^{\prime}_{mv}=\delta_{m_{m}m_{v}}\left[\delta_{\kappa_{m}\kappa_{v}}S(mv)+{\rm i}\eta\delta_{\kappa_{m},-\kappa_{v}}S^{\prime\prime}(mv)\right]\,,

where S′′(ma)S^{\prime\prime}(ma) and S′′(mv)S^{\prime\prime}(mv) are PNC singles amplitudes. The fact that a PM singles amplitude indeed breaks down into two mutually exclusive components, a PP-even real part and a PP-odd imaginary part, is proved in appendix B.

Next, we consider the PM double amplitudes ρmnab\rho^{\prime}_{mnab} and ρmnvb\rho^{\prime}_{mnvb}. As discussed in Sec. V, the PM Coulomb matrix element g~ijkl\tilde{g}^{\prime}_{ijkl} retains its angular structure, Eq. (59), but the reduced matrix element ZL(ijkl)Z^{\prime}_{L}(ijkl) acquires a PP-odd imaginary part. Since the PP double excitation coefficients have the same angular decomposition as the Coulomb matrix elements and the weak interaction conserves total angular momentum, one may decompose the PM double amplitudes as

ρ~mnab\displaystyle\tilde{\rho}^{\prime}_{mnab} =LJL(mnab)S~L(mnab),\displaystyle=\sum_{L}J_{L}(mnab)\tilde{S}^{\prime}_{L}(mnab)\,, (95)
ρ~mnvb\displaystyle\tilde{\rho}^{\prime}_{mnvb} =LJL(mnvb)S~L(mnvb),\displaystyle=\sum_{L}J_{L}(mnvb)\tilde{S}^{\prime}_{L}(mnvb)\,,

where S~L(mnab)\tilde{S}^{\prime}_{L}(mnab) and S~L(mnvb)\tilde{S}^{\prime}_{L}(mnvb) are the reduced double amplitudes.

Similarly to the case of the PM single amplitudes, it may be shown that the the reduced double amplitudes decompose into real and imaginary parts

S~L(mnab)\displaystyle\tilde{S}^{\prime}_{L}(mnab) =S~L(mnab)+iηS~L′′(mnab),\displaystyle=\tilde{S}_{L}(mnab)+{\rm i}\eta\tilde{S}^{\prime\prime}_{L}(mnab)\,, (96)
S~L(mnvb)\displaystyle\tilde{S}^{\prime}_{L}(mnvb) =S~L(mnvb)+iηS~L′′(mnvb).\displaystyle=\tilde{S}_{L}(mnvb)+{\rm i}\eta\tilde{S}^{\prime\prime}_{L}(mnvb)\,.

Here the real part S~L(mnab)\tilde{S}_{L}(mnab) vanishes if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is odd whereas the imaginary part iηS~L′′(mnab){\rm i}\eta\tilde{S}^{\prime\prime}_{L}(mnab) vanishes if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even. The same rules apply for S~L(mnvb)\tilde{S}_{L}(mnvb) and iηS~L′′(mnvb){\rm i}\eta\tilde{S}^{\prime\prime}_{L}(mnvb).

The proof that the reduced double amplitudes indeed separate into mutually exclusive real and imaginary parts with opposite parity selection rules proceeds in a similar manner as for single-excitation coefficients, c.f. appendix B.

Finally, we consider the PM valence triple amplitudes. Again, since the weak interaction does not break the total angular momentum selection rules, ρ~mnrvab\tilde{\rho}^{\prime}_{mnrvab} has the following angular decomposition [69]

ρ~mnrvab=LLh[Uncaptioned image]S~LLh(mnrvab),\tilde{\rho}^{\prime}_{mnrvab}=\sum_{LL^{\prime}h}\begin{array}[]{l}\includegraphics*[scale={0.5}]{triples.pdf}\end{array}\tilde{S}^{\prime}_{LL^{\prime}h}(mnrvab)\,, (97)

where hh is a half integer coupling angular momentum and LL and LL^{\prime} are integer coupling momenta. The formula for writing the algebraic expression corresponding to the angular diagram in Eq. (97) may be found in Ref. [71].

The PM reduced triple amplitude S~LLh(mnrvab)\tilde{S}^{\prime}_{LL^{\prime}h}(mnrvab) does not depend on the magnetic quantum numbers. Similar to the reduced double amplitudes, S~LLh(mnrvab)\tilde{S}^{\prime}_{LL^{\prime}h}(mnrvab) may be decomposed into a PP-even real part and a PP-odd imaginary part, i.e.,

S~LLh(mnrvab)\displaystyle\tilde{S}^{\prime}_{LL^{\prime}h}(mnrvab) =S~LLh(mnrvab)\displaystyle=\tilde{S}_{LL^{\prime}h}(mnrvab)
+iηS~LLh′′(mnrvab),\displaystyle+{\rm i}\eta\tilde{S}^{\prime\prime}_{LL^{\prime}h}(mnrvab)\,, (98)

where S~LLh(mnrvab)\tilde{S}_{LL^{\prime}h}(mnrvab) vanishes if m+n+r+v+a+b\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b} is odd and S~LLh′′(mnrvab)\tilde{S}^{\prime\prime}_{LL^{\prime}h}(mnrvab) vanishes if m+n+r+v+a+b\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b} is even. The proof that valence triple amplitudes separate into mutually exclusive real and imaginary parts of opposite parities proceeds in a similar manner as for single- and double-excitation coefficients, c.f. appendix B.

We have described the angular and parity structure of the PM-CC single, double, and valence triples amplitudes. Let us now turn our attention to the correlation corrections to the energy expressed in the PM basis. Up to the level of valence triples, the valence energy correction may be written as

δEv=δESD+δECC+δEvT,\delta E^{\prime}_{v}=\delta E^{\prime}_{\mathrm{SD}}+\delta E^{\prime}_{\mathrm{CC}}+\delta E^{\prime}_{\mathrm{vT}}\,, (99)

where δESD\delta E^{\prime}_{\mathrm{SD}} represents the linear singles-doubles corrections

δESD\displaystyle\delta E^{\prime}_{\mathrm{SD}} =mag~vavmρma\displaystyle=\sum_{ma}\tilde{g}^{\prime}_{vavm}\rho^{\prime}_{ma}
+12mabg~abvmρ~mvab+12mnbg~vbmnρ~mnvb,\displaystyle+\frac{1}{2}\sum_{mab}\tilde{g}^{\prime}_{abvm}\tilde{\rho}^{\prime}_{mvab}+\frac{1}{2}\sum_{mnb}\tilde{g}^{\prime}_{vbmn}\tilde{\rho}^{\prime}_{mnvb}\,, (100)

while δECC\delta E^{\prime}_{\mathrm{CC}} contains contributions from nonlinear singles-doubles terms

δECC\displaystyle\delta E^{\prime}_{\mathrm{CC}} =abnrg~abnr[ρvbρnrvaρnbρ~vrvaρnvρvrab]\displaystyle=\sum_{abnr}\tilde{g}^{\prime}_{abnr}\left[\rho^{\prime}_{vb}\rho^{\prime}_{nrva}-\rho^{\prime}_{nb}\tilde{\rho}^{\prime}_{vrva}-\rho^{\prime}_{nv}\rho^{\prime}_{vrab}\right]
+anrg~avnrρnaρrv+abng~abnvρvaρnb,\displaystyle+\sum_{anr}\tilde{g}^{\prime}_{avnr}\rho^{\prime}_{na}\rho^{\prime}_{rv}+\sum_{abn}\tilde{g}^{\prime}_{abnv}\rho^{\prime}_{va}\rho^{\prime}_{nb}\,, (101)

and δEvT\delta E^{\prime}_{\mathrm{vT}} is the valence triples term

δEvT\displaystyle\delta E^{\prime}_{\mathrm{vT}} =12abmngabmnρ~vmnvab.\displaystyle=\frac{1}{2}\sum_{abmn}g^{\prime}_{abmn}\tilde{\rho}^{\prime}_{vmnvab}\,. (102)

For completeness, we also present the correlation correction to the core energy δEc\delta E^{\prime}_{c}, although it is not needed in the CC calculations. Since we do not include core triples in our formalism, the correlation correction to the core energy has the form

δEc=12abmngabmnρ~mnab+12abmng~abmnρmaρnb.\delta E^{\prime}_{c}=\frac{1}{2}\sum_{abmn}g^{\prime}_{abmn}\tilde{\rho}_{mnab}+\frac{1}{2}\sum_{abmn}\tilde{g}^{\prime}_{abmn}\rho_{ma}\rho_{nb}\,. (103)

Physically, the energy corrections presented here must be real-valued. Nevertheless, that they are so is not immediately clear from Eqs. (VII), (101), (102) and (103) alone. However, once the decomposition of the PM Coulomb matrix elements and CC amplitudes into real and imaginary parts of opposite parities is taken into account, the reality of the CC energy corrections becomes apparent.

For example, consider the term mag~vavmρma\sum_{ma}\tilde{g}_{vavm}\rho_{ma} in Eq. (VII). Since vv appears twice in g~vavm\tilde{g}_{vavm}, its (nominal) parity is (1)m+a(-1)^{\ell_{m}+\ell_{a}}, which is the same as that of ρma\rho_{ma}. Thus, g~vavm\tilde{g}_{vavm} and ρma\rho_{ma} are either both real or imaginary simultaneously so their product is always real. We may also consider, for example, the term abnrg~abnrρvbρnrva\sum_{abnr}\tilde{g}_{abnr}\rho_{vb}\rho_{nrva} in Eq. (101). Suppose now that a+n+r\ell_{a}+\ell_{n}+\ell_{r} is odd and that b\ell_{b} and v\ell_{v} are even. Then g~abnr\tilde{g}_{abnr} and ρnrva\rho_{nrva} are both imaginary while ρvb\rho_{vb} is real. As a result, the product of these three terms are real and the same argument applies to other cases. The upshot here is that if the total parities of several quantities is even then their product is real. Since the indices of the terms contributing to the correlation energy always appear in pairs, the total parity of each contribution is even and thus they are all real.

Moreover, since each imaginary quantity comes with the small factor η\eta, the PM correlation corrections to the energies of the core EcE_{c} and a valence state EvE_{v} are given by

δEc=δEc+O(η2),\displaystyle\delta E^{\prime}_{c}=\delta E_{c}+O(\eta^{2})\,, (104)
δEv=δEv+O(η2),\displaystyle\delta E^{\prime}_{v}=\delta E_{v}+O(\eta^{2})\,,

where δEc\delta E_{c} and δEv\delta E_{v} are the correlation corrections calculated using PP bases. This fact is in agreement with the general observation that the weak interaction does not produce energy shifts up to O(η2)O(\eta^{2}).

In a conventional PP-CC scheme, the correlation energies are used as a test for the convergence pattern of the CC amplitudes. Analogously, Eqs. (104) show that in a PM-CC calculation, the correlation energies can be used to test the convergence of the real parts of the CC amplitudes. They do not, however, provide information about the convergence of the imaginary parts. We control the convergence patterns of these PP-odd components by directly observing the largest change from iteration to iteration.

The fact that a cluster amplitude’s real and imaginary parts are of opposite parities allows us to formulate a strategy for implementing the PM-CCSDvT code as follows. First, the conventional PP-CC program is executed until it converges. The resulting PP cluster amplitudes from this program are, up to O(η2)O(\eta^{2}), the real components of their PM counterparts. These PP amplitudes are then used as the initial values in a modified PM-CC code. This modified code uses the complex-valued PM matrix elements zijz^{\prime}_{ij} and g~ijkl\tilde{g}^{\prime}_{ijkl} (with parity selection rules modified accordingly) to compute the imaginary PNC part of the PM cluster amplitudes. The convergence of these imaginary parts are checked via their relative changes from iteration to iteration.

Once the PM cluster amplitudes and correlation energies have been found, one may use the obtained wave functions for two valence states Ψw\Psi^{\prime}_{w} and Ψv\Psi^{\prime}_{v} to evaluate various matrix elements, such as that of the electric dipole operator entering PNC amplitude,

Zwv=Ψw|iji|z|jaiaj|ΨvΨw|ΨwΨv|Ψv.Z^{\prime}_{wv}=\frac{\langle\Psi^{\prime}_{w}|\sum_{ij}\langle i^{\prime}|z|j^{\prime}\rangle\,a^{\prime\dagger}_{i}a^{\prime}_{j}|\Psi^{\prime}_{v}\rangle}{\sqrt{\langle\Psi^{\prime}_{w}|\Psi^{\prime}_{w}\rangle\langle\Psi^{\prime}_{v}|\Psi^{\prime}_{v}\rangle}}\,. (105)

The corresponding CCSDvT expressions are given in Ref. [69]. The “dressing” of lines and vertices in expressions for matrix elements is the same as discussed in Ref. [74]. The only difference is that all the PP quantities are to be replaced by PM ones.

VIII Discussion

We have discussed how a conventional coupled-cluster (CC) calculation which uses parity-proper (PP) single-electron basis functions may be generalized to use parity-mixed (PM) basis functions instead. In this PM version of the CC method, the parity non-conserving (PNC) electron-nucleus weak interaction is incorporated into the zeroth-order single-electron Dirac-Hartree-Fock (DHF) Hamiltonian. Such a PM-CC formulation has the advantage over the traditional PP-CC method for several reasons.

Firstly, in a conventional PP-CC calculation of the PNC amplitude where the sum-over-states approach is used, c.f. Eq. (1), contributions to the PNC amplitude are often split into a “main” term, coming from low-lying excited states, and a “tail” term, coming from highly-excited and core-excited intermediate states. A typical breakdown of these contributions [46, 47] is given in Table 1.

It may be observed from Table 1 that although the “main” and “tail” terms in the CCSDvT method have the same absolute uncertainty, the fractional inaccuracy of the former is at the level of 0.2%, whereas that of the latter reaches 10%. In the PM-CC approach, summing over states and thus the artificial separation into “main” and “tail” terms are avoided. As a result, the fractional inaccuracies of all the contributions are anticipated to be at the level of 0.2%0.2\%. Since the “tail” contributes only 2% to the PV amplitude, this means that one of the largest sources of error in Table 1 will be effectively removed. Thereby, with the PM-CCSDvT approach, we anticipate improving the current 0.5% theoretical uncertainty [49] to \sim 0.2%, reaching the new improved accuracy level in the low-energy test of the electroweak sector of the Standard Model.

Coulomb interaction corrections
Main (n=69n=6-9) 0.8823(18)
Tail 0.0175(18)
Total correlated 0.8998(25)
Other corrections
Breit, Ref. [39] -0.0054(5)
QED, Ref. [44] -0.0024(3)
Neutron skin, Ref. [48] -0.0017(5)
eee-e weak interaction, Ref. [32] 0.0003
Final 0.8906(26)
Table 1: Contributions to the parity violating amplitude EPV for the 6S1/27S1/26S_{1/2}\rightarrow 7S_{1/2} transition in 133Cs in units of 1011i|e|a0QW/N10^{-11}i|e|a_{0}Q_{W}/N. Here, N=78N=78 is the number of neutrons in the 133Cs nucleus. The results are from the CCSDvT calculations [46, 47] in the sum-over-state approach. Improving the accuracy of the tail contribution (shown in red) is the goal of the current work.

Secondly, the lowest-order DHF result in this PM approach is only 3% away from the more accurate CCSDvT value. This is to be compared with the traditional DHF result which is off by 18%. This indicates that the correlation corrections in the PM approach are substantially smaller than in the conventional PP method. Depending on the MBPT convergence pattern, one can generically expect an improved theoretical accuracy. In addition, the upgrade of existing and well-tested large-scale PP-CC codes to PM-CC ones is relatively straightforward.

The implementation of a PM-CC code requires a basis of PM single-electron orbitals, which are eigen-states of the PM-DHF Hamiltonian (4). In this paper, we presented several methods through which these eigen-states may be obtained with high accuracy. We note here that the symmetry of Eq. (1) with respect to the exchange DzHWD_{z}\leftrightarrow H_{W} suggests an alternative approach to the APV problem: instead of using the Hamiltonian  (4), one adds an operator λDz\lambda D_{z} (λ\lambda may be thought of as the strength of an external electric field 𝐄=λ𝐳^{\bf E}=\lambda\hat{\bf z}) to the PP atomic Hamiltonian HH and solves for the eigen-states Ψ(λ)\Psi(\lambda) of H+λDzH+\lambda D_{z}. With Ψ(λ)\Psi(\lambda) obtained, one may then proceed to computing the expectation value Ψ(λ)|HW|Ψ(λ)\bra{\Psi(\lambda)}H_{W}\ket{\Psi(\lambda)}, whence the PNC transition amplitude (1) may be calculated by taking the first derivative with respect to λ\lambda. This method has certain advantages such as the relatively simple form of the operator DzD_{z}. However, since DzD_{z} is a tensor operator of rank one, as opposed to HWH_{W} which is a pseudoscalar, it can couple orbitals with different total angular momenta, thus leading to a drastic increase in the number of allowed angular channels in a CC calculation.

With the PM bases obtained, we proceeded to computing the PM matrix elements of inter-electron Coulomb interaction and the electric dipole operator. The former are needed for the computation of correlation corrections to the single-electron wave functions while the latter are needed to calculate the PNC amplitude. We demonstrated the numerical accuracy of our PM approach by using these PM matrix elements in a random-phase approximation (RPA) calculation of the PNC amplitude, obtaining a 0.04% agreement with a previous RPA result [75].

Finally, we presented the extension of the conventional PP-CC method to a PM-CC formalism. We also proved rigorously that a PM cluster amplitude is a complex number which decomposes into mutually exclusive PP-even real part and PP-odd imaginary part. An immediate consequence of this decomposition is that the correlations energies computed from these amplitudes are, reassuringly, real. More importantly, that a PM cluster amplitude is either real or imaginary depending on its nominal parity allows us to formulate a strategy for the PM-CC program.

A full implementation of the PM-CCSDvT calculation based on the strategy mapped out here will be a subject of our future work. The result of this computation will help with the interpretation of the next generation searches for new physics with atomic parity violation (APV) [22, 23]. In addition, since there are multiple implementations of relativistic PP-CC methods, especially in the quantum chemistry community, our theoretical formulation may be useful in the work of other groups.

Acknowledgements

We would like to thank Walter Johnson for valuable discussion. This work was supported in part by the U.S. National Science Foundation grant PHY-1912465, by the Sara Louise Hartman endowed professorship in Physics, and by the Center for Fundamental Physics at Northwestern University.

Appendix A Parity-mixed coupled-cluster equations

In this appendix, we present the equations for the singles, doubles and valence triples CC amplitudes in their explicit form. For ease of presentation, we shall suppress all primes on the quantities involved. It should still be understood, however, that the quantities appearing here are of PM character and are generally complex numbers. For convenient, we will use the notation εijkεi+εj+εk+\varepsilon_{ijk\dots}\equiv\varepsilon_{i}+\varepsilon_{j}+\varepsilon_{k}+\dots to denote sums of single-electron energies [69, 73].

The equation for the core single-excitation coefficients reads

(εaεm)ρma=XSDs+i=13Ais,\displaystyle(\varepsilon_{a}-\varepsilon_{m})\rho_{ma}=X^{s}_{\rm SD}+\sum_{i=1}^{3}A^{s}_{i}\,, (106)

where XSDsX^{s}_{\rm SD} is the linearized singles-doubles term

XSDs\displaystyle X^{s}_{\rm SD} =bng~mbanρnb+12bnrg~mbnrρ~nrab\displaystyle=\sum_{bn}\tilde{g}_{mban}\rho_{nb}+\frac{1}{2}\sum_{bnr}\tilde{g}_{mbnr}\tilde{\rho}_{nrab}
12bcng~bcanρ~mnbc,\displaystyle-\frac{1}{2}\sum_{bcn}\tilde{g}_{bcan}\tilde{\rho}_{mnbc}\,, (107)

while A1,2,3sA^{s}_{1,2,3} are all the nonlinear singles-doubles terms

A1s\displaystyle A^{s}_{1} =drsg~mdrsρraρsdcdsg~cdasρmcρsd,\displaystyle=\sum_{drs}\tilde{g}_{mdrs}\rho_{ra}\rho_{sd}-\sum_{cds}\tilde{g}_{cdas}\rho_{mc}\rho_{sd}\,, (108)
A2s\displaystyle A^{s}_{2} =12cdrsg~cdrsρ~rsdaρmc12cdrsg~cdsrρ~smcdρra\displaystyle=-\frac{1}{2}\sum_{cdrs}\tilde{g}_{cdrs}\tilde{\rho}_{rsda}\rho_{mc}-\frac{1}{2}\sum_{cdrs}\tilde{g}_{cdsr}\tilde{\rho}_{smcd}\rho_{ra}
+cdrsg~cdrsρ~rmcaρsd,\displaystyle+\sum_{cdrs}\tilde{g}_{cdrs}\tilde{\rho}_{rmca}\rho_{sd}\,, (109)
A3s\displaystyle A^{s}_{3} =cdrsg~cdsrρmcρrdρsa.\displaystyle=-\sum_{cdrs}\tilde{g}_{cdsr}\rho_{mc}\rho_{rd}\rho_{sa}\,. (110)

The equation for the core double-excitation coefficients reads

(εabεmn)ρmnab=XSDd+i=16Aid,(\varepsilon_{ab}-\varepsilon_{mn})\rho_{mnab}=X^{d}_{\rm SD}+\sum_{i=1}^{6}A^{d}_{i}\,, (111)

where the linearized singles-doubles term is given by

XSDd\displaystyle X^{d}_{\rm SD} =gmnab+14cdg~cdabρ~mncd+14rsg~mnrsρ~rsab\displaystyle=g_{mnab}+\frac{1}{4}\sum_{cd}\tilde{g}_{cdab}\tilde{\rho}_{mncd}+\frac{1}{4}\sum_{rs}\tilde{g}_{mnrs}\tilde{\rho}_{rsab}
+[rgmnrbρracgcnabρmc\displaystyle+\left[\sum_{r}g_{mnrb}\rho_{ra}-\sum_{c}g_{cnab}\rho_{mc}\right.
+crg~cnrbρ~mrac+(abmn)],\displaystyle\left.+\sum_{cr}\tilde{g}_{cnrb}\tilde{\rho}_{mrac}+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (112)

and the nonlinear singles-doubles terms are given by

A1d\displaystyle A^{d}_{1} =rsgmnrsρraρsb+cdgcdabρmcρnd\displaystyle=\sum_{rs}g_{mnrs}\rho_{ra}\rho_{sb}+\sum_{cd}g_{cdab}\rho_{mc}\rho_{nd}
[drg~mdarρrbρnd+(abmn)],\displaystyle-\left[\sum_{dr}\tilde{g}_{mdar}\rho_{rb}\rho_{nd}+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (113)
A2d\displaystyle A^{d}_{2} =[cdrg~cdrbρndρ~rmcacdrg~cdarρrdρmncb\displaystyle=\left[\sum_{cdr}\tilde{g}_{cdrb}\rho_{nd}\tilde{\rho}_{rmca}-\sum_{cdr}\tilde{g}_{cdar}\rho_{rd}\rho_{mncb}\right.
+cdrgcdraρrbρnmcd+crsg~ncrsρrbρ~smca\displaystyle+\sum_{cdr}g_{cdra}\rho_{rb}\rho_{nmcd}+\sum_{crs}\tilde{g}_{ncrs}\rho_{rb}\tilde{\rho}_{smca}
+crsg~ncrsρscρmrab14crsg~ncrsρmcρ~srab\displaystyle+\sum_{crs}\tilde{g}_{ncrs}\rho_{sc}\rho_{mrab}-\frac{1}{4}\sum_{crs}\tilde{g}_{ncrs}\rho_{mc}\tilde{\rho}_{srab}
+(abmn)],\displaystyle\left.+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (114)
A3d\displaystyle A^{d}_{3} =[cdrgcdarρndρmcρrbcrsgmcrsρncρraρsb\displaystyle=\left[\sum_{cdr}g_{cdar}\rho_{nd}\rho_{mc}\rho_{rb}-\sum_{crs}g_{mcrs}\rho_{nc}\rho_{ra}\rho_{sb}\right.
+(abmn)],\displaystyle\left.+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (115)
A4d=\displaystyle A_{4}^{d}= cdtugcdtuρtuabρmncd+cdtug~cdtuρ~mtacρ~undb\displaystyle\sum_{cdtu}g_{cdtu}\rho_{tuab}\rho_{mncd}+\sum_{cdtu}\tilde{g}_{cdtu}\tilde{\rho}_{mtac}\tilde{\rho}_{undb}
[cdtug~cdtu(ρtubdρmnac+ρmucdρntba)\displaystyle-\left[\sum_{cdtu}\tilde{g}_{cdtu}\left(\rho_{tubd}\rho_{mnac}+\rho_{mucd}\rho_{ntba}\right)\right.
+(abmn)],\displaystyle\left.+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (116)
A5d\displaystyle A^{d}_{5} =cdtugcdtu(ρtaρubρmncd+ρmcρndρtuab)\displaystyle=\sum_{cdtu}g_{cdtu}\left(\rho_{ta}\rho_{ub}\rho_{mncd}+\rho_{mc}\rho_{nd}\rho_{tuab}\right)
[cdtug~cdutρtbρucρmnad+cdtug~cdtuρtcρndρmuab\displaystyle-\left[\sum_{cdtu}\tilde{g}_{cdut}\rho_{tb}\rho_{uc}\rho_{mnad}+\sum_{cdtu}\tilde{g}_{cdtu}\rho_{tc}\rho_{nd}\rho_{muab}\right.
cdtug~cdtuρtbρncρ~muad+(abmn)],\displaystyle\left.\sum_{cdtu}\tilde{g}_{cdtu}\rho_{tb}\rho_{nc}\tilde{\rho}_{muad}+\begin{pmatrix}a\leftrightarrow b\\ m\leftrightarrow n\end{pmatrix}\right]\,, (117)
A6d\displaystyle A^{d}_{6} =cdtugcdtuρtaρubρmcρnd.\displaystyle=\sum_{cdtu}g_{cdtu}\rho_{ta}\rho_{ub}\rho_{mc}\rho_{nd}\,. (118)

The equation for valence singles reads

(εvεm+δEv)ρmv\displaystyle(\varepsilon_{v}-\varepsilon_{m}+\delta E_{v})\rho_{mv} =(XSDs)av\displaystyle=\left(X^{s}_{\rm SD}\right)_{a\rightarrow v}
+i=13(Ais)av+Bs,\displaystyle+\sum_{i=1}^{3}\left(A^{s}_{i}\right)_{a\rightarrow v}+B^{s}\,, (119)

where BsB^{s} stands for the contribution from valence triples

Bs=12abnrgabnrρ~mnrvab.B^{s}=\frac{1}{2}\sum_{abnr}g_{abnr}\tilde{\rho}_{mnrvab}\,. (120)

The equation for valence doubles reads

(εavεmn+δEv)ρmnva\displaystyle(\varepsilon_{av}-\varepsilon_{mn}+\delta E_{v})\rho_{mnva} =(XSDd)av\displaystyle=\left(X^{d}_{\rm SD}\right)_{a\rightarrow v}
+i=15(Aid)av+Bd,\displaystyle+\sum_{i=1}^{5}\left(A^{d}_{i}\right)_{a\rightarrow v}+B^{d}\,, (121)

where BdB^{d} represents the effect of valence triples on valence doubles

Bd\displaystyle B^{d} =12rbc(gbcarρ~mnrvbc+gbcvrρ~nmrabc)\displaystyle=-\frac{1}{2}\sum_{rbc}\left(g_{bcar}\tilde{\rho}_{mnrvbc}+g_{bcvr}\tilde{\rho}_{nmrabc}\right)
+12rsb(gbnrsρ~msrvab+gbmrsρ~snrvab).\displaystyle+\frac{1}{2}\sum_{rsb}\left(g_{bnrs}\tilde{\rho}_{msrvab}+g_{bmrs}\tilde{\rho}_{snrvab}\right)\,. (122)

The equation for valence triples reads

(εabvεmnr+δEv)ρ~mnrvab=B1t+B2t,(\varepsilon_{abv}-\varepsilon_{mnr}+\delta E_{v})\tilde{\rho}_{mnrvab}=B^{t}_{1}+B^{t}_{2}\,, (123)

where

B1t\displaystyle B^{t}_{1} =+s(g~nrsvρ~msab+g~rmsvρ~nsab+g~mnsvρ~rsab)\displaystyle=+\sum_{s}\left(\tilde{g}_{nrsv}\tilde{\rho}_{msab}+\tilde{g}_{rmsv}\tilde{\rho}_{nsab}+\tilde{g}_{mnsv}\tilde{\rho}_{rsab}\right)
c(g~mcvaρ~nrcbg~mcvbρ~nrca+g~ncvaρ~rmcb\displaystyle-\sum_{c}\left(\tilde{g}_{mcva}\tilde{\rho}_{nrcb}-\tilde{g}_{mcvb}\tilde{\rho}_{nrca}+\tilde{g}_{ncva}\tilde{\rho}_{rmcb}\right.
g~ncvbρ~rmca+g~rcvaρ~mncbg~rcvbρ~mnca),\displaystyle\left.-\tilde{g}_{ncvb}\tilde{\rho}_{rmca}+\tilde{g}_{rcva}\tilde{\rho}_{mncb}-\tilde{g}_{rcvb}\tilde{\rho}_{mnca}\right)\,, (124)
B2t\displaystyle B^{t}_{2} =c(g~mcabρ~nrvc+g~ncabρ~rmvc+g~rcabρ~mnvc)\displaystyle=\sum_{c}\left(\tilde{g}_{mcab}\tilde{\rho}_{nrvc}+\tilde{g}_{ncab}\tilde{\rho}_{rmvc}+\tilde{g}_{rcab}\tilde{\rho}_{mnvc}\right)
+s(g~nrsbρ~msvag~nrsaρ~msvb+g~rmsbρ~nsva\displaystyle+\sum_{s}\left(\tilde{g}_{nrsb}\tilde{\rho}_{msva}-\tilde{g}_{nrsa}\tilde{\rho}_{msvb}+\tilde{g}_{rmsb}\tilde{\rho}_{nsva}\right.
g~rmsaρ~nsvb+g~mnsbρ~rsvag~mnsaρ~rsvb).\displaystyle\left.-\tilde{g}_{rmsa}\tilde{\rho}_{nsvb}+\tilde{g}_{mnsb}\tilde{\rho}_{rsva}-\tilde{g}_{mnsa}\tilde{\rho}_{rsvb}\right)\,. (125)

The formulas for the valence energy correction δEv\delta E_{v} in these equations were given in Eqs. (VII), (101), and (102) in the main text.

Appendix B Parity decomposition of coupled-cluster amplitudes

In this appendix, we prove that PM cluster amplitudes decompose into real parts with even parities and imaginary parts with odd parities. More specifically, we show that (i) the singles amplitude ρma\rho^{\prime}_{ma} (ρmv\rho^{\prime}_{mv}) is purely real if the sum m+a\ell_{m}+\ell_{a} (m+v\ell_{m}+\ell_{v}) is even but is purely imaginary if the sum is odd, (ii) the doubles amplitude ρmnab\rho^{\prime}_{mnab} (ρmnvb\rho^{\prime}_{mnvb}) is purely real if the sum m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} (m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}) is even but is purely imaginary if the sum is odd, and (iii) the triples amplitude ρmnrvab\rho^{\prime}_{mnrvab} is purely real if the sum m+n+r+v+a+b\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b} is even but is purely imaginary if the sum is odd. Note that although we limit the current discussion to valence triples, the proof here applies to cluster amplitudes of all ranks.

There are several ways through which the selections rules imposed on the PM cluster amplitudes may be demonstrated. Here, present two such methods, namely, a proof by induction on the cluster equations (see appendix A) and a proof which uses the parity operator (see below).

B.1 Proof using the cluster equations

B.1.1 Parity-proper cluster amplitudes

We begin by deriving the parity selection rule imposed on the PP amplitudes. This serves as a starting point which motivates and generalizes well to the case of PM amplitudes. For definiteness, we concentrate on the PP core single and double amplitudes and prove by induction that

ρma\displaystyle\rho_{ma} mod2(m+a+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{a}+1)\,, (126a)
ρmnab\displaystyle\rho_{mnab} mod2(m+n+a+b+1).\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}+1)\,. (126b)

From the definition (90a), one observes that ρ~mnab\tilde{\rho}_{mnab} has the same selection rule as ρmnab\rho_{mnab}, Eq. (126b). The selection rules for PP valence single, double, and triple amplitudes follow from those for core singles and doubles in a trivial way.

To prove the selection rules (126) by induction, we consider solving the PP version of Eqs. (106) and (111) iteratively. As initial values, we take ρma(0)=0\rho^{(0)}_{ma}=0 and ρmnab(0)=0\rho^{(0)}_{mnab}=0. Equations (106) and (111) then give, after the first iteration

ρma(1)\displaystyle\rho^{(1)}_{ma} =0,\displaystyle=0\,, (127)
ρmnab(1)\displaystyle\rho^{(1)}_{mnab} =gmnab.\displaystyle=g_{mnab}\,.

Equation (126a) is satisfied trivially while Eq. (126b) is satisfied due to the selection rules on the Coulomb matrix elements gmnabg_{mnab}.

We now assume that Eqs. (126) are satisfied by ρma(n)\rho^{(n)}_{ma} and ρmnab(n)\rho^{(n)}_{mnab} for n2n\geq 2. Let us investigate the driving terms (XSDs)(n)(X_{\rm SD}^{s})^{(n)} and (Ais)(n)(A_{i}^{s})^{(n)} on the right-hand side of the single equation, Eq. (106). A close inspection of these terms shows that they vanish if a+m\ell_{a}+\ell_{m} is odd. For example, the term g~mban(n)ρnb(n)\tilde{g}^{(n)}_{mban}\rho^{(n)}_{nb} in Eq. (A) is nonzero only if both m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} and n+b\ell_{n}+\ell_{b} are even, which implies that a+m\ell_{a}+\ell_{m} is even. As a result, Eq. (106) implies that

ρma(n+1)=(XSDs)(n)+i=13(Ais)(n)ϵaϵm\rho_{ma}^{(n+1)}=\frac{(X_{\rm SD}^{s})^{(n)}+\sum_{i=1}^{3}(A_{i}^{s})^{(n)}}{\epsilon_{a}-\epsilon_{m}} (128)

is zero if a+m\ell_{a}+\ell_{m} is odd.

Analogously, one may show that the driving terms (XSDd)(n)(X_{\rm SD}^{d})^{(n)} and (Aid)(n)(A_{i}^{d})^{(n)} on the right-hand side of the double equation, Eq. (111) vanish if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is odd. For example, the term g~cdab(n)ρ~mncd(n)\tilde{g}^{(n)}_{cdab}\tilde{\rho}^{(n)}_{mncd} in Eq. (A) vanishes unless c+d+a+b\ell_{c}+\ell_{d}+\ell_{a}+\ell_{b} and m+n+c+d\ell_{m}+\ell_{n}+\ell_{c}+\ell_{d} are both even. That these sums are both even is equivalent to requiring m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} to be even. Similar arguments apply to all other terms in Eqs. (A)–(118). As a result, Eq. (111) implies that

ρmnab(n+1)=(XSDd)(n)+i=16(Aid)(n)ϵabϵmn\rho_{mnab}^{(n+1)}=\frac{(X_{\rm SD}^{d})^{(n)}+\sum_{i=1}^{6}(A_{i}^{d})^{(n)}}{\epsilon_{ab}-\epsilon_{mn}} (129)

is zero if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is odd. Due to the definition (90a), this selection rule for ρmnab(n+1)\rho_{mnab}^{(n+1)} also applies to ρ~mnab(n+1)\tilde{\rho}_{mnab}^{(n+1)}.

By the principle of induction, we conclude that the selection rules (126) are satisfied by all core single and double amplitudes. An argument along the same line shows that the valence singles, doubles and triples satisfy similar selection rules

ρmv\displaystyle\rho_{mv} mod2(m+v+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{v}+1)\,, (130)
ρmnva\displaystyle\rho_{mnva} mod2(m+n+v+a+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{v}+\ell_{a}+1)\,,
ρmnrvab\displaystyle\rho_{mnrvab} mod2(m+n+r+v+a+b+1).\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}+1)\,.

B.1.2 Parity-mixed cluster amplitudes

We have proved by induction the selection rules imposed on the PP cluster amplitudes from the PP cluster equations. Here, we generalize this method to show that the PM cluster amplitudes may be decomposed into real and imaginary parts with opposite parities. Again, we concentrate on the PM core singles and doubles, proving that

Re(ρma)\displaystyle{\rm Re}(\rho^{\prime}_{ma}) mod2(m+a+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{a}+1)\,, (131a)
Im(ρma)\displaystyle{\rm Im}(\rho^{\prime}_{ma}) mod2(m+a),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{a})\,, (131b)
Re(ρmnab)\displaystyle{\rm Re}(\rho_{mnab}) mod2(m+n+a+b+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}+1)\,, (131c)
Im(ρmnab)\displaystyle{\rm Im}(\rho_{mnab}) mod2(m+n+a+b).\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b})\,. (131d)

Again, the selection rules for g~mnab\tilde{g}^{\prime}_{mnab} are the same as those for gmnabg^{\prime}_{mnab} and the selection rules for PM valence single, double, and triple amplitudes follow directly from those for PM core singles and doubles.

Similarly to the PP case, we consider solving the PM version of Eqs. (106) and (111) iteratively. As before, we take ρma(0)=0\rho^{\prime(0)}_{ma}=0 and ρmnab(0)=0\rho^{\prime(0)}_{mnab}=0. Equations (106) and (111) then give, after the first iteration

ρma(1)\displaystyle\rho^{\prime(1)}_{ma} =0,\displaystyle=0\,, (132)
ρmnab(1)\displaystyle\rho^{\prime(1)}_{mnab} =gmnab.\displaystyle=g^{\prime}_{mnab}\,.

Equation (131a) and (131b) are satisfied trivially while Eqs. (131c) and (131d) are satisfied due to the selection rules on the PM Coulomb matrix elements gmnabg^{\prime}_{mnab} (see Sec. V).

We now assume that Eqs. (131) are satisfied for ρma(n)\rho^{\prime(n)}_{ma} and ρmnab(n)\rho^{\prime(n)}_{mnab} with n2n\geq 2. It is then straightforward to show that the driving terms (XSDs)(n)(X^{\prime s}_{\rm SD})^{(n)} and (Ais)(n)(A^{\prime s}_{i})^{(n)} (we have used the prime to emphasize that these quantities are of PM character) on the right-hand side of the single equation, Eq. (106), are real if a+m\ell_{a}+\ell_{m} is even and are purely imaginary if a+m\ell_{a}+\ell_{m} is odd. For example, consider again the term g~mban(n)ρnb(n)\tilde{g}^{\prime(n)}_{mban}\rho^{\prime(n)}_{nb} in Eq. (A). By the induction assumptions, g~mban(n)\tilde{g}^{\prime(n)}_{mban} is real if m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} is even and is imaginary if m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} is odd. Similarly, ρnb(n)\rho^{\prime(n)}_{nb} is real if m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} is even and is imaginary if m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} is odd. As a result, the product g~mban(n)ρnb(n)\tilde{g}^{\prime(n)}_{mban}\rho^{\prime(n)}_{nb} is real if m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} and n+b\ell_{n}+\ell_{b} have the same parity which can only be satisfied if m+a\ell_{m}+\ell_{a} is even. On the other hand, if m+a\ell_{m}+\ell_{a} is odd then m+b+a+n\ell_{m}+\ell_{b}+\ell_{a}+\ell_{n} and n+b\ell_{n}+\ell_{b} have opposite parities and g~mban(n)ρnb(n)\tilde{g}^{\prime(n)}_{mban}\rho^{\prime(n)}_{nb} is imaginary.

As a result, Eq. (106) implies that

ρma(n+1)=(XSDs)(n)+i=13(Ais)(n)ϵaϵm\rho^{\prime(n+1)}_{ma}=\frac{(X^{\prime s}_{\rm SD})^{(n)}+\sum_{i=1}^{3}(A^{\prime s}_{i})^{(n)}}{\epsilon_{a}-\epsilon_{m}} (133)

is purely real if a+m\ell_{a}+\ell_{m} is even and purely imaginary if a+m\ell_{a}+\ell_{m} is odd.

Analogously, it may be shown that the driving terms (XSDd)(n)(X^{\prime d}_{\rm SD})^{(n)} and (Aid)(n)(A^{\prime d}_{i})^{(n)} on the right-hand side of the double equation, Eq. (111), are real if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even and purely imaginary if this sum is odd. Consider again, for example, the term g~cdab(n)ρ~mncd(n)\tilde{g}^{\prime(n)}_{cdab}\tilde{\rho}^{\prime(n)}_{mncd} in Eq. (A). Whether this product is purely real or imaginary depends on the parities of its factors. This dependence is shown explicitly in Table 2. We observe from this table that g~cdab(n)ρ~mncd(n)\tilde{g}^{\prime(n)}_{cdab}\tilde{\rho}^{\prime(n)}_{mncd} is purely real if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even and purely imaginary if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even is odd. Similar arguments apply to all other terms in Eqs. (A)–(118). As a result, one finds from Eq. (111) that

ρmnab(n+1)=(XSDd)(n)+i=16(Aid)(n)ϵabϵmn\rho^{\prime(n+1)}_{mnab}=\frac{(X^{\prime d}_{\rm SD})^{(n)}+\sum_{i=1}^{6}(A^{\prime d}_{i})^{(n)}}{\epsilon_{ab}-\epsilon_{mn}} (134)

is purely real if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even and purely imaginary if m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b} is even is odd. Due to the definition (90a), the same selection rules hold for ρ~mnab(n+1)\tilde{\rho}^{\prime(n+1)}_{mnab}.

By the principle of induction, we conclude that the selection rules (131) are satisfied by all PM core single and double amplitudes. An argument along the same line shows that the PM valence singles, doubles and triples satisfy similar conditions, i.e.,

Re(ρmv)\displaystyle{\rm Re}(\rho_{mv}) mod2(m+v+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{v}+1)\,, (135)
Im(ρmv)\displaystyle{\rm Im}(\rho_{mv}) mod2(m+v),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{v})\,,
Re(ρmnva)\displaystyle{\rm Re}(\rho_{mnva}) mod2(m+n+v+a+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{v}+\ell_{a}+1)\,,
Im(ρmnva)\displaystyle{\rm Im}(\rho_{mnva}) mod2(m+n+v+a),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{v}+\ell_{a})\,,
Re(ρmnrvab)\displaystyle{\rm Re}(\rho_{mnrvab}) mod2(m+n+r+v+a+b+1),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}+1)\,,
Im(ρmnrvab)\displaystyle{\rm Im}(\rho_{mnrvab}) mod2(m+n+r+v+a+b),\displaystyle\propto{\rm mod}_{2}(\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b})\,,
a+b\ell_{a}+\ell_{b} c+d\ell_{c}+\ell_{d} m+n\ell_{m}+\ell_{n} g~cdab(n)\tilde{g}^{\prime(n)}_{cdab} ρ~mncd(n)\tilde{\rho}^{\prime(n)}_{mncd} m+n\ell_{m}+\ell_{n} g~cdab(n)\tilde{g}^{\prime(n)}_{cdab}
+a+b+\ell_{a}+\ell_{b} ×ρ~mncd(n)\times\tilde{\rho}^{\prime(n)}_{mncd}
e e e r r e r
e e o r i o i
e o e i i e r
e o o i r o i
o e e i r o i
o e o i i e r
o o e r i o i
o o o r r e r
Table 2: The dependence of the reality of the term g~cdab(n)ρ~mncd(n)\tilde{g}^{\prime(n)}_{cdab}\tilde{\rho}^{\prime(n)}_{mncd} in Eq. (A) on the parity of the sum m+n+a+b\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}. Here, “e” stands for even, “o” for odd, “r” for real, and “i” for imaginary.

B.2 Proof using the parity operator

In the last subsection, we have proved that the PM cluster amplitudes decompose into real and imaginary parts of opposite parities by using induction. Here, we provide a different proof using the parity operator in second-quantized form (see Sec. III)

Π\displaystyle\Pi =μ=0Ne1(μ!)2{a}{m}(1)i=1μai+mi\displaystyle=\sum_{\mu=0}^{N_{e}}\frac{1}{(\mu!)^{2}}\sum_{\{a\}\{m\}}(-1)^{\sum_{i=1}^{\mu}\ell_{a_{i}}+\ell_{m_{i}}}
×am1amμaa1aaμ|Ψv(0)Ψv(0)|\displaystyle\times a^{\dagger}_{m_{1}}\ldots a^{\dagger}_{m_{\mu}}a_{a_{1}}\ldots a_{a_{\mu}}\ket{\Psi^{(0)}_{v}}\bra{\Psi^{(0)}_{v}}
×aaμaa1amμam1.\displaystyle\times a^{\dagger}_{a_{\mu}}\ldots a^{\dagger}_{a_{1}}a_{m_{\mu}}\ldots a_{m_{1}}\,. (136)

B.2.1 Parity-proper cluster amplitudes

We again begin our consideration by deriving the parity selection rule imposed on the PP amplitudes. We provide here a formal treatment which motivates and generalizes well to the case of PM amplitudes. In Sec. III, we presented the second-quantized form of the parity operator Π\Pi. Using Eq. (B.2), it may be checked that the lowest order state Ψv(0)\Psi^{(0)}_{v}, comprising of one valence electron above a closed-shell core, satisfies

Π|Ψv(0)=(1)v|Ψv(0),\Pi\ket{\Psi^{(0)}_{v}}=(-1)^{\ell_{v}}\ket{\Psi^{(0)}_{v}}\,, (137)

where we have again used the fact that the closed-shell core has even parity.

Since the inter-electron Coulomb interaction is PP-even, we require that the correlation corrections to the zeroth-order wave function also satisfy Eq. (137), i.e.,

Π|Ψv=(1)v|Ψv,\Pi\ket{\Psi_{v}}=(-1)^{\ell_{v}}\ket{\Psi_{v}}\,, (138)

where |Ψv=Ω|Ψv(0)\ket{\Psi_{v}}=\Omega\ket{\Psi^{(0)}_{v}} is the PP equivalent of the wave function defined in Eq. (VII).

Expanding the wave operator Ω\Omega into singles, doubles, triples, etc., we may write the wave function Ψv\Psi_{v} as

|Ψv\displaystyle\ket{\Psi_{v}} =(1+maρmaamaa+12!mnabρmnabamanabaa\displaystyle=\left(1+\sum_{ma}\rho_{ma}a^{\dagger}_{m}a_{a}+\frac{1}{2!}\sum_{mnab}\rho_{mnab}a^{\dagger}_{m}a^{\dagger}_{n}a_{b}a_{a}\right.
+mvρmvamav+12!mnaρmnvaamanaaav\displaystyle+\sum_{m\neq v}\rho_{mv}a^{\dagger}_{m}a_{v}+\frac{1}{2!}\sum_{mna}\rho_{mnva}a^{\dagger}_{m}a^{\dagger}_{n}a_{a}a_{v}
+13!mnrabρmnrvabamanarabaaav+)|Ψv(0),\displaystyle\left.+\frac{1}{3!}\sum_{mnrab}\rho_{mnrvab}a^{\dagger}_{m}a^{\dagger}_{n}a^{\dagger}_{r}a_{b}a_{a}a_{v}+\ldots\right)\ket{\Psi^{(0)}_{v}}\,, (139)

which gives

Π|Ψv\displaystyle\Pi\ket{\Psi_{v}} =(1)v[1+ma(1)m+aρmaamaa\displaystyle=(-1)^{\ell_{v}}\Bigg{[}1+\sum_{ma}(-1)^{\ell_{m}+\ell_{a}}\rho_{ma}a^{\dagger}_{m}a_{a}
+mv(1)m+vρmvamav\displaystyle+\sum_{m\neq v}(-1)^{\ell_{m}+\ell_{v}}\rho_{mv}a^{\dagger}_{m}a_{v}
+12!mnab(1)m+n+a+bρmnabamanabaa\displaystyle+\frac{1}{2!}\sum_{mnab}(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}}\rho_{mnab}a^{\dagger}_{m}a^{\dagger}_{n}a_{b}a_{a}
+12!mna(1)m+n+a+vρmnvaamanaaav\displaystyle+\frac{1}{2!}\sum_{mna}(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{v}}\rho_{mnva}a^{\dagger}_{m}a^{\dagger}_{n}a_{a}a_{v}
+13!mnrab(1)m+n+r+v+a+bρmnrvab\displaystyle+\frac{1}{3!}\sum_{mnrab}(-1)^{\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}}\rho_{mnrvab}
×amanarabaaav+]|Ψv(0).\displaystyle\times a^{\dagger}_{m}a^{\dagger}_{n}a^{\dagger}_{r}a_{b}a_{a}a_{v}+\ldots\Bigg{]}\ket{\Psi^{(0)}_{v}}\,. (140)

Substituting Eqs. (B.2.1) and (B.2.1) into Eq. (137) and comparing like terms, we obtain the following parity selection rules for the PP cluster amplitudes

(1)m+aρma\displaystyle(-1)^{\ell_{m}+\ell_{a}}\rho_{ma} =ρma,\displaystyle=\rho_{ma}\,, (141)
(1)m+vρmv\displaystyle(-1)^{\ell_{m}+\ell_{v}}\rho_{mv} =ρmv,\displaystyle=\rho_{mv}\,,
(1)m+n+a+bρmnab\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}}\rho_{mnab} =ρmnab,\displaystyle=\rho_{mnab}\,,
(1)m+n+v+aρmnva\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{v}+\ell_{a}}\rho_{mnva} =ρmnva,\displaystyle=\rho_{mnva}\,,
(1)m+n+r+v+a+bρmnrvab\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}}\rho_{mnrvab} =ρmnrvab.\displaystyle=\rho_{mnrvab}\,.

The first of Eqs. (141) implies that if m+a\ell_{m}+\ell_{a} is odd then ρma=0\rho_{ma}=0. Similar parity selection rules follow from the rest of Eqs. (141).

B.2.2 Parity-mixed cluster amplitudes

We have shown that the use of the parity operator Π\Pi allows us to formally prove the parity selection rules imposed on the PP amplitudes. To extend this formalism to the case of PM amplitudes, we first consider the effect of Π\Pi on a PM single-electron orbital ψi\psi^{\prime}_{i}. Since the PM orbital ψi\psi^{\prime}_{i} splits into two components with opposite parities, Eq. (22), we have

Π|ψi\displaystyle\Pi\ket{\psi^{\prime}_{i}} =(1)i|ψi+(1)i¯iη|ψ¯i\displaystyle=(-1)^{\ell_{i}}\ket{\psi_{i}}+(-1)^{\ell_{\bar{i}}}{\rm i}\eta\ket{\bar{\psi}_{i}}
=(1)i(|ψiiη|ψ¯i)\displaystyle=(-1)^{\ell_{i}}\left(\ket{\psi_{i}}-{\rm i}\eta\ket{\bar{\psi}_{i}}\right)
=(1)iPη|ψi,\displaystyle=(-1)^{\ell_{i}}P_{\eta}\ket{\psi^{\prime}_{i}}\,, (142)

where we have introduced the operator PηP_{\eta} which changes η\eta to η-\eta. We see that the action of Π\Pi on ψi\psi^{\prime}_{i} is equivalent to multiplying ψi\psi^{\prime}_{i} with its nominal parity (1)i(-1)^{\ell_{i}} and applying the operator PηP_{\eta}. Note that since changing the sign of η\eta does not affect terms O(η2)O(\eta^{2}), Eq. (B.2.2) is correct up to O(η3)O(\eta^{3}).

To find the PM equivalence of Eq. (137), we first expand the PM creation operator aia^{\prime\dagger}_{i} in terms of its PP counterparts

ai=ai+iηj¯γij¯aj¯,a^{\prime\dagger}_{i}=a^{\dagger}_{i}+{\rm i}\eta\sum_{\bar{j}}\gamma_{i\bar{j}}a^{\dagger}_{\bar{j}}\,, (143)

where we have used Eqs. (36) and (47) (see, for example, Ref. [76] for a detailed discussion on the effects of basis rotation on the second-quantization operators). As a result, up to O(η)O(\eta), we have

|Ψv(0)\displaystyle\ket{\Psi^{\prime(0)}_{v}} =avaa1aan1|0=avaa1aan1|0\displaystyle=a^{\prime\dagger}_{v}a^{\prime\dagger}_{a_{1}}\dots a^{\prime\dagger}_{a_{n-1}}\ket{0}=a^{\dagger}_{v}a^{\dagger}_{a_{1}}\dots a^{\dagger}_{a_{n-1}}\ket{0}
+iη(j¯γvj¯aj¯)aa1aan1|0\displaystyle+{\rm i}\eta\left(\sum_{\bar{j}}\gamma_{v\bar{j}}a^{\dagger}_{\bar{j}}\right)a^{\dagger}_{a_{1}}\dots a^{\dagger}_{a_{n-1}}\ket{0}
+iηav(j¯γa1j¯aj¯)aan1|0\displaystyle+{\rm i}\eta a^{\dagger}_{v}\left(\sum_{\bar{j}}\gamma_{a_{1}\bar{j}}a^{\dagger}_{\bar{j}}\right)\dots a^{\dagger}_{a_{n-1}}\ket{0}
++iηavaa1(j¯γan1j¯aj¯)|0,\displaystyle+\dots+{\rm i}\eta a^{\dagger}_{v}a^{\dagger}_{a_{1}}\dots\left(\sum_{\bar{j}}\gamma_{a_{n-1}\bar{j}}a^{\dagger}_{\bar{j}}\right)\ket{0}\,, (144)

which makes it clear that

Π|Ψv(0)=(1)vPη|Ψv(0).\Pi\ket{\Psi^{\prime(0)}_{v}}=(-1)^{\ell_{v}}P_{\eta}\ket{\Psi^{\prime(0)}_{v}}\,. (145)

Again, we have used the fact that the nominal parity of a closed-shell core is even to remove the factor (1)aa(-1)^{\sum_{a}\ell_{a}} in Eq. (145). Again, we see that the action of Π\Pi on Ψv(0)\Psi^{\prime(0)}_{v} is equivalent to multiplying Ψv(0)\Psi^{\prime(0)}_{v} with its nominal parity factor and changing ηη\eta\rightarrow-\eta, where the nominal parity of the many-electron state Ψv(0)\Psi^{\prime(0)}_{v} is defined as that of its valence orbital (the closed-shell core has even nominal parity).

We now consider the effect of correlations on the nominal parity of the zeroth-order state. Although the operator VcV^{\prime}_{c} in Eq. (4), which comprises of the inter-electron Coulomb interaction and the PNC-DHF potential, is not PP-even, it is predominantly so. In other words, we may write

Vc=VcPC+iηVcPNCV^{\prime}_{c}=V^{\rm PC}_{c}+{\rm i}\eta V_{c}^{\rm PNC} (146)

where VcPCV_{c}^{\rm PC} and VcPNCV_{c}^{\rm PNC} are respectively its parity conserving and parity non-conserving parts. As a result, the perturbation VcV^{\prime}_{c} preserves, up to O(η)O(\eta), the nominal parity of the zeroth-order state.

This may be demonstrated more rigorously if we assume, without loss of generality, that Ψv(0)\Psi^{\prime(0)}_{v} is predominantly even and write, for brevity, Ψv(0)=e(0)+iηo(0)\Psi^{\prime(0)}_{v}={\rm e}^{(0)}+{\rm i}\eta{\rm o}^{(0)}, where the letter “e” denotes a PP-even wave function and the letter “o” denotes a PP-odd wave function. To the first order in VcV^{\prime}_{c}, the correlation correction to Ψv(0)\Psi^{\prime(0)}_{v} has the structure

|δΨv\displaystyle\ket{\delta\Psi^{\prime}_{v}} eiηo|Vc|e(0)+iηo(0)|eiηo\displaystyle\sim\bra{{\rm e}-{\rm i}\eta{\rm o}}V^{\prime}_{c}\ket{{\rm e}^{(0)}+{\rm i}\eta{\rm o}^{(0)}}\ket{{\rm e}-{\rm i}\eta{\rm o}}
+oiηe|Vc|e(0)+iηo(0)|oiηe,\displaystyle+\bra{{\rm o}-{\rm i}\eta{\rm e}}V^{\prime}_{c}\ket{{\rm e}^{(0)}+{\rm i}\eta{\rm o}^{(0)}}\ket{{\rm o}-{\rm i}\eta{\rm e}}\,, (147)

where, for brevity, we have dropped the energy denominator and the summation over intermediate states. Expanding Eq. (B.2.2) and keeping only terms up to O(η)O(\eta), we have

|δΨv\displaystyle\ket{\delta\Psi^{\prime}_{v}} e|VcPC|e(0)|e+iηo|VcPNC|e(0)|o\displaystyle\sim\bra{{\rm e}}V_{c}^{\rm PC}\ket{{\rm e}^{(0)}}\ket{\rm e}+{\rm i}\eta\bra{{\rm o}}V_{c}^{\rm PNC}\ket{{\rm e}^{(0)}}\ket{\rm o}
+iηe|VcPNC|o(0)|o,\displaystyle+{\rm i}\eta\bra{{\rm e}}V_{c}^{\rm PNC}\ket{{\rm o}^{(0)}}\ket{\rm o}\,, (148)

which clearly shows that, parity-wise, δΨv\delta\Psi^{\prime}_{v} and thus Ψv\Psi^{\prime}_{v} have the same structure as Ψv(0)\Psi^{\prime(0)}_{v}. More explicitly, we require that

Π|Ψv=(1)vPη|Ψv.\Pi\ket{\Psi^{\prime}_{v}}=(-1)^{\ell_{v}}P_{\eta}\ket{\Psi^{\prime}_{v}}\,. (149)

We may now use Eq. (149) to derive selection rules on the PM cluster amplitudes similar for those in Eqs. (141). For this purpose, we first note the effect of PηP_{\eta} on the ρ\rho^{\prime}. Since the the weak interaction introduces imaginary components to the cluster amplitudes, namely, ρma=ρma+iηρma′′\rho^{\prime}_{ma}=\rho_{ma}+i\eta\rho^{\prime\prime}_{ma} and so on, we see that changing ηη\eta\rightarrow-\eta is the same as taking the complex conjugates of these amplitudes, i.e.,

Pηρma\displaystyle P_{\eta}\rho^{\prime}_{ma} =(ρma),\displaystyle=(\rho^{\prime}_{ma})^{*}\,, (150)
Pηρmv\displaystyle P_{\eta}\rho^{\prime}_{mv} =(ρmv),\displaystyle=(\rho^{\prime}_{mv})^{*}\,,
Pηρmnab\displaystyle P_{\eta}\rho^{\prime}_{mnab} =(ρmnab),\displaystyle=(\rho^{\prime}_{mnab})^{*}\,,
Pηρmnva\displaystyle P_{\eta}\rho^{\prime}_{mnva} =(ρmnva),\displaystyle=(\rho^{\prime}_{mnva})^{*}\,,
Pηρmnrvab\displaystyle P_{\eta}\rho^{\prime}_{mnrvab} =(ρmnrvab).\displaystyle=(\rho^{\prime}_{mnrvab})^{*}\,.

Using Eq. (150), we may now write Eq. (149) in a more explicit form. Remembering that |Ψv=Ω|Ψv(0)\ket{\Psi^{\prime}_{v}}=\Omega^{\prime}\ket{\Psi^{\prime(0)}_{v}}, we may expand the wave operator Ω\Omega^{\prime} into singles, doubles, and triples, obtaining

Π|Ψv\displaystyle\Pi\ket{\Psi^{\prime}_{v}} =(1)v[1+ma(1)m+aρmaPηamaa\displaystyle=(-1)^{\ell_{v}}\left[1+\sum_{ma}(-1)^{\ell_{m}+\ell_{a}}\rho^{\prime}_{ma}P_{\eta}a^{\prime\dagger}_{m}a^{\prime}_{a}\right.
+mv(1)m+vρmvPηamav\displaystyle\left.+\sum_{m\neq v}(-1)^{\ell_{m}+\ell_{v}}\rho^{\prime}_{mv}P_{\eta}a^{\prime\dagger}_{m}a^{\prime}_{v}\right.
+12!mnab(1)m+n+a+bρmnabPηamanabaa\displaystyle\left.+\frac{1}{2!}\sum_{mnab}(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}}\rho^{\prime}_{mnab}P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{b}a^{\prime}_{a}\right.
+12!mna(1)m+n+a+vρmnvaPηamanaaav\displaystyle\left.+\frac{1}{2!}\sum_{mna}(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{v}}\rho^{\prime}_{mnva}P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{a}a^{\prime}_{v}\right.
+13!mnrvab(1)m+n+r+v+a+bρmnrvab\displaystyle\left.+\frac{1}{3!}\sum_{mnrvab}(-1)^{\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}}\rho^{\prime}_{mnrvab}\right.
×Pηamanarabaaav+]|Ψv(0),\displaystyle\times P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime\dagger}_{r}a^{\prime}_{b}a^{\prime}_{a}a^{\prime}_{v}+\ldots\Bigg{]}\ket{\Psi^{(0)}_{v}}\,, (151)

and

Pη|Ψv\displaystyle P_{\eta}\ket{\Psi^{\prime}_{v}} =[1+ma(ρma)Pηamaa\displaystyle=\left[1+\sum_{ma}(\rho^{\prime}_{ma})^{*}P_{\eta}a^{\prime\dagger}_{m}a^{\prime}_{a}\right.
+mv(ρmv)Pηamav\displaystyle\left.+\sum_{m\neq v}(\rho^{\prime}_{mv})^{*}P_{\eta}a^{\prime\dagger}_{m}a^{\prime}_{v}\right.
+12!mnab(ρmnab)Pηamanabaa\displaystyle\left.+\frac{1}{2!}\sum_{mnab}(\rho^{\prime}_{mnab})^{*}P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{b}a^{\prime}_{a}\right.
+12!mna(ρmnva)Pηamanaaav\displaystyle+\frac{1}{2!}\sum_{mna}(\rho^{\prime}_{mnva})^{*}P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime}_{a}a^{\prime}_{v}
+13!mnrvab(ρmnrvab)\displaystyle+\frac{1}{3!}\sum_{mnrvab}(\rho^{\prime}_{mnrvab})^{*}
×Pηamanarabaaav+]|Ψv(0).\displaystyle\times P_{\eta}a^{\prime\dagger}_{m}a^{\prime\dagger}_{n}a^{\prime\dagger}_{r}a^{\prime}_{b}a^{\prime}_{a}a^{\prime}_{v}+\ldots\Bigg{]}\ket{\Psi^{(0)}_{v}}\,. (152)

Substituting Eqs. (B.2.2) and (B.2.2) into Eq. (149) and comparing like terms, we obtain the following selection rules for the PM cluster amplitudes

(1)m+aρma\displaystyle(-1)^{\ell_{m}+\ell_{a}}\rho^{\prime}_{ma} =(ρma),\displaystyle=(\rho^{\prime}_{ma})^{*}\,, (153)
(1)m+vρmv\displaystyle(-1)^{\ell_{m}+\ell_{v}}\rho^{\prime}_{mv} =(ρmv),\displaystyle=(\rho^{\prime}_{mv})^{*}\,,
(1)m+n+a+bρmnab\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{a}+\ell_{b}}\rho^{\prime}_{mnab} =(ρmnab),\displaystyle=(\rho^{\prime}_{mnab})^{*}\,,
(1)m+n+v+aρmnva\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{v}+\ell_{a}}\rho^{\prime}_{mnva} =(ρmnva),\displaystyle=(\rho^{\prime}_{mnva})^{*}\,,
(1)m+n+r+v+a+bρmnrvab\displaystyle(-1)^{\ell_{m}+\ell_{n}+\ell_{r}+\ell_{v}+\ell_{a}+\ell_{b}}\rho^{\prime}_{mnrvab} =(ρmnrvab).\displaystyle=(\rho^{\prime}_{mnrvab})^{*}\,.

Clearly, if ψm\psi^{\prime}_{m} and ψa\psi^{\prime}_{a} have the same nominal parity then the first of Eqs. (153) implies that ρma\rho^{\prime}_{ma} is real. On the other hand, if ψm\psi^{\prime}_{m} and ψa\psi^{\prime}_{a} have opposite nominal parities then this equation implies that ρma\rho^{\prime}_{ma} is imaginary. The rest of Eqs. (153) have the same interpretation for ρmv\rho^{\prime}_{mv} and other double and valence triple amplitudes.

References