The frequency-independent Coulomb–Breit operator gives rise to the most accurate treatment of two-electron interaction in the non-quantum-electrodynamics regime. The Breit interaction in the Coulomb gauge consists of magnetic and gauge contributions. The high computational cost of the gauge term limits the application of the Breit interaction in relativistic molecular calculations. In this work, we apply the Pauli component integral–density matrix contraction scheme for gauge interaction with a maximum spin- and component separation scheme. We also present two different computational algorithms for evaluating gauge integrals. One is the generalized Obara–Saika algorithm, where the Laplace transformation is used to transform the gauge operator into Gaussian functions and the Obara–Saika recursion is used for reducing the angular momentum. The other algorithm is the second derivative of Coulomb interaction evaluated with Rys-quadrature. This work improves the efficiency of performing Dirac–Hartree–Fock with the variational treatment of Breit interaction for molecular systems. We use this formalism to examine relativistic trends in the Periodic Table and analyze the relativistic two-electron interaction contributions in heavy-element complexes.

Four-component Dirac–Hartree–Fock (DHF) is the starting point for accurate all-electron relativistic calculations.1–4 The frequency-independent instantaneous two-electron interaction provides the most accurate description5–8 of electron–electron interaction before going to a genuine relativistic quantum electrodynamics theory of many-electron systems.9–14 Derived from quantum electrodynamics, the form of the instantaneous two-electron interaction operator, accurate to O(c2), depends on the choice of gauge. In the Feynman gauge, it can be written as

VeeF=i=1Nj>i(gC(i,j)+gG(i,j)),
(1)
gC(i,j)=1rij,
(2)
gG(i,j)=αiαjrij,
(3)

where {i, j} are electron indices. Equation (3) is also known as the Gaunt term. The components of the α matrices are defined as

αi,J=02σJσJ02,J={x,y,z},
(4)

and σ consists of Pauli matrices,

I=1001,σx=0110,σy=0ii0,σz=1001.
(5)

In the Coulomb gauge, the instantaneous two-electron operator is

VeeC=i=1Nj>i(gC(i,j)+gB(i,j)),
(6)
gB(i,j)=12αiαjrij+αirijαjrijrij3,
(7)

where the Breit term [Eq. (7)] includes the magnetic interaction and an additional gauge term. Note that the magnetic interaction of the Breit term in the Coulomb gauge is half of the Gaunt term in the Feynman gauge.

While the Dirac–Coulomb operator accounts for the leading two-electron energy correction in the relativistic regime, the frequency-independent Breit or Gaunt term cannot be ignored for accurate description of electronic structures of late-row and heavy elements.5–8 The difference in the two-electron interaction in Feynman and Coulomb gauge arises from different truncation schemes in the instantaneous interaction approximation, leading to gauge-dependent results. Although both forms of two-electron interactions [VeeF in Eq. (1) and VeeC in Eq. (6)] are accurate to O(c2), there has been work showing that the Breit operator in the Coulomb gauge is more accurate than the Gaunt operator in the Feynman gauge.7,8 It was also suggested that the variational treatment of the frequency-independent Breit interaction is important for both correlation calculations and property evaluations.15,16 DHF with the Coulomb and Gaunt interactions has been implemented in quantum chemistry software packages for atomic and molecular systems.17–30 However, the gauge term in the Breit interaction is computationally more expensive in both integral evaluation and integral–density contraction.

In the 1980s, frequency-dependent and -independent Breit interactions were included perturbatively in a multi-configuration Dirac-Fock formalism with finite difference methods in atomic systems.31–33 Later on, variational treatment of the frequency-independent Breit term in atomic systems with kinetically balanced Slater type basis34 and Gaussian basis35–37 was realized computationally.

Molecular calculation with the Breit interaction came much later. Since radial-angular separation is not applicable for molecular systems, an analytical integral evaluation algorithm for the gauge term in Gaussian type orbitals is required.38–40 The first practical gauge integral evaluation using the McMurchie–Davidson algorithm was implemented and applied for molecular systems by Quiney and co-workers.41,42 Recently, a gauge integral with the generalized Rys-quadrature method was developed for molecules with the density fitting technique.28,43 Still, the high computational cost of the gauge term limits its application in molecular relativistic calculations.44 

Some techniques have been applied to reduce the computational cost of Dirac–Hartree–Fock calculations, such as the quaternion formalism,23 resolution of the identity,28,45 and quasi-four component approximations.46,47 Recently, we introduced a Pauli matrix representation with an optimal spin- and component separation algorithm that results in a minimal floating-point count algorithm for integral–density contraction.30 

In this work, we generalize the integral–density contraction in the Pauli spinor representation in the previous work30 to gauge term contraction. A restricted kinetically balanced (RKB) Gaussian basis is used in this work. Component- and spin-separations48 are applied for the Fock matrix and density matrices for maximal optimization of the contraction scheme. We also derived and developed two new methods for gauge term integral evaluation in Gaussian type orbitals: one uses the generalized Obara–Saika recursion, and the other is the second derivative of Coulomb interaction with Rys quadrature. This work includes the gauge term variationally in a relativistic four-component Dirac–Hartree–Fock framework for molecular calculations.

All expressions in this work are in atomic units. The eigenfunction of the Dirac Hamiltonian ψ has a four-spinor form,

ψ=ϕLϕS,
(8)

where ϕL and ϕS are spinors of the large and small components, expanded in spinor basis,

ϕpL=τμ=1Ncμτ,pLχμτL,ϕpS=τμ=1Ncμτ,pSχμτS,
(9)

where τ ∈ {α, β} and N is the number of spatial basis functions. The large component basis is defined as

χμαL=χμ0,χμβL=0χμ,
(10)

where χμ is the one component spatial basis function. The relationship between the large and small component spinor basis can be defined via the restricted kinetic balance (RKB) condition49–51 to ensure the correct nonrelativistic limit of the positive energy states.52,53 The RKB condition can be written as

χμτS=12cσpχμτL,
(11)

where c is the speed of light and p = −i is the linear momentum operator.

Introducing the RKB condition [Eq. (11)] into the Dirac equation results in the four-component Dirac formalism in matrix representation,54 

FLLFLSFSLFSSCL+CLCS+CS=IS00I12mc2TCL+CLCS+CSϵ+00ϵ,
(12)

where T and S are the kinetic energy and overlap matrices, respectively. The solution of Eq. (12) consists of sets of positive/negative eigenvalues ({ϵ+}, {ϵ}) with corresponding molecular orbital coefficients CL+CS+ and CLCS for the positive/negative energy solutions.

The computation of integrals of operators in Eq. (6) is the dominant cost in building the Dirac–Coulomb–Breit Hartree–Fock matrix and is usually done in the atomic-orbital (AO) basis. To build the AO Dirac–Coulomb–Breit Hartree–Fock, one can scatter the four-component density matrix into 16 component- and spin-blocked density matrices,

Dμτ1,ντ2XY=i=1Necμτ1,iXcντ2,iY*,X,Y={L,S},
(13)

where τ1, τ2, τ3, τ4 ∈ {α, β}. With the RKB condition [Eq. (11)], component- and spin-blocked integrals can be constructed from four-index scalar AO integrals. Then, one can build the Fock matrix by contracting AO spinor integrals with component- and spin-blocked density matrices, resulting in the contraction scheme in the two-spinor basis [see Eqs. (A1)(A4) in  Appendix A].

In a recent study,30 we introduced a minimal floating-point-operation (FLOP) count algorithm based on the Pauli quaternion formalism of maximal component- and spin-separation. In this approach, AO density matrices are further component- and spin-separated and directly contracted with scalar integrals with a minimal FLOP approach but at the cost of an increased arithmetic complexity.

In this work, we extend the Pauli quaternion formalism for building the gauge term in the Dirac–Coulomb–Breit Hartree–Fock equation. Note that the magnetic term [first term in Eq. (7)] is one-half of the Gaunt term, which was previously derived and implemented.30 We refer readers to Ref. 30 for detailed implementation of the Gaunt term the with Pauli quaternion formalism.

In order to build the gauge term contribution in the Pauli component basis, both gauge integrals and density matrices need to be expressed in the Pauli matrix representation. The Pauli components of the density matrix can be transformed from the spin-blocked matrices,54 

Ds=Dαα+Dββ,Dz=DααDββ,Dx=Dαβ+Dβα,Dy=i(DαβDβα),

for each component-block (LL, SS, LS, SL).

The separation of gauge integrals that contract with the density matrices in the Pauli component basis utilizes the Dirac identity and orthogonality of the quaternion. Mathematical derivations are lengthy and complex, and we only present the key equations and the final working expressions [Eqs. (A17)(A30) in  Appendix A].

Collecting all Pauli components, the two-spinor Fock matrix can be constructed:

FXY=12FsXY+FzXYFxXYiFyXYFxXY+iFyXYFsXYFzXY,
(14)

where X, Y ∈ {L, S}. The Dirac–Hartree–Fock equation can be solved self-consistently.

In this work, we derived and implemented two methods to evaluate the gauge integral with the operator form defined as

(α1r12)(α2r12)2r123=α1r12r122r123α2.

In  Appendix B, we show the working equations to evaluate the gauge integral using the Rys-quadrature approach in terms of second derivatives of the four-center two-electron Coulomb integral. In  Appendix C, we derived the recursion relations using the generalized Obara–Saika algorithm. Both methods are implemented and benchmarked against each other, and the resulting integrals only differ by less than 10−12 a.u. Integral code optimization requires extensive work, and therefore, we will not provide timing comparisons between Rys-quadrature and Obara–Saika algorithms for the gauge term.

In this section, we focus on analyzing the memory requirement and computational cost for building the Dirac–Fock matrix from four-index electron-repulsion-integrals (ERIs). We use the in-core formalism where all four-index integrals are computed and stored in memory for the theoretical cost analysis. In this analysis, we used a Kramers-unrestricted condition where all four Pauli components are non-zero. The memory requirement and FLOP count for Dirac–Coulomb and Gaunt have been previously analyzed in Ref. 30. Note that the magnetic contribution in the Breit term has the same computational cost as the Gaunt term. Therefore, the previous cost analyses were directly used in this work with the addition of the gauge contribution.

The gauge term is built from 16 four-index scalar (or one-component) integrals shown in Eqs. (A11)(A14). In the two-spinor RKB approach, these 16 types of gauge ERIs are assembled into 64 two-spinor ERIs, and the integral–density contraction is carried out in the two-spinor form [see Eqs. (A1)(A4)]. In contrast, the Pauli component approach does not assemble the 16 gauge ERIs into a large dimension, although several auxiliary scalar integrals [e.g., (σ · σ), (σ × σ), and (σσ)xz + (σσ)zx] can be constructed to reduce the cost of contraction. Instead, we scatter the Dirac–Fock and density matrices into the Pauli representation so that all gauge ERIs remain in the scalar form and the integral–density contraction is done in Pauli quaternion representation (see Sec. II B). Tables I and II list the estimated computational costs in terms of memory requirement and FLOP count for forming the Dirac–Fock matrix from integrals.

TABLE I.

Analysis of memory requirement in terms of the equivalent number of unique real-valued one-component four-index electron-repulsion-integral tensors needed for forming the four-component Dirac–Coulomb–Breit Hamiltonian. N number of atomic basis functions are used in this estimate. Dense matrices with no integral or matrix symmetry are considered.

Equivalent no. of real-valued N4 ERI tensors
Four-component schemeDirac–CoulombaBreitbTotal
Pauli matrix RKB basis 36 35 71 
Two-spinor RKB basisc 96 128 224 
Equivalent no. of real-valued N4 ERI tensors
Four-component schemeDirac–CoulombaBreitbTotal
Pauli matrix RKB basis 36 35 71 
Two-spinor RKB basisc 96 128 224 
a

Dirac–Coulomb including the (SS|SS) contribution.

b

Analysis of the magnetic contribution in the Breit term is from Ref. 30.

c

Two-spinor integrals are stored as complex-valued quantities, occupying two double-precision storage.

TABLE II.

Analysis of FLOP count in double-precision operations for forming the four-component Dirac–Coulomb–Breit Fock matrix (FLL, FSS, FLS). N number of atomic basis functions are used in this estimate. FLOP count is separated into Coulomb J and exchange K contractions. Dense matrices with no integral or matrix symmetry are considered.

Number of integral–density contractions (J, K)
Four-component schemeDirac–CoulombbBreitcTotal FLOPa
Pauli matrix RKB basis (21, 53) (44, 189) 1228N4 
Two-spinor RKB basis (64, 48) (64, 96) 2176N4 
Number of integral–density contractions (J, K)
Four-component schemeDirac–CoulombbBreitcTotal FLOPa
Pauli matrix RKB basis (21, 53) (44, 189) 1228N4 
Two-spinor RKB basis (64, 48) (64, 96) 2176N4 
a

Floating point operation (FLOP) is defined as an arithmetic procedure (+, −, ×, ÷) of two double-precision numbers. Because integrals for the Pauli matrix RKB basis are stored as real-valued matrices and density matrices are complex-valued, each contraction step requires 4N4 FLOPs. For the two-spinor RKB basis, because integrals are complex-valued, each contraction step requires 8N4 FLOPs.

b

Dirac–Coulomb including the (SS|SS) contribution.

c

Analysis of the magnetic contribution in the Breit term is from Ref. 30.

Table I shows that the Pauli matrix formalism has a much smaller memory storage requirement compared to the two-spinor matrix form. This is mainly because in the Pauli matrix basis, all integrals are stored in one-component real-valued scalar form. In contrast, in the two-spinor basis, scalar integrals are assembled into two-component complex-valued tensors. As a result, the in-core integral storage requirement is much bigger in the two-spinor basis than in the Pauli matrix basis.

Table II gives the FLOP count for each type of Coulomb J and exchange K contractions with the density. For both the Dirac–Coulomb and the Breit terms, we see that the Pauli matrix formalism requires a much smaller number of J contractions than the two-spinor approach. However, the number of K matrix formations is much bigger in the Pauli basis than in the two-spinor basis. Since all integrals and density matrices in the two-spinor basis are complex-valued, each integral–density contraction requires 8N4 FLOPs. In contrast, each integral–density contraction needs 4N4 FLOPs in the Pauli basis because integrals are real-valued. As a result, the computational cost in terms of FLOP count is much more efficient in the Pauli basis.

The purpose of the data presented here is to display the magnitude of the effect of the Dirac–Coulomb–Breit Hamiltonian in atomic systems. In particular, DHF energies are compared to one another in this section. Integral computation is performed with the LIBCINT integral library.55 All calculations in this section are performed with a development version of the Chronus Quantum software package.56 The speed of light utilized in this section is 137.035 999 679 94 a.u.

Ground state DHF energies were computed utilizing the Dirac–Coulomb and Dirac–Coulomb–Breit Hamiltonians. All atoms are defined to be neutral in these calculations. A core Hamiltonian guess is utilized, such that no spin multiplicity need to be defined for the wavefunction guess. The self-consistent-field optimization proceeds in the Kramers unrestricted framework in which the lowest Ne number of positive energy orbitals is singly occupied. In order to analyze the importance of relativistic two-electron corrections in Dirac Hartree–Fock, we also compare the results obtained using the (LL|LL) approximation (previously referred to as the Bare–Coulomb30 approximation), i.e., only (LL|LL) integrals are used without contributions from the small component. The four-component Hamiltonians presented here differ only in their treatment of the two-electron interaction. This allows for the isolation of the effect of various two-electron corrections, particularly the terms of the Dirac–Coulomb–Breit Hamiltonian presented here.

The ground state energies of selected atomic species from across the Periodic Table were computed and relativistic two-electron contributions are compared in Table III. To obtain the ground state energies, we have carried out separate self-consistent-field (SCF) calculations with each of the three Hamiltonians and the (LL|LL) approximation. Relativistic two-electron corrections are defined as the differences between different SCF energies (see the caption of Table III). The uncontracted ANO-RCC basis58–59 set was utilized for all cases. The results are listed in Table III and plotted in Fig. 1. As the atomic number Z increases, the Dirac–Coulomb (ΔEDC) and Breit (ΔEBreit) corrections increase as expected, as shown in Fig. 1(a). Compared to ΔEDC, the Breit correction becomes less important as the atomic number increases, as indicated by the decreasing ratio of ΔEBreitΔEDC [see Fig. 1(b)].

TABLE III.

Ground state SCF energies (Hartrees) for selected atoms using finite width nuclei and the uncontracted ANO-RCC basis.58–59 ΔEDC = EDCELLLL, ΔEBreit = EDCBEDC, ΔEMag.=12(EDCGEDC), and ΔEGauge = ΔEBreit − ΔEMag., where EDCB, EDCG, EDC, and ELLLL are the ground state SCF energies calculated with the Dirac–Coulomb–Breit (DCB), Dirac–Coulomb–Gaunt (DCG), and Dirac–Coulomb (DC) Hamiltonians, respectively, and the (LL|LL) approximation.

ElementZELLLLEDCΔEDCEDCBΔEBreitΔEMag.ΔEGauge
−99.540 92 −99.504 74 0.036 18 −99.493 29 0.011 44 0.005 99 0.005 46 
Na 11 −162.159 79 −162.077 29 0.082 49 −162.053 92 0.023 37 0.012 38 0.010 99 
Si 14 −289.662 42 −289.448 67 0.213 76 −289.393 78 0.054 89 0.029 42 0.025 47 
Cl 17 −461.406 68 −460.947 16 0.459 51 −460.838 80 0.108 36 0.058 57 0.049 79 
Ca 20 −680.572 25 −679.709 87 0.862 37 −679.518 88 0.191 00 0.103 88 0.087 11 
Ti 22 −854.108 69 −852.857 54 1.251 14 −852.592 70 0.264 84 0.144 48 0.120 36 
Cu 29 −1 657.190 12 −1 653.453 40 3.736 72 −1 652.775 81 0.677 59 0.372 57 0.305 02 
Ag 47 −5 339.393 10 −5 314.625 70 24.767 40 −5 311.055 72 3.569 98 1.985 61 1.584 36 
53 −7 155.395 18 −7 115.798 49 39.596 69 −7 110.390 59 5.407 90 3.014 22 2.393 67 
Ce 58 −8 917.383 68 −8 861.043 87 56.339 81 −8 853.658 36 7.385 51 4.121 82 3.263 68 
Pr 59 −9 298.560 18 −9 238.265 44 60.294 74 −9 230.432 81 7.832 63 4.372 22 3.460 41 
Pm 61 −10 091.178 92 −10 022.330 95 68.847 97 −10 013.546 34 8.784 62 4.905 35 3.879 26 
Sm 62 −10 502.861 37 −10 429.356 22 73.505 15 −10 420.062 39 9.293 83 5.190 33 4.103 50 
Tb 65 −11 801.625 31 −11 712.825 12 88.800 18 −11 701.892 28 10.932 8 6.108 49 4.824 36 
Ho 67 −12 721.922 60 −12 621.573 98 100.348 62 −12 609.433 93 12.140 05 6.784 47 5.355 58 
Lu 71 −14 699.426 40 −14 572.526 42 126.899 98 −14 557.677 23 14.849 19 8.301 18 6.548 01 
Au 79 −19 231.541 95 −19 035.579 05 195.962 91 −19 013.938 49 21.640 56 12.101 75 9.538 81 
ElementZELLLLEDCΔEDCEDCBΔEBreitΔEMag.ΔEGauge
−99.540 92 −99.504 74 0.036 18 −99.493 29 0.011 44 0.005 99 0.005 46 
Na 11 −162.159 79 −162.077 29 0.082 49 −162.053 92 0.023 37 0.012 38 0.010 99 
Si 14 −289.662 42 −289.448 67 0.213 76 −289.393 78 0.054 89 0.029 42 0.025 47 
Cl 17 −461.406 68 −460.947 16 0.459 51 −460.838 80 0.108 36 0.058 57 0.049 79 
Ca 20 −680.572 25 −679.709 87 0.862 37 −679.518 88 0.191 00 0.103 88 0.087 11 
Ti 22 −854.108 69 −852.857 54 1.251 14 −852.592 70 0.264 84 0.144 48 0.120 36 
Cu 29 −1 657.190 12 −1 653.453 40 3.736 72 −1 652.775 81 0.677 59 0.372 57 0.305 02 
Ag 47 −5 339.393 10 −5 314.625 70 24.767 40 −5 311.055 72 3.569 98 1.985 61 1.584 36 
53 −7 155.395 18 −7 115.798 49 39.596 69 −7 110.390 59 5.407 90 3.014 22 2.393 67 
Ce 58 −8 917.383 68 −8 861.043 87 56.339 81 −8 853.658 36 7.385 51 4.121 82 3.263 68 
Pr 59 −9 298.560 18 −9 238.265 44 60.294 74 −9 230.432 81 7.832 63 4.372 22 3.460 41 
Pm 61 −10 091.178 92 −10 022.330 95 68.847 97 −10 013.546 34 8.784 62 4.905 35 3.879 26 
Sm 62 −10 502.861 37 −10 429.356 22 73.505 15 −10 420.062 39 9.293 83 5.190 33 4.103 50 
Tb 65 −11 801.625 31 −11 712.825 12 88.800 18 −11 701.892 28 10.932 8 6.108 49 4.824 36 
Ho 67 −12 721.922 60 −12 621.573 98 100.348 62 −12 609.433 93 12.140 05 6.784 47 5.355 58 
Lu 71 −14 699.426 40 −14 572.526 42 126.899 98 −14 557.677 23 14.849 19 8.301 18 6.548 01 
Au 79 −19 231.541 95 −19 035.579 05 195.962 91 −19 013.938 49 21.640 56 12.101 75 9.538 81 
FIG. 1.

(a) Contributions of Dirac–Coulomb and Breit relativistic corrections to the ground state SCF energy for selected elements. (b) Ratio between ΔEBreit and ΔEDC ground state energies. (c) Contribution of magnetic and gauge terms of the Breit Hamiltonian to the ground state SCF energy for selected elements. (d) Ratio between the gauge and magnetic portions of the Breit contribution to ground state energies.

FIG. 1.

(a) Contributions of Dirac–Coulomb and Breit relativistic corrections to the ground state SCF energy for selected elements. (b) Ratio between ΔEBreit and ΔEDC ground state energies. (c) Contribution of magnetic and gauge terms of the Breit Hamiltonian to the ground state SCF energy for selected elements. (d) Ratio between the gauge and magnetic portions of the Breit contribution to ground state energies.

Close modal

We plot in Fig. 1(c) the magnetic and gauge contributions (ΔEMag and ΔEGauge) in the Dirac–Coulomb–Breit Hamiltonian and their ratio ΔEGaugeΔEMag in Fig. 1(d). Both the magnetic and gauge terms grow exponentially with respect to the atomic number and both contributions are equally important in the Breit Hamiltonian. Figure 1(d) shows that the ratio ΔEGaugeΔEMag approaches unity toward light elements and ∼0.78 toward heavy elements. Recall that the Breit operator is the Gaunt term in the Feynman gauge, and in the Coulomb gauge, consists of half of the Gaunt term (i.e., magnetic contribution) and the gauge term. Therefore, another way to understand Fig. 1(d) is that the energetic difference between the Feynman gauge and the Coulomb gauge is approaching ∼0% for light elements and ∼20% for heavy elements.

In this section, we analyze the contribution of the Breit Hamiltonian to the ground state energy and its effect on the electronic structure characteristics of selected heavy-element compounds from the actinyl oxycation series. Linear geometries and experimental bond lengths of 1.76, 1.75, and 1.74 Å were utilized for uranyl, neptunyl, and plutonyl, respectively.60 

Table IV gives the energetic decomposition of different relativistic two-electron interaction terms in the total energy of uranyl, neptunyl, and plutonyl. The uncontracted ANO-RCC-VDZP basis set was used for both the oxygen and actinide atoms. The main trends seen across the actinyl series reflect the growing nuclear charge of the actinide atom and mirror the atomic periodic trends seen in Sec. III B. Although the relative contribution of the Breit term is small (0.1% of the total energy), its effect on electronic structure characteristics of the molecular bonding could be significant.

TABLE IV.

Ground state SCF energies (Hartrees) and relativistic two-electron contributions in three actinyls using finite width nuclei and uncontracted ANO-RCC-VDZP basis. ΔEDC = EDCELLLL, ΔEBreit = EDCBEDC, ΔEMag.=12(EDCGEDC), and ΔEGauge = ΔEBreit − ΔEMag, where EDCB, EDCG, EDC, and ELLLL are the ground state SCF energies calculated with the Dirac–Coulomb–Breit (DCB), Dirac–Coulomb–Gaunt (DCG), and Dirac–Coulomb (DC) Hamiltonians, respectively, and the (LL|LL) approximation.

CompoundFormulaELLLLEDCΔEDCEDCBΔEBreitΔEMag.ΔEGauge
Uranyl UO22+ −28 572.422 12 −28 202.267 63 370.154 49 −28 164.716 23 37.551 40 20.984 33 16.567 07 
Neptunyl NpO22+ −29 383.809 43 −28 996.386 29 387.423 14 −28 957.313 71 39.072 57 21.831 54 17.241 03 
Plutonyl PuO22+ −30 211.253 61 −29 805.949 66 405.303 94 −29 765.307 59 40.642 07 22.705 34 17.936 74 
CompoundFormulaELLLLEDCΔEDCEDCBΔEBreitΔEMag.ΔEGauge
Uranyl UO22+ −28 572.422 12 −28 202.267 63 370.154 49 −28 164.716 23 37.551 40 20.984 33 16.567 07 
Neptunyl NpO22+ −29 383.809 43 −28 996.386 29 387.423 14 −28 957.313 71 39.072 57 21.831 54 17.241 03 
Plutonyl PuO22+ −30 211.253 61 −29 805.949 66 405.303 94 −29 765.307 59 40.642 07 22.705 34 17.936 74 

Density of states plots were generated using the occupied Dirac–Coulomb and Dirac–Coulomb–Breit molecular orbitals for the three actinyls presented in Table IV. These plots are presented in Fig. 2 and show a total large component density as well as a large component contribution from each angular momentum orbital type, i.e., s, p, d, and f characters. Orbital character is summed over the three atoms in the molecule.

FIG. 2.

Dirac–Hartree–Fock density of states plots of frontier molecular orbitals of actinyls with Dirac–Coulomb and Dirac–Coulomb–Breit Hamiltonians. Occupied molecular orbitals are shown with energies shifted with respect to the highest occupied molecular orbital (HOMO) for (a) UO22+ using Dirac–Coulomb, (b) UO22+ using Dirac–Coulomb–Breit, (c) NpO22+ using Dirac–Coulomb, (d) NpO22+ using Dirac–Coulomb–Breit, (e) PuO22+ using Dirac–Coulomb, and (f) PuO22+ using Dirac–Coulomb–Breit. Each orbital is broadened using Lorentzian broadening with a half-width at half-maximum of 0.03 eV. Partial density of states is calculated using Mulliken population analysis. The s, p, d, and f characters are summed over each of the three atoms in the molecule.

FIG. 2.

Dirac–Hartree–Fock density of states plots of frontier molecular orbitals of actinyls with Dirac–Coulomb and Dirac–Coulomb–Breit Hamiltonians. Occupied molecular orbitals are shown with energies shifted with respect to the highest occupied molecular orbital (HOMO) for (a) UO22+ using Dirac–Coulomb, (b) UO22+ using Dirac–Coulomb–Breit, (c) NpO22+ using Dirac–Coulomb, (d) NpO22+ using Dirac–Coulomb–Breit, (e) PuO22+ using Dirac–Coulomb, and (f) PuO22+ using Dirac–Coulomb–Breit. Each orbital is broadened using Lorentzian broadening with a half-width at half-maximum of 0.03 eV. Partial density of states is calculated using Mulliken population analysis. The s, p, d, and f characters are summed over each of the three atoms in the molecule.

Close modal

Figures 2(a) and 2(b) show that the highest occupied orbitals of uranyl consist of six 5f-2p and six 6d-2p bonding orbitals with a very small contribution of s character. The occupied nonbonding 5f electron molecular orbitals in neptunyl and plutonyl are the next highest energy orbitals, appearing lower in energy than the 12 bonding orbitals.

Comparing the DOS plots for this series, the energies of bonding molecular orbitals in neptunyl are more spread out than for those of uranyl and plutonyl. Comparing the DOS plots computed using the Dirac–Coulomb and Dirac–Coulomb–Breit Hamiltonians, we see that the Breit term does not significantly alter the character of the orbitals or the general ordering of the molecular orbitals presented here, but it does alter their relative energetics. For example, a distinct sharpening of the highest energy peak in each of these systems can be observed in the Dirac–Coulomb–Breit case. This is a result of narrowing of the frontier orbital energy gap near HOMO. However, for deeper valence orbitals, there is no noticeable difference in the DOS plots computed using Dirac–Coulomb and Dirac–Coulomb–Breit Hamiltonians.

In order to investigate the effect of the Breit operator on energetic observables, the dissociation energies of uranyl, neptunyl, and plutonyl into 2 × O2− and U(VI), Np(VI), and Pu(VI), respectively, were calculated with both Dirac–Coulomb and Dirac–Coulomb–Breit. To compute the SCF energies of atomic An(VI) cations, a core Hamiltonian guess is utilized, such that no spin multiplicity need to be defined for the wavefunction guess. The SCF optimization proceeds in the Kramers unrestricted framework in which the lowest Ne number of positive energy orbitals is singly occupied. These dissociation energies are given in Table V. Note that this study is meant to analyze the importance of the Breit contribution to bond dissociation energy at the Dirac–Hartree–Fock level. Accurate bond dissociation energies will require calculations using electron correlation methods.

TABLE V.

Ground state bond dissociation energies of three actinyls using finite width nuclei and uncontracted ANO-RCC-VDZP basis. Edissoc.=2×E(O2)+E(An(VI))E(AnO22+). ΔEdissoc.Breit=Edissoc.DCBEdissoc.DC.

CompoundFormulaEdissoc.DC (eV)Edissoc.DCB (eV)ΔEdissoc.Breit (eV)
Uranyl UO22+ 178.907 65 179.108 90 0.201 25 
Neptunyl NpO22+ 182.579 94 182.802 70 0.222 76 
Plutonyl PuO22+ 186.566 79 186.810 01 0.243 22 
CompoundFormulaEdissoc.DC (eV)Edissoc.DCB (eV)ΔEdissoc.Breit (eV)
Uranyl UO22+ 178.907 65 179.108 90 0.201 25 
Neptunyl NpO22+ 182.579 94 182.802 70 0.222 76 
Plutonyl PuO22+ 186.566 79 186.810 01 0.243 22 

Dissociation energies, and therefore bonding energies, can be seen to be higher when calculated with Dirac–Coulomb–Breit than with the Dirac–Coulomb Hamiltonian in these compounds, suggesting that the Breit Hamiltonian stabilizes the chemical bonds between actinide and oxygen. This can be understood from the DOS plots in Fig. 2. The Breit Hamiltonian narrows the energy gap among frontier orbitals near HOMO. As a result, orbitals lower than HOMO will contribute more to the molecular bonding with Dirac–Coulomb–Breit than with the Dirac–Coulomb Hamiltonian. The difference between the calculated dissociation energies grows larger with an increasing atomic number of the actinide atom.

In this work, we applied Pauli spinor representation in the restricted–kinetic–balance condition for four-component Dirac–Hartree–Fock calculations with the Coulomb–Breit two-electron interaction. Along with the density–integral contraction formalism, we introduced two different methods for calculating frequency-independent gauge term integrals with the Gaussian basis. One method utilizes the generalized Obara–Saika algorithm, and the other is the second derivative of Coulomb interaction with Rys quadrature.

With this maximally component- and spin-separated formalism, the computational cost of density–integral contraction for gauge term in the Breit interaction is minimized. It is shown that this Pauli spinor representation is equivalent to the two-spinor formalism but with a much lower computational cost.

We applied this method for calculating atomic energies in order to analyze the periodic trend of the Breit term. It was shown that the Breit contribution becomes relatively less important compared to the Dirac–Coulomb term as the atomic number increases. The analysis also suggests that the difference between the Breit terms in the Feynman gauge and the Coulomb gauge approaches ∼20% toward the heavy elements.

We also studied the effect of the Breit term on the electronic structure characteristics of actinyl oxycations. The analysis shows that the Breit term modifies the relative energies of frontier molecular orbitals near HOMO. As a result, it leads to a slightly stronger chemical bond.

Future development will be focused on combining this Pauli spinor formalism with atomic balance52,61 and minimal-FLOP Cholesky integral decomposition62 to further reduce the computational cost of four-component Dirac–Hartree–Fock.

X.L. acknowledges support from the U.S. Department of Energy, Office of Science, Basic Energy Sciences, in the Heavy-Element Chemistry program (Grant No. DE-SC0021100) for the development of electronic structure methods within the four-component Dirac framework.

The authors have no conflicts to disclose.

Shichao Sun: Data curation (equal); Formal analysis (lead); Investigation (lead); Methodology (lead); Writing – original draft (equal); Writing – review & editing (equal). Jordan Ehrman: Data curation (equal); Writing – original draft (equal); Writing – review & editing (equal). Qiming Sun: Conceptualization (supporting); Software (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Xiaosong Li: Conceptualization (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

The gauge term contribution to the Fock matrix in the two-spinor basis reads

Vμτ1,ντ2g,LL=12λτ4κτ3Dλτ4,κτ3SS(μτ1Lσλτ4S|κτ3Sσντ2L)3,
(A1)
Vμτ1,ντ2g,SS=12λτ4κτ3Dλτ4,κτ3LL(μτ1Sσλτ4L|κτ3Lσντ2S)3,
(A2)
Vμτ1,ντ2g,LS=12λτ4κτ3Dλτ4,κτ3LS(μτ1Lσντ2S|κτ3Sσλτ4L)3+Dλτ4,κτ3SL×[(μτ1Lσντ2S|κτ3Lσλτ4S)3(μτ1Lσλτ4S|κτ3Lσντ2S)3],
(A3)
Vμτ1,ντ2g,SL=12λτ4κτ3Dλτ4,κτ3SL(μτ1Sσντ2L|κτ3Lσλτ4S)3+Dλτ4,κτ3LS×[(μτ1Sσντ2L|κτ3Sσλτ4L)3(μτ1Sσλτ4L|κτ3Sσντ2L)3],
(A4)

where subscript “3” denotes that the integral is 1r123 in contrast to 1r12 and superscript g denotes the gauge term contribution.

In this work, we perform component- and spin-separation for gauge integrals. We use the Dirac identity,1(σμ)(σν)=Iμν+iσμ×ν, and introduce ξτ as the spin representation,

ξα=10,ξβ=01.

We have the following identities:

ξασxξβ=1,ξβσxξα=1,ξασyξβ=i,ξβσyξα=i,ξασzξα=1,ξβσzξβ=1,ξαIξα=1,ξβIξβ=1,

otherwise zero.

The gauge integral in Eq. (A1) can be written as

(μτ1Lσ(1)λτ4S|κτ3Sσ(2)ντ2L)3=ξτ1(1)ξτ4(1)ξτ3(2)ξτ2(2)I(1)I(2)κr12r12λ+iI(2)κr12σ(1)r12×λ+iI(1)σ(2)κ×r12r12λσ(2)κ×r12σ(1)r12×λ(μλ|κν)3.
(A5)

Similarly, the integral in the first term of Eq. (A3) is

(μτ1Lσ(1)ντ2S|κτ3Sσ(2)λτ4L)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)νr12r12κ+iI(2)r12κσ(1)r12×ν+iI(1)νr12σ(2)κ×r12σ(1)r12×νσ(2)κ×r12(μν|κλ)3.
(A6)

The integral in the second term of Eq. (A3) is

(μτ1Lσ(1)ντ2S|κτ3Lσ(2)λτ4S)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)νr12r12λ+iI1νr12σ(2)r12×λ+iI2σ(1)r12×νr12λσ(1)r12×νσ(2)r12×λ(μν|κλ)3,
(A7)

and the exchange integral of Eq. (A3) is

(μτ1Lσ(1)λτ4S|κτ3Lσ(2)ντ2S)3=ξτ1(1)ξτ4(1)ξτ3(2)ξτ2(2)I(1)I(2)λr12r12ν+iI(1)λr12σ(2)r12×ν+iI(2)σ(1)r12×λr12νσ(1)r12×λσ(2)r12×ν(μλ|κν)3,
(A8)

and the exchange integral of Eq. (A2),

(μτ1Sσ(1)λτ4L|κτ3Lσ(2)ντ2S)3=ξτ1(1)ξτ4(1)ξτ3(2)ξτ2(2)I(1)I(2)μr12r12ν+iI(1)μr12σ(2)r12×ν+iI(2)σ(1)μ×r12r12νσ(1)μ×r12σ(2)r12×ν(μλ|κν)3,
(A9)

where κ is the nuclear coordinate derivative of the atomic basis χκ. I(2) and σ(2) are the identity and Pauli matrix for electron 2.

For simplicity and brevity, we introduce a simplified notation for two-spinor integrals,

(μτ1Lσντ2S|κτ3Sσλτ4L)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)(ss)+iI(2)σ(1)(σs)+iI(1)(sσ)σ(2)+σ(1)(σσ)σ(2)(μν̄|κ̄λ)3.
(A10)

The “-” over ν, κ means derivatives are taken of these orbitals. To expand the spinor basis integral (μτ1Lσντ2S|κτ3Sσλτ4L)3 as linear combination of second derivatives of spatial basis integrals as shown in Eq. (A10), we introduce the following short notations for the second derivative of spatial basis gauge integrals:

(ss)(μν̄|κ̄λ)3=νr12r12κ(μν|κλ)3,
(A11)
(σs)J(μν̄|κ̄λ)3=(r12×ν)J(r12κ)(μν|κλ)3,
(A12)
(sσ)J(μν̄|κ̄λ)3=(νr12)(κ×r12)J(μν|κλ)3,
(A13)
(σσ)JK(μν̄|κ̄λ)3=(r12×ν)J(κ×r12)K(μν|κλ)3,J,K{x,y,z},
(A14)

where r12r12 in Eqs. (A11)(A14) is the operator evaluated in the integral. We move it outside of the ()3 for brevity. We give the definition for integrals with derivatives on the second and third centers.

Other types of integrals where the derivatives are taken on the second and fourth centers, or the first and fourth centers, can be derived from Eqs. (A11)(A14) using symmetry. For the Vg,LL, Vg,LS, and Vg,SS terms, we need gauge spinor integrals of the type (LS|SL), (SL|LS), (LS|LS). By using the symmetry of the spinor integrals, we can express the gauge integrals of the type (SL|LS), (LS|LS) by Pauli components of (LS|SL). Comparing Eq. (A9) with Eq. (A6), we have

(μτ1Sσντ2L|κτ3Lσλτ4S)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)[I(1)I(2)(ss)+iI(2)σ(1)(σs)+iI(1)(sσ)σ(2)+σ(1)(σσ)σ(2)](μ̄ν|κλ̄)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)(ss)iI(2)σ(1)(σs)iI(1)(sσ)σ(2)+σ(1)(σσ)σ(2)(νμ̄|λ̄κ)3.
(A15)

Similarly, comparing Eq. (A7) with Eq. (A6), we have

(μτ1Lσντ2S|κτ3Lσλτ4S)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)(ss)+iI(2)σ(1)(σs)+iI(1)(sσ)σ(2)+σ(1)(σσ)σ(2)(μν̄|κλ̄)3=ξτ1(1)ξτ2(1)ξτ3(2)ξτ4(2)I(1)I(2)(ss)iI(2)σ(1)(σs)+iI(1)(sσ)σ(2)+σ(1)(σσ)σ(2)(μν̄|λ̄κ)3.
(A16)

The discussion above illustrates the mathematical expressions and notations needed to derive the gauge contribution to the Fock matrix in the component- and spin-separated form. We skip the detailed derivations and only present final working expressions. Note that the gauge contribution to the Fock matrix has a prefactor of 1(2c)2. For brevity, we omit the prefactor in all equations.

The gauge term contribution to the Pauli components of the LL block of the Fock matrix is an exchange type,

Vμν,sg,LL=12Dλκ,sSS(ss)+iDλκ,zSS(sσ)z+Dλκ,xSS(sσ)x+Dλκ,ySS(sσ)y+iDλκ,zSS(σs)z+Dλκ,xSS(σs)x+Dλκ,ySS(σs)y+Dλκ,sSS(σσ)iDλκ,zSS(σ×σ)ziDλκ,ySS(σ×σ)yiDλκ,xSS(σ×σ)x(μλ̄|κ̄ν)3,
(A17)
Vμν,zg,LL=12Dλκ,zSS(ss)+iDλκ,sSS(sσ)ziDλκ,ySS(sσ)x+iDλκ,xSS(sσ)y+iDλκ,sSS(σs)z+iDλκ,ySS(σs)xiDλκ,xSS(σs)y+Dλκ,zSS(σσ)xx(σσ)yy+(σσ)zz+iDλκ,sSS(σ×σ)z+Dλκ,xSS(σσ)xz+(σσ)zx+Dλκ,ySS(σσ)yz+(σσ)zy(μλ̄|κ̄ν)3,
(A18)
Vμν,xg,LL=12Dλκ,xSS(ss)+iiDλκ,ySS(sσ)z+Dλκ,sSS(sσ)xiDλκ,zSS(sσ)y+iiDλκ,ySS(σs)z+Dλκ,sSS(σs)x+iDλκ,zSS(σs)y+Dλκ,xSS(σσ)xx(σσ)yy(σσ)zz+Dλκ,ySS(σσ)xy+(σσ)yx+Dλκ,zSS(σσ)xz+(σσ)zx+iDλκ,sSS(σ×σ)x(μλ̄|κ̄ν)3,
(A19)
Vμν,yg,LL=12Dλκ,ySS(ss)+iiDλκ,xSS(sσ)z+iDλκ,zSS(sσ)x+Dλκ,sSS(sσ)y+iiDλκ,xSS(σs)ziDλκ,zSS(σs)x+Dλκ,sSS(σs)y+Dλκ,ySS(σσ)xx+(σσ)yy(σσ)zz+Dλκ,xSS(σσ)xy+(σσ)yx+iDλκ,sSS(σ×σ)y+Dλκ,zSS(σσ)yz+(σσ)zy(μλ̄|κ̄ν)3,
(A20)

where (σ × σ)x, (σ × σ)y, (σ × σ)z denote (σσ)yz − (σσ)zy, (σσ)zx − (σσ)xz, (σσ)xy − (σσ)yx, respectively.

Similarly, the Pauli components of the SS block of the Fock matrix are also an exchange type,

Vμν,sg,SS=12Dλκ,sLL(ss)+iDλκ,zLL(sσ)z+Dλκ,xLL(sσ)x+Dλκ,yLL(sσ)y+iDλκ,zLL(σs)z+Dλκ,xLL(σs)x+Dλκ,yLL(σs)y+Dλκ,sLL(σσ)iDλκ,zLL(σ×σ)ziDλκ,yLL(σ×σ)yiDλκ,xLL(σ×σ)x(μ̄λ|κν̄)3,
(A21)
Vμν,zg,SS=12Dλκ,zLL(ss)+iDλκ,sLL(sσ)ziDλκ,yLL(sσ)x+iDλκ,xLL(sσ)y+iDλκ,sLL(σs)z+iDλκ,yLL(σs)xiDλκ,xLL(σs)y+Dλκ,zLL(σσ)xx(σσ)yy+(σσ)zz+iDλκ,sLL(σ×σ)z+Dλκ,xLL(σσ)xz+(σσ)zx+Dλκ,yLL(σσ)yz+(σσ)zy(μ̄λ|κν̄)3,
(A22)
Vμν,xg,SS=12Dλκ,xLL(ss)+iiDλκ,yLL(sσ)z+Dλκ,sLL(sσ)xiDλκ,zLL(sσ)y+iiDλκ,yLL(σs)z+Dλκ,sLL(σs)x+iDλκ,zLL(σs)y+Dλκ,xLL(σσ)xx(σσ)yy(σσ)zz+Dλκ,yLL(σσ)xy+(σσ)yx+Dλκ,zLL(σσ)xz+(σσ)zx+iDλκ,sLL(σ×σ)x(μ̄λ|κν̄)3,
(A23)
Vμν,yg,SS=12Dλκ,yLL(ss)+iiDλκ,xLL(sσ)z+iDλκ,zLL(sσ)x+Dλκ,sLL(sσ)y+iiDλκ,xLL(σs)ziDλκ,zLL(σs)x+Dλκ,sLL(σs)y+Dλκ,yLL(σσ)xx+(σσ)yy(σσ)zz+Dλκ,xLL(σσ)xy+(σσ)yx+iDλκ,sLL(σ×σ)y+Dλκ,zLL(σσ)yz+(σσ)zy(μ̄λ|κν̄)3.
(A24)

The gauge contribution to the Pauli components of the LS and SL blocks of the Fock matrix includes both Coulomb and exchange types of integrals. For the LS block, the Coulomb part is

Vμν,sg,LS=Dλκ,sLS(ss)+iJ=x,y,zDλκJLS(sσ)J(μν̄|κ̄λ)3+Dλκ,sSL(ss)+iJ=x,y,zDλκJSL(sσ)J(μν̄|κλ̄)3,
(A25)
Vμν,Jg,LS=iDλκ,sLS(σs)J+K=x,y,zDλκ,KLS(σσ)JK(μν̄|κ̄λ)3+iDλκ,sSL(σs)J+K=x,y,zDλκ,KSL(σσ)JK(μν̄|κλ̄)3,J,K{x,y,z}.
(A26)

The exchange part of LS block is

Vμν,sg,LS=12Dλκ,sSL(ss)+iDλκ,zSL(sσ)z+Dλκ,xSL(sσ)x+Dλκ,ySL(sσ)y+iDλκ,zSL(σs)z+Dλκ,xSL(σs)x+Dλκ,ySL(σs)y+Dλκ,sSL(σσ)iDλκ,zSL(σ×σ)ziDλκ,ySL(σ×σ)yiDλκ,xSL(σ×σ)x(μλ̄|κν̄)3,
(A27)
Vμν,zg,LS=12Dλκ,zSL(ss)+iDλκ,sSL(sσ)ziDλκ,ySL(sσ)x+iDλκ,xSL(sσ)y+iDλκ,sSL(σs)z+iDλκ,ySL(σs)xiDλκ,xSL(σs)y+Dλκ,zSL(σσ)xx(σσ)yy+(σσ)zz+iDλκ,sSL(σ×σ)z+Dλκ,xSL(σσ)xz+(σσ)zx+Dλκ,ySL(σσ)yz+(σσ)zy(μλ̄|κν̄)3,
(A28)
Vμν,xg,LS=12Dλκ,xSL(ss)+iiDλκ,ySL(sσ)z+Dλκ,sSL(sσ)xiDλκ,zSL(sσ)y+iiDλκ,ySL(σs)z+Dλκ,sSL(σs)x+iDλκ,zSL(σs)y+Dλκ,xSL(σσ)xx(σσ)yy(σσ)zz+Dλκ,ySL(σσ)xy+(σσ)yx+Dλκ,zSL(σσ)xz+(σσ)zx+iDλκ,sSL(σ×σ)x(μλ̄|κν̄)3,
(A29)
Vμν,yg,LS=12Dλκ,ySL(ss)+iiDλκ,xSL(sσ)z+iDλκ,zSL(sσ)x+Dλκ,sSL(sσ)y+iiDλκ,xSL(σs)ziDλκ,zSL(σs)x+Dλκ,sSL(σs)y+Dλκ,ySL(σσ)xx+(σσ)yy(σσ)zz+Dλκ,xSL(σσ)xy+(σσ)yx+iDλκ,sSL(σ×σ)y+Dλκ,zSL(σσ)yz+(σσ)zy(μλ̄|κν̄)3.
(A30)

The SL block can be obtained by taking the complex conjugate transpose of the LS block,

Vg,SL=Vg,LS.

The frequency independent gauge term operator is

(α1r12)(α2r12)2r123=α1r12r122r123α2.
(B1)

Define the Cartesian Gaussian basis function as

φ(r-R; n,ζ)=(xRx)nx(yRy)ny(zRz)nzexp[ζ(r-R)2].
(B2)

The integral of the gauge operator in the Gaussian basis is a 3 × 3 second order tensor that has six independent elements: xx, xy, xz, yy, yz, and zz. A component of the gauge integral is defined as

(ab|n12|cd)3,
(B3)

where subscript 3 denotes |r12|−3. n12=(n12x,n12y,n12z) is a vector that denotes the component of the operator, x12n12xy12n12yz12n12z. Thus, xx, xy, xz, yy, yz, and zz components correspond to n12 = (2, 0, 0), (1, 1, 0), (1, 0, 1), (0, 2, 0), (0, 1, 1), and (0, 0, 2), respectively.

The first-order derivative for the Coulomb operator is

11r12=r12r123.
(B4)

We can use integration by parts to compute the first-order derivatives of the two-electron integrals,

ab|11r12|cd=(a(1)b|cd)+(ab(1)|cd).
(B5)

The superscript (1) refers to the gradient of basis function with respect to electron 1. The four-index notations are

ab|11r12|cd=dr1dr1φ*(r1A;a)φ(r1B;b)11r12φ*(r2C;c)φ(r2D;d),(a(u)b|cd)=dr1dr1(1uφ*(r1A;a))φ(r1B;b)1r12φ*(r2C;c)φ(r2D;d).

The superscript (u) in (a(u)b|cd) refers to taking the first derivative with respect to u ∈ {x, y, z} for the particular basis function.

By using the relation [Eq. (B4)], we can decompose the gauge interactions into two terms,

(α1r12)(α2r12)2r123=12(α1r12)α211r12=12(α1r1)α211r1212(α1r2)α211r12.

Taking the integration by parts relation [Eq. (B5)], we then derive an expression to compute a gauge integral from five regular Coulomb integrals,

ab|(α1r12)(α2r12)2r123|cd=12u,v=x,y,zα1,uα2,v×(a(v)r1ub|cd)+(ar1ub(v)|cd)+δuv(ab|cd)(a(v)b|cr2ud)(ab(v)|cr2ud).
(B6)

All five Coulomb integral terms above can be evaluated with the regular electron repulsion integral algorithms.

In the LIBCINT integral library,55 the gauge integrals are computed using the Rys-quadrature approach. Based on the working expression in Eq. (B6), when calling Rys-quadrature for the regular Coulomb integrals, one needs to solve the Rys-polynomials for the total angular moment of the entire Coulomb integral with the total angular momentum raised by two. This roughly has the same cost as the second-order derivatives of Coulomb integrals. In addition, when the restricted–kinetic–balanced basis is used for small components, the total angular momentum of the Coulomb integral is raised by another two. This leads to roughly the same cost as fourth-order derivatives of regular Coulomb integrals.

1. Laplace transformation of the operator

In general, (1|r|)λ can be written in the form of Laplace transform,

1|r|λ=0du2uλ1Γ(λ/2)eu2r2=0du2uλ1Γ(λ/2)φ(r;0,u2).
(C1)

The μ, ν elements of the tensor operator can be expressed as the Laplace transform of Cartesian Gaussian with a total angular momentum of two,

(r12)μ(r12)νr123=0du2u2Γ(3/2)(r12)μ(r12)νeu2(r12)2=0du2u2Γ(3/2)φ(r1r2;1μ+1ν,u2),
(C2)

where 1μ denotes a vector whose μ component is 1, otherwise 0. Thus, the gauge integral can be written as Laplace transform,

(ab|n12|cd)3=0du2u2Γ(3/2)[ab|n12|cd],
(C3)

where [ab|n12|cd] is the overlap integral,

[ab|n12|cd]=dr1dr2φ*(r1A;a,ζa)φ(r1B;b,ζb)×φ(r1r2;n12,u2)φ*(r2C;c,ζc)φ(r2D;d,ζd).
(C4)

Similarly, integral of the operator (r12)xn12x(r12)yn12y(r12)zn12z|r12| can be written as

(ab|n12|cd)1=0du2u2Γ(1/2)[ab|n12|cd],
(C5)

and Eq. (C5) becomes the Coulomb integral when n12 = (0, 0, 0).

2. Recursion relations for auxiliary integrals

In order to use the generalized Obara–Saika recursion introduced in Ref. 63 to compute the gauge term, we define auxiliary integrals,

(ab|n12|cd)3(m)=0du2u2Γ(3/2)[ab|n12|cd]u2ρ+u2m,(ab|n12|cd)1(m)=0du2Γ(1/2)[ab|n12|cd]u2ρ+u2m,
(C6)

where

ρ=ζηζ+η,ζ=ζa+ζb,η=ζc+ζd,

and m is the non-negative integer. The gauge integral is (ab|n12|cd)3(0).

First, we derived the recursion relation for n12 using Eq. (128) in Ref. 63,

(ab|n12+1μ|cd)λ(m)=12ζNμ(n12)(ab|n121μ|cd)λ(m)(ab|n121μ|cd)λ(m+1)+12ηNμ(n12)(ab|n121μ|cd)λ(m)(ab|n121μ|cd)λ(m+1)+Pμ(ab|n12|cd)λ(m)(ab|n12|cd)λ(m+1)+12ζNμ(a)((a-1μ)b|n12|cd)λ(m)((a-1μ)b|n12|cd)λ(m+1)+12ζNμ(b)(a(b-1μ)|n12|cd)λ(m)(a(b-1μ)|n12|cd)λ(m+1)Qμ(ab|n12|cd)λ(m)(ab|n12|cd)λ(m+1)+12ηNμ(c)(ab|n12|(c-1μ)d)λ(m)(ab|n12|(c-1μ)d)λ(m+1)+12ηNμ(d)(ab|n12|c(d-1μ))λ(m)(ab|n12|c(d-1μ))λ(m+1),
(C7)

where λ can be 1 or 3. Although Eq. (C7) is not directly used in the integral evaluation, it is important for deriving the vertical recursion for electron 2 [Eq. (C11)] and the initial integral [Eq. (C14)].

The vertical recursion for the auxiliary Cartesian Gaussian integral for electron 1 (a and b) is

((a+1μ)b|n12|cd)3(m)=(P-A)μ(ab|n12|cd)3(m)+(W-P)μ(ab|n12|cd)3(m+1)+ρ2ζΓ(1/2)Γ(3/2)Nμ(n12)(ab|n121μ|cd)1(m+1)+12ζNμ(a)((a-1μ)b|n12|cd)3(m)ρζ((a-1μ)b|n12|cd)3(m+1)+12ζNμ(b)(a(b-1μ)|n12|cd)3(m)ρζ(a(b-1μ)|n12|cd)3(m+1)+12(ζ+η)Nμ(c)(ab|n12|(c-1μ)d)3(m+1)+Nμ(d)(ab|n12|(c(d-1μ))3(m+1),
(C8)

where μ, ν ∈ {x, y, z}, P=ζaA+ζbBζ, Q=ζcC+ζdDη, and W=ζP+ηQζ+η. Since n12x+n12y+n12z=2, we can define n121μ = 1ν. The integral of the third term of Eq. (C8) can be split into three Coulomb integrals,

(ab|n121μ|cd)1(m+1)=((a+1)νb|cd)1(m+1)(ab|(c+1)νd)1(m+1)+(A-C)ν(ab|cd)1(m+1).
(C9)

To derive the vertical recursion for electron 2, we first note that

(ab|n12+1μ|cd)3(m)=((a+1)μb|n12|cd)3(m)(ab|n12|(c+1)μd)3(m)+(A-C)μ(ab|n12|cd)3(m).
(C10)

Combining Eq. (C10) with the vertical recursion of electron 1 [Eq. (C8)] and the recursion for n12 [Eq. (C7)], we have

(ab|n12|(c+1μ)d)3(m)=(Q-C)μ(ab|n12|cd)3(m)+(W-Q)μ(ab|n12|cd)3(m+1)ρ2ηΓ(1/2)Γ(3/2)Nμ(n12)(ab|n121μ|cd)1(m+1)+12ηNμ(c)(ab|n12|(c-1μ)d)3(m)ρη(ab|n12|(c-1μ)d)3(m+1)+12ηNμ(d)(ab|n12|c(d-1μ))3(m)ρη(ab|n12|c(d-1μ))3(m+1)+12(ζ+η)Nμ(a)((a-1μ)b|n12|cd)3(m+1)+Nμ(b)(a(b-1μ)|n12|(cd)3(m+1).
(C11)

Note that the third term of Eq. (C11) is different from that in Eq. (C8) by a negative sign.

Horizontal recursions for electron 1 and 2 can be obtained from vertical recursions Eqs. (C8) and (C11), respectively,

(a(b+1μ)|n12|cd)3(m)=((a+1μ)b|n12|cd)3(m)+(A-B)μ(ab|n12|cd)3(m),
(C12)
(ab|n12|c(d+1μ))3(m)=(ab|n12|(c+1μ)d)3(m)+(C-D)μ(ab|n12|cd)3(m).
(C13)

With horizontal [Eqs. (C12) and (C13)] and vertical recursion relations [Eqs. (C8) and (C11)], the gauge integral can be expressed in terms of regular ERIs and (00|n12|00)3(m) integrals.

3. (00|n12|00)3(m) auxiliary integrals

Since n12 has a total angular momentum of 2, we can write it as n12 = 1μ + 1ν. The (00|n12|00)3(m) auxiliary integral of the gauge operator can be written in terms of auxiliary integrals of the Coulomb operator using Eq. (C7),

(00|1μ+1ν|00)3(m)=12ζ+12ηNμ(1ν)(00|1ν1μ|00)3(m)00|1ν1μ|003(m+1)+(PμQμ)(00|1ν|00)3(m)(00|1ν|00)3(m+1)=δμν21ζ+1η(00|0|00)1(m+1)Γ(1/2)Γ(3/2)ρ+(PμQμ)Γ(1/2)Γ(3/2)ρ(PνQν)×(00|0|00)1(m+1)(00|0|00)1(m+2),
(C14)

where we use the relation (ab|n12|cd)3(m)(ab|n12|cd)3(m+1)=Γ(1/2)Γ(3/2)ρ(ab|n12|cd)1(m+1). This relation can be derived

(ab|n12|cd)3(m)(ab|n12|cd)3(m+1)=0du2u2Γ(3/2)(ab|n12|cd)u2ρ+u2mu2ρ+u2m+1=0du2u2Γ(3/2)(ab|n12|cd)u2ρ+u2m1u2ρ+u2=01dtρ1/2(1t2)3/22Γ(3/2)ρt21t2(ab|n12|cd)t2m(1t2)=01dtρ1/2(1t2)3/22Γ(3/2)ρ(ab|n12|cd)t2m+2=0du2ρΓ(3/2)(ab|n12|cd)u2ρ+u2m+1=Γ(1/2)Γ(3/2)ρ0du2Γ(1/2)(ab|n12|cd)u2ρ+u2m+1=Γ(1/2)Γ(3/2)ρ(ab|n12|cd)1(m+1),
(C15)

where we used t2=u2ρ+u2, du=ρ1/2(1t2)3/2dt, (ρρ+u2)3/2=(1t2)3/2, and note that Γ(1/2)Γ(3/2)=2.

Using Eq. (C14), (00|n12|00)3(m) integral can be evaluated using

(00|0|00)1(m)=2π5/2ζηζ+ηKabKcdFm(T),
(C16)
Kab=expζaζbζ(A-B)2,
(C17)
Kcd=expζcζdη(C-D)2,
(C18)

where the Boys function Fm(T)=01dtt2mexp(Tt2) with T = ρ(P − Q)2.

1.
K. G.
Dyall
and
K.
Fægri
, Jr.
,
Introduction to Relativistic Quantum Chemistry
(
Oxford University Press
,
2007
).
2.
M.
Reiher
and
A.
Wolf
,
Relativistic Quantum Chemistry
, 2nd ed. (
Wiley VCH
,
2015
).
3.
I. P.
Grant
,
Relativistic Quantum Theory of Atoms and Molecules: Theory and Computation
(
Springer Science & Business Media
,
2007
), Vol. 40.
4.
I.
Lindgren
,
Relativistic Many-Body Theory: A New Field-Theoretical Approach
(
Springer
,
Switzerland
,
2016
).
5.
J. B.
Mann
and
W. R.
Johnson
, “
Breit interaction in multielectron atoms
,”
Phys. Rev. A
4
,
41
51
(
1971
).
6.
K.-N.
Huang
,
M.
Aoyagi
,
M. H.
Chen
,
B.
Crasemann
, and
H.
Mark
, “
Neutral-atom electron binding energies from relaxed-orbital relativistic Hartree-Fock-Slater calculations 2 ≤ Z ≤ 106
,”
At. Data Nucl. Data Tables
18
,
243
291
(
1976
).
7.
C. T.
Chantler
,
T. V. B.
Nguyen
,
J. A.
Lowe
, and
I. P.
Grant
, “
Convergence of the Breit interaction in self-consistent and configuration-interaction approaches
,”
Phys. Rev. A
90
,
062504
(
2014
).
8.
K.
Kozioł
,
C. A.
Giménez
, and
G. A.
Aucar
, “
Breit corrections to individual atomic and molecular orbital energies
,”
J. Chem. Phys.
148
,
044113
(
2018
).
9.
C.
Thierfelder
and
P.
Schwerdtfeger
, “
Quantum electrodynamic corrections for the valence shell in heavy many-electron atoms
,”
Phys. Rev. A
82
,
062503
(
2010
).
10.
P.
Pyykkö
, “
The physics behind chemistry and the Periodic Table
,”
Chem. Rev.
112
,
371
384
(
2012
).
11.
P.
Pyykkö
, “
Relativistic effects in chemistry: More common than you thought
,”
Annu. Rev. Phys. Chem.
63
,
45
64
(
2012
).
12.
W.
Liu
and
I.
Lindgren
, “
Going beyond ‘no-pair relativistic quantum chemistry
,’”
J. Chem. Phys.
139
,
014108
(
2013
).
13.
W.
Liu
, “
Advances in relativistic molecular quantum mechanics
,”
Phys. Rep.
537
,
59
89
(
2014
).
14.
W.
Liu
, “
Essentials of relativistic quantum chemistry
,”
J. Chem. Phys.
152
,
180901
(
2020
).
15.
I. P.
Grant
, “
Relativistic self-consistent fields
,”
Proc. R. Soc. London, Ser. A
262
,
555
576
(
1961
).
16.
E.
Lindroth
,
A.-M.
Martensson-Pendrill
,
A.
Ynnerman
, and
P.
Oster
, “
Self-consistent treatment of the Breit interaction, with application to the electric dipole moment in thallium
,”
J. Phys. B
22
,
2447
2464
(
1989
).
17.
A.
Mohanty
and
E.
Clementi
, “
Dirac-Fock self-consistent field method for closed-shell molecules with kinetic balance and finite nuclear size
,”
Int. J. Quantum Chem.
39
,
487
517
(
1991
).
18.
O.
Visser
,
L.
Visscher
,
P. J. C.
Aerts
, and
W. C.
Nieuwpoort
, “
Relativistic all-electron molecular Hartree-Fock-Dirac-(Breit) calculations on CH4, SiH4, GeH4, SnH4, PbH4
,”
Theor. Chem. Acc.
81
,
405
416
(
1992
).
19.
L.
Visscher
,
O.
Visser
,
P. J. C.
Aerts
,
H.
Merenga
, and
W. C.
Nieuwpoort
, “
Relativistic quantum chemistry: The MOLFDIR program package
,”
Comput. Phys. Commun.
81
,
120
144
(
1994
).
20.
L.
Visscher
, “
Approximate molecular relativistic Dirac-Coulomb calculations using a simple Coulombic correction
,”
Theor. Chem. Acc.
98
,
68
70
(
1997
).
21.
J.
Styszyński
,
X.
Cao
,
G. L.
Malli
, and
L.
Visscher
, “
Relativistic all-electron Dirac-Fock-Breit calculations on xenon fluorides (XeFn, n = 1, 2, 4, 6)
,”
J. Comput. Chem.
18
,
601
608
(
1997
).
22.
B. T.
Saue
,
K.
Faegri
,
T.
Helgaker
, and
O.
Gropen
, “
Principles of direct 4-component relativistic SCF: Application to caesium auride
,”
Mol. Phys.
91
,
937
950
(
1997
).
23.
T.
Saue
and
H. J. A.
Jensen
, “
Quaternion symmetry in relativistic molecular calculations: The Dirac–Hartree–Fock method
,”
J. Chem. Phys.
111
,
6211
6222
(
1999
).
24.
H. M.
Quiney
,
H.
Skaane
, and
I. P.
Grant
, “
Ab initio relativistic quantum chemistry: Four-components good, two-components bad!
,”
Adv. Quantum Chem.
32
,
1
49
(
1998
).
25.
I. P.
Grant
and
H. M.
Quiney
, “
Application of relativistic theories and quantum electrodynamics to chemical problems
,”
Int. J. Quantum Chem.
80
,
283
297
(
2000
).
26.
T.
Yanai
,
T.
Nakajima
,
Y.
Ishikawa
, and
K.
Hirao
, “
A new computational scheme for the Dirac–Hartree–Fock method employing an efficient integral algorithm
,”
J. Chem. Phys.
114
,
6526
6538
(
2001
).
27.
T.
Yanai
,
T.
Nakajima
,
Y.
Ishikawa
, and
K.
Hirao
, “
A highly efficient algorithm for electron repulsion integrals over relativistic four-component Gaussian-type spinors
,”
J. Chem. Phys.
116
,
10122
10128
(
2002
).
28.
M. S.
Kelley
and
T.
Shiozaki
, “
Large-scale Dirac–Fock–Breit method using density fitting and 2-spinor basis functions
,”
J. Chem. Phys.
138
,
204113
(
2013
).
29.
T.
Saue
,
R.
Bast
,
A. S. P.
Gomes
,
H. J. A.
Jensen
,
L.
Visscher
,
I. A.
Aucar
,
R.
Di Remigio
,
K. G.
Dyall
,
E.
Eliav
,
E.
Fasshauer
,
T.
Fleig
,
L.
Halbert
,
E. D.
Hedegård
,
B.
Helmich-Paris
,
M.
Iliaš
,
C. R.
Jacob
,
S.
Knecht
,
J. K.
Laerdahl
,
M. L.
Vidal
,
M. K.
Nayak
,
M.
Olejniczak
,
J. M. H.
Olsen
,
M.
Pernpointner
,
B.
Senjean
,
A.
Shee
,
A.
Sunaga
, and
J. N. P.
van Stralen
, “
The DIRAC code for relativistic molecular calculations
,”
J. Chem. Phys.
152
,
204104
(
2020
).
30.
S.
Sun
,
T. F.
Stetina
,
T.
Zhang
,
H.
Hu
,
E. F.
Valeev
,
Q.
Sun
, and
X.
Li
, “
Efficient four-component Dirac–Coulomb–Gaunt Hartree–Fock in the Pauli spinor representation
,”
J. Chem. Theory Comput.
17
,
3388
3402
(
2021
).
31.
J. P.
Desclaux
, “
Multiconfiguration relativistic DIRAC-FOCK program
,”
Comput. Phys. Commun.
9
,
31
45
(
1975
).
32.
I. P.
Grant
and
B. J.
McKenzie
, “
The transverse electron-electron interaction in atomic structure calculations
,”
J. Phys. B
13
,
2671
(
1980
).
33.
K. G.
Dyall
,
I. P.
Grant
,
C. T.
Johnson
,
F. A.
Parpia
, and
E. P.
Plummer
, “
GRASP: A general-purpose relativistic atomic structure program
,”
Comput. Phys. Commun.
55
,
425
456
(
1989
).
34.
H. M.
Quiney
,
I. P.
Grant
, and
S.
Wilson
, “
The Dirac equation in the algebraic approximation. V. Self-consistent field studies including the Breit interaction
,”
J. Phys. B
20
,
1413
1422
(
1987
).
35.
Y.
Ishikawa
, “
Dirac–Fock Gaussian basis calculations: Inclusion of the Breit interaction in the self-consistent field procedure
,”
Chem. Phys. Lett.
166
,
321
325
(
1990
).
36.
Y.
Ishikawa
,
H. M.
Quiney
, and
G. L.
Malli
, “
Dirac-Fock-Breit self-consistent-field method: Gaussian basis-set calculations on many-electron atoms
,”
Phys. Rev. A
43
,
3270
3278
(
1991
).
37.
F. A.
Parpia
,
A. K.
Mohanty
, and
E.
Clementi
, “
Relativistic calculations for atoms: Self-consistent treatment of Breit interaction and nuclear volume effect
,”
J. Phys. B
25
,
1
16
(
1992
).
38.
F.
Rosicky
, “
On interelectronic magnetic and retardation effects within relativistic Hartree-Fock theory
,”
Chem. Phys. Lett.
85
,
195
198
(
1982
).
39.
P.
Chandra
and
R. J.
Buenker
, “
Relativistic integrals over Breit–Pauli operators using general Cartesian Gaussian functions. II. Two-electron interactions
,”
J. Chem. Phys.
79
,
366
372
(
1983
).
40.
A. K.
Mohanty
, “
Dirac–Fock self-consistent field method for closed-shell molecules including Breit interaction
,”
Int. J. Quantum Chem.
42
,
627
662
(
1992
).
41.
H. M.
Quiney
,
H.
Skaane
, and
I. P.
Grant
, “
Relativistic calculation of electromagnetic interactions in molecules
,”
J. Phys. B
30
,
L829
L834
(
1997
).
42.
H. M.
Quiney
,
H.
Skaane
, and
I. P.
Grant
, “
Relativistic, quantum electrodynamic and many-body effects in the water molecule
,”
Chem. Phys. Lett.
290
,
473
480
(
1998
).
43.
T.
Shiozaki
, “
Communication: An efficient algorithm for evaluating the Breit and spin–spin coupling integrals
,”
J. Chem. Phys.
138
,
111101
(
2013
).
44.
L.
Visscher
, “
The Dirac equation in quantum chemistry: Strategies to overcome the current computational problems
,”
J. Comput. Chem.
23
,
759
766
(
2002
).
45.
M.
Repisky
,
S.
Komorovsky
,
M.
Kadek
,
L.
Konecny
,
U.
Ekström
,
E.
Malkin
,
M.
Kaupp
,
K.
Ruud
,
O. L.
Malkina
, and
V. G.
Malkin
, “
ReSpect: Relativistic spectroscopy DFT program package
,”
J. Chem. Phys.
152
,
184101
(
2020
).
46.
W.
Liu
and
D.
Peng
, “
Infinite-order quasirelativistic density functional method based on the exact matrix quasirelativistic theory
,”
J. Chem. Phys.
125
,
044102
(
2006
).
47.
D.
Peng
,
W.
Liu
,
Y.
Xiao
, and
L.
Cheng
, “
Making four- and two-component relativistic density functional methods fully equivalent based on the idea of ‘from atoms to molecule
,’”
J. Chem. Phys.
127
,
104106
(
2007
).
48.
K. G.
Dyall
, “
An exact separation of the spin-free and spin-dependent terms of the Dirac–Coulomb–Breit Hamiltonian
,”
J. Chem. Phys.
100
,
2118
2127
(
1994
).
49.
R. E.
Stanton
and
S.
Havriliak
, “
Kinetic balance: A partial solution to the problem of variational safety in Dirac calculations
,”
J. Chem. Phys.
81
,
1910
1918
(
1984
).
50.
Y.
Ishikawa
,
R. C.
Binning
, and
K. M.
Sando
, “
Dirac-Fock discrete-basis calculations on the beryllium atom
,”
Chem. Phys. Lett.
101
,
111
114
(
1983
).
51.
K. G.
Dyall
and
K.
Fægri
, “
Kinetic balance and variational bounds failure in the solution of the Dirac equation in a finite Gaussian basis set
,”
Chem. Phys. Lett.
174
,
25
32
(
1990
).
52.
W.
Liu
, “
Ideas of relativistic quantum chemistry
,”
Mol. Phys.
108
,
1679
1706
(
2010
).
53.
Q.
Sun
,
W.
Liu
, and
W.
Kutzelnigg
, “
Comparison of restricted, unrestricted, inverse, and dual kinetic balances for four-component relativistic calculations
,”
Theor. Chem. Acc.
129
,
423
436
(
2011
).
54.
A.
Petrone
,
D. B.
Williams-Young
,
S.
Sun
,
T. F.
Stetina
, and
X.
Li
, “
An efficient implementation of two-component relativistic density functional theory with torque-free auxiliary variables
,”
Eur. Phys. J. B
91
,
169
(
2018
).
55.
Q.
Sun
, “
Libcint: An efficient general integral library for Gaussian basis functions
,”
J. Comput. Chem.
36
,
1664
1671
(
2015
).
56.
D. B.
Williams-Young
,
A.
Petrone
,
S.
Sun
,
T. F.
Stetina
,
P.
Lestrange
,
C. E.
Hoyer
,
D. R.
Nascimento
,
L.
Koulias
,
A.
Wildman
,
J.
Kasper
,
J. J.
Goings
,
F.
Ding
,
A. E.
DePrince
 III
,
E. F.
Valeev
, and
X.
Li
, “
The Chronus quantum (ChronusQ) software package
,”
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
10
,
e1436
(
2020
).
57.
B. O.
Roos
,
R.
Lindh
,
P.-Å.
Malmqvist
,
V.
Veryazov
, and
P.-O.
Widmark
, “
Main group atoms and dimers studied with a new relativistic ANO basis set
,”
J. Phys. Chem. A
108
,
2851
2858
(
2004
).
58.
B. O.
Roos
,
R.
Lindh
,
P.-Å.
Malmqvist
,
V.
Veryazov
, and
P.-O.
Widmark
, “
New relativistic ANO basis sets for transition metal atoms
,”
J. Phys. Chem. A
109
,
6575
6579
(
2005
).
59.
B. O.
Roos
,
R.
Lindh
,
P.-Å.
Malmqvist
,
V.
Veryazov
,
P.-O.
Widmark
, and
A. C.
Borin
, “
New relativistic atomic natural orbital basis sets for lanthanide atoms with applications to the Ce diatom and LuF3
,”
J. Phys. Chem. A
112
,
11431
11435
(
2008
).
60.
D.-C.
Sergentu
,
T. J.
Duignan
, and
J.
Autschbach
, “
Ab initio study of covalency in the ground versus core-excited states and X-ray absorption spectra of actinide complexes
,”
J. Phys. Chem. Lett.
9
,
5583
5591
(
2018
).
61.
L.
Visscher
,
P. J. C.
Aerts
,
O.
Visser
, and
W. C.
Nieuwpoort
, “
Kinetic balance in contracted basis sets for relativistic calculations
,”
Int. J. Quantum Chem.
40
,
131
(
1991
).
62.
T.
Zhang
,
X.
Liu
,
E. F.
Valeev
, and
X.
Li
, “
Toward the minimal floating operation count Cholesky decomposition of electron repulsion integrals
,”
J. Phys. Chem. A
125
,
4258
4265
(
2021
).
63.
S.
Obara
and
A.
Saika
, “
General recurrence formulas for molecular integrals over Cartesian Gaussian functions
,”
J. Chem. Phys.
89
,
1540
(
1988
).