Drachmann’s regularization approach is implemented for floating explicitly correlated Gaussians (fECGs) and molecular systems. Earlier applications of drachmannized relativistic corrections for molecular systems were hindered due to the unknown analytic matrix elements of 1/rix1/rjy-type operators with fECGs. In the present work, one of the 1/r factors is approximated by a linear combination of Gaussians, which results in calculable integrals. The numerical approach is found to be precise and robust over a range of molecular systems and nuclear configurations, and thus, it opens the route toward an automated evaluation of high-precision relativistic corrections over potential energy surfaces of polyatomic systems. Furthermore, the newly developed integration approach makes it possible to construct the matrix representation of the square of the electronic Hamiltonian relevant for energy lower-bound as well as time-dependent computations of molecular systems with a flexible and high-precision fECG basis representation.

The leading-order (α2Eh) relativistic corrections1 have been computed to high precision for two-, three, and four-electron atoms2–8 and two- and three-electron diatomic molecules.6,9–14 We are aware of a single computation of a polyatomic system (H3+) for which the regularized leading-order relativistic corrections were computed (at the equilibrium structure).12 

The terms contributing to the leading-order relativistic correction are large in absolute value and have opposite signs, so the correction values must be converged to several digits to arrive at precise final values. Alternative routes to the mainstream perturbation theory approach have been explored and are promising, potentially also for spectroscopic applications,11,15–22 but currently available only for two-electron (or two-spin-1/2-fermion19) systems.

Hence, it is relevant to pursue also the traditional approach1 and develop automatization schemes for the precise evaluation of the involved singular operators over generally applicable (explicitly correlated) Gaussian basis sets. Due to the incorrect description of the (electron–electron and electron–nucleus) coalescence points of the non-relativistic wave function represented as a linear combination of explicitly correlated Gaussians (ECGs), the highly localized, short-range operator expectation values must be regularized to speed up convergence with respect to the basis size.2,6,12 There have been two major directions to correct for this deficiency of general basis sets: (a) using operator identities, which replace local operators with global expressions, often called “drachmannization,” named after the pioneer of this technique,2 and (b) using integral transformation techniques,6,12 which map the short-range behavior to the long-range, for which analytic asymptotic expressions can be derived, and can be used to replace the long-tail part of the integral.

Despite the engaging formal theoretical background12 of the integral transformation technique,6 it is critical to be able to identify a lower threshold value for the long-range part, which is to be represented by the analytic asymptotic expression. Numerical experience shows that a considerable amount of manual fiddling is required to find the optimal threshold value (separating the important physical features represented by numerical integration and the asymptotic limiting function).12,13 This arbitrariness introduces large uncertainties.

For a precise and automated computation of the leading-order relativistic correction value over a multi-dimensional potential energy surface, potentially at hundreds or thousands of nuclear geometries, a robust methodology is required, which can be used in an automated fashion without much manual adjustment.

The drachmannization approach appears to be more robust than the integral transformation technique since it does not contain any parameters to adjust, but the necessary ECG integrals for molecular systems have been thought to be incalculable (in a closed, analytic form).23 In particular, molecular computations with clamped nuclei can be efficiently carried out using the so-called floating explicitly correlated Gaussians (fECGs), in which the shifted (“floating”) centers (treated as variational parameters) are necessary to efficiently account for the significant wave function amplitude off-centered from the origin. The ⟨1/rix1/rjy⟩-type integral expressions (i, j label electron coordinates; x, y label electron or nuclear coordinates), necessary for the evaluation of the drachmannized relativistic corrections, are not known for fECGs, unless the ECG centers are fixed at the origin (zero shift).

Interestingly, the same type of incalculable integrals appear in lower-bound theory, and most importantly its Pollak–Martinazzo24,25 version, which has been demonstrated to be powerful for atomic systems (zero-shift ECGs with analytic integrals),26,27 but no molecular computations have been performed due to the missing integral expressions. Potentially, similar integrals will be required in time-dependent simulations recently proposed with explicitly correlated Gaussians.28 

In this work, we develop a numerical approach for the computation of the ⟨1/rix1/rjy⟩-type integrals with fECGs by writing one of the 1/r factors as a linear combination of Gaussians. The representation of 1/r as a linear combination of Gaussians has already been successfully used for deformed explicitly correlated Gaussians by Varga and co-workers,29 using the mathematical scheme of Beylkin and Monzón.30 

As a result, a robust numerical drachmannization scheme is developed in this work, and regularized relativistic corrections (computable in a black-box fashion) are reported for two-, three-, and four-electron di- and triatomic molecular systems.

As a starting point, the non-relativistic energy and wave function are computed by solving the electronic Schrödinger equation,
(1)
with the Hamiltonian
(2)
including the potential energy
(3)
where the nuclei have charge number Za and are clamped at Ra(a = 1, …, N). The generally applicable and flexible ansatz used to represent the wave function is
(4)
where A is the anti-symmetrization operator (relevant for indistinguishable fermions) and χ is the spin function written as a linear combination of elementary spin functions coupled to the total electronic spin of the system.31,32 For the spatial part, we use floating explicitly correlated Gaussian functions (fECGs),
(5)
where r collects the 3n Cartesian coordinates of the electrons in the system, AμRn×n is a positive-definite, symmetric matrix, and sμR3n will be referred to as the “shift vector.” I3 is the three-dimensional unit matrix. The Aμ and sμ parameters are optimized by minimization of the non-relativistic energy.
In the non-relativistic quantum electrodynamics approach,1 the non-relativistic energy is perturbatively corrected by relativistic and QED contributions in terms of increasing orders of the α fine-structure constant,
(6)
The E(2) second-order term, known as the leading-order relativistic correction, can be computed as the expectation value of the Breit–Pauli Hamiltonian,1 
(7)
which is written for singlet (or spin averaged) electronic states as
(8)
with
(9)
The terms of HBP(2)in Eq. (8) have traditional names; they are the mass-velocity term, the one-electron Darwin term, the two-electron Darwin term (combined with the spin–spin Fermi contact interaction), and the orbit–orbit term, respectively. All terms, except the orbit–orbit contribution, include highly localized, singular operators, for which the expectation value with the non-relativistic wave function is inefficiently represented over “smooth” fECG basis functions, i.e., the expectation value converges slowly with respect to the number of basis functions.12 
The drachmannization technique2 uses operator identities valid for the exact non-relativistic wave function, (T + V)Ψ = EΨ in Eqs. (1) and (2), to replace singular operators with more globally acting operators,
(10)
(11)
and
(12)
During the evaluation of these expectation values, the following integrals appear:
(13)
for which no analytic result is known with fECG basis functions. In the present work, we propose a numerical approach to compute these integrals to high precision.
We can approximate the 1/r function as a linear combination of Gaussian functions. For practical purposes, this approximation is formulated as
(14)
with the weights and points chosen according to Beylkin and Monzón,30,
(15)
where the step variable is h=baM. The approximation has three input parameters: the M number of Gaussian functions and the a < 0 and b > 0 boundary parameters. To better understand the origin of these parameters, we note that the underlying approximation idea starts with considering the following integral:30,33
(16)
which can be calculated by the s = ln (t/r2)/2 substitution. Then, for α = 1,
(17)
and the integral is approximated using the trapezoidal rule. In this numerical approximation, the lower and upper limits are set to some (finite) a and b values, respectively. The numerical integration limits a and b are chosen so that the value of the integrand is negligible at the boundaries and beyond them. Then, the application of the trapezoidal rule results in Eqs. (14) and (15). The fast convergence of the trapezoidal rule, motivating this form of the integrand, was pointed out in Ref. 33.

First, we have tested the Gaussian approximation for three special cases, for which analytic integral expressions are also available. Figure 1 highlights the robustness and excellent performance of the numerical approach. As a result of extensive numerical testing, we conclude that high-precision results can be obtained robustly without manual fiddling of the parameters of Eq. (15). For further computations, we have selected the following set of values: M = 200, a = −31, and b = 31, which is a computationally well-affordable choice and allows for sufficiently precise results (close to machine precision in double-precision arithmetic) without further adjustment. An alternative, similarly useful and practical parameterization was M = 400, a = −45, and b = 45, for which the computational cost remains low, and was used to test the (M, a, b) = (200, −31, 31) parameterization. Further refinement of these parameters is possible in the future, but these choices have already provided excellent results with a low computational cost and thereby open the route for the computation of relativistic corrections for polyelectronic and polyatomic molecular systems with unprecedented precision in an automated manner.

FIG. 1.

Relative error on a logarithmic scale, log |δ|, of the Gaussian approximation, Eqs. (14) and (15) (with −a = b), tested for special cases for which analytic expressions are also available: (a) Ukl=φk1rijφl; (b) Ukl=φk1rij1rpqφls=0; and (c) Ukl=φk1rij2φl. δ=[Ukl(analytic)Ukl]/Ukl. Randomly selected pairs of basis functions were used to prepare the figure (deposited in the supplementary material), and they exemplify the general behavior observed for all basis sets used in this study.

FIG. 1.

Relative error on a logarithmic scale, log |δ|, of the Gaussian approximation, Eqs. (14) and (15) (with −a = b), tested for special cases for which analytic expressions are also available: (a) Ukl=φk1rijφl; (b) Ukl=φk1rij1rpqφls=0; and (c) Ukl=φk1rij2φl. δ=[Ukl(analytic)Ukl]/Ukl. Randomly selected pairs of basis functions were used to prepare the figure (deposited in the supplementary material), and they exemplify the general behavior observed for all basis sets used in this study.

Close modal

Tables I and II present the test results of this numerical drachmannization approach for two-, three-, and four-electron atoms and the two-electron hydrogen molecule, for which benchmark-quality data were already available in the literature. In these examples, we computed the ground state for each system, but the approach is similarly applicable to excited states, too.

TABLE I.

Demonstration of the numerical drachmannization approach for the ground state of atoms, He: 11S, Li: 22S, and Be: 21S, and comparison with high-precision reference data from the literature. δEnr = ErefEnr in Eh. All computations have been carried out with the same (M, a, b) = (200, −31, 31) parameterization of the Gaussian expansion, Eqs. (14) and (15). Convergence tables are provided in the supplementary material.

δEnr18ii2i2rgi,aZaδ(ria)rgi,jδ(rij)rg
He 6 × 10−11 −13.522 054 7.241 717 15 0.106 345 364 
He3  1 × 10−15 −13.522 016 81 7.241 717 274 0.106 345 371 2 
Li 3 × 10−10 −78.557 5 41.527 76 0.544 322 
Li4a 2 × 10−13 −78.556 143 062 41.527 828 927 0.544 329 79 
Be 5 × 10−7 −270.706 5 141.475 80 1.605 301 5 
Be7a 3 × 10−10 −270.703 76 141.476 010 4 1.605 305 33 
δEnr18ii2i2rgi,aZaδ(ria)rgi,jδ(rij)rg
He 6 × 10−11 −13.522 054 7.241 717 15 0.106 345 364 
He3  1 × 10−15 −13.522 016 81 7.241 717 274 0.106 345 371 2 
Li 3 × 10−10 −78.557 5 41.527 76 0.544 322 
Li4a 2 × 10−13 −78.556 143 062 41.527 828 927 0.544 329 79 
Be 5 × 10−7 −270.706 5 141.475 80 1.605 301 5 
Be7a 3 × 10−10 −270.703 76 141.476 010 4 1.605 305 33 
a

More precise non-relativistic energy, Enr, is available from Refs. 5 and 35 for Li and Be, respectively, but we could use the regularized relativistic corrections available from the cited work for comparison. The basis sets for the present computations were taken from earlier work.15,27

TABLE II.

Illustration of the enhanced convergence rate by the numerical drachmannization approach developed in this work for the example of the H2 molecule (X1Σg+, Rp–p = 1.40 bohr). δEnr = ErefEnr in Eh. All computations have been carried out with the same (M, a, b) = (200, −31, 31) parameterization of the Gaussian expansion, Eqs. (14) and (15). Reference 10: Enr = −1.174 475 714 220 443 4(5) Eh.

δEnr18ii2i2i,aZaδ(ria)i,jδ(rij)HBP(2)
Direct 
1.7 × 10−7 −1.65 0.918 0.0169 −0.205 
1.6 × 10−8 −1.65 0.919 0.0168 −0.205 
1.4 × 10−9 −1.65 0.919 0.016 78 −0.2055 
3.8 × 10−10 −1.654 0.919 0.016 77 −0.2055 
Numerical drachmannization 
1.7 × 10−7 −1.655 0.919 33 0.016 742 −0.206 
1.6 × 10−8 −1.6548 0.919 335 5 0.016 743 1 −0.2058 
1.4 × 10−9 −1.654 80 0.919 335 7 0.016 743 21 −0.205 75 
3.8 × 10−10 −1.654 79 0.919 335 8 0.016 743 24 −0.205 733 
Reference 10  
“Direct” −1.654 837 5(1) 0.919 37(7) 0.016 743 5(4) −0.205 682 (8) 
“St. reg.” −1.654 745 (2) 0.919 336 211 2(18) 0.016 743 278 1(7) −0.205 689 (2) 
“rECG & m.reg.” −1.654 744 522 5(1) 0.919 336 211(2) 0.016 743 278 3(3) −0.205 688 526(7) 
δEnr18ii2i2i,aZaδ(ria)i,jδ(rij)HBP(2)
Direct 
1.7 × 10−7 −1.65 0.918 0.0169 −0.205 
1.6 × 10−8 −1.65 0.919 0.0168 −0.205 
1.4 × 10−9 −1.65 0.919 0.016 78 −0.2055 
3.8 × 10−10 −1.654 0.919 0.016 77 −0.2055 
Numerical drachmannization 
1.7 × 10−7 −1.655 0.919 33 0.016 742 −0.206 
1.6 × 10−8 −1.6548 0.919 335 5 0.016 743 1 −0.2058 
1.4 × 10−9 −1.654 80 0.919 335 7 0.016 743 21 −0.205 75 
3.8 × 10−10 −1.654 79 0.919 335 8 0.016 743 24 −0.205 733 
Reference 10  
“Direct” −1.654 837 5(1) 0.919 37(7) 0.016 743 5(4) −0.205 682 (8) 
“St. reg.” −1.654 745 (2) 0.919 336 211 2(18) 0.016 743 278 1(7) −0.205 689 (2) 
“rECG & m.reg.” −1.654 744 522 5(1) 0.919 336 211(2) 0.016 743 278 3(3) −0.205 688 526(7) 

New numerical results for two-, three-, and four-electron di- and triatomic molecular systems are reported in Table III. These results are the first computation of these leading-order relativistic corrections or improve upon the already available values in the literature. Furthermore, we report the curve of relativistic corrections to the H2+Ã2Σu+ electronic state (Fig. 2) to demonstrate the general applicability of the numerical drachmannization approach along potential energy curves.

TABLE III.

Molecular systemsa: numerically drachmannized expectation values and the leading-order relativistic correction. −δEnr is the estimated convergence error, in Eh of the (variational) non-relativistic energy. All converged and 1–2 additional digits are shown. All computations have been carried out with the same (M, a, b) = (200, −31, 31) parameterization of the Gaussian expansion. Convergence tables are provided in the supplementary material.

δEnr18ii2i2rgi,aZaδ(ria)rgi,jδ(rij)rgHOOHBP(2)
H2+ 10−12b −0.624 952 638 6 0.318 293 258 6 ⋯ ⋯ −0.124 978 757 0 
b34c 10−9 −0.624 952 7 0.318 293 2 ⋯ ⋯ −0.124 978 9 
H3+c 10−9 −1.933 449 7 1.089 654 23 0.018 334 666 −0.057 217 520 −0.221 442 3 
i12c 10−9 −1.933 424 1.089 655 0.018 335 −0.057 218 −0.221 422 
HeH+ 10−9 −13.419 497 7.216 244 8 0.101 121 71 −0.141 241 61 −1.907 804 
i15c 10−9 −13.419 29 7.216 25 0.101 122 −0.141 242 −1.907 58 
He2+ 10−8 −23.867 58 12.537 54 0.117 596 4 −0.086 577 −3.890 79 
i13c 10−8 −23.866 14 12.537 51 0.117 63 −0.086 58 −3.889 30 
H3 10−8 −2.277 870 1.236 566 0.016 712 −0.046 780 1 −0.329 754 2 
He2 10−6 −24.157 271 12.660 991 0.119 967 0 −0.073 154 −3.965 700 
δEnr18ii2i2rgi,aZaδ(ria)rgi,jδ(rij)rgHOOHBP(2)
H2+ 10−12b −0.624 952 638 6 0.318 293 258 6 ⋯ ⋯ −0.124 978 757 0 
b34c 10−9 −0.624 952 7 0.318 293 2 ⋯ ⋯ −0.124 978 9 
H3+c 10−9 −1.933 449 7 1.089 654 23 0.018 334 666 −0.057 217 520 −0.221 442 3 
i12c 10−9 −1.933 424 1.089 655 0.018 335 −0.057 218 −0.221 422 
HeH+ 10−9 −13.419 497 7.216 244 8 0.101 121 71 −0.141 241 61 −1.907 804 
i15c 10−9 −13.419 29 7.216 25 0.101 122 −0.141 242 −1.907 58 
He2+ 10−8 −23.867 58 12.537 54 0.117 596 4 −0.086 577 −3.890 79 
i13c 10−8 −23.866 14 12.537 51 0.117 63 −0.086 58 −3.889 30 
H3 10−8 −2.277 870 1.236 566 0.016 712 −0.046 780 1 −0.329 754 2 
He2 10−6 −24.157 271 12.660 991 0.119 967 0 −0.073 154 −3.965 700 
a

The electronic state and molecular geometry was H2+, Ã2Σu+: Rp–p = 12.00 bohr; H2, X1Σg+: Rp–p = 1.40 bohr; H3+, 1A1: Rp–p = 1.65 bohr, ϑp1p2p3=60o; H3, 12Σ+: Rp1p2=1.4015 bohr, Rp2p3=5.8113 bohr, ϑp1p2p3=180o; HeH+, 11Σ+: Rp–α = 1.46 bohr; He2+, 12Σu+: Rαα = 2.042 bohr; He2, a(1)3Σu+: Rαα = 2.00 bohr.

b

The non-relativistic energy has been reported with higher precision in Ref. 36.

c

Earlier available computation with regularization type: integral transform (i)6,12, special basis set (b).34 

FIG. 2.

Regularized relativistic corrections to the H2+Ã2Σu+ electronic state near the local minimum (near Rpp = 12.5 bohr) computed by the automated numerical drachmannization approach proposed in this work. The computed values confirm (and are more precise) than the ones reported by Kennedy and Howells using special basis functions.34 

FIG. 2.

Regularized relativistic corrections to the H2+Ã2Σu+ electronic state near the local minimum (near Rpp = 12.5 bohr) computed by the automated numerical drachmannization approach proposed in this work. The computed values confirm (and are more precise) than the ones reported by Kennedy and Howells using special basis functions.34 

Close modal

This work reports a robust and generally applicable numerical drachmannization approach for the computation of the leading-order relativistic corrections with floating explicitly correlated Gaussian basis functions. Formerly incalculable ⟨1/rix1/rjy⟩-type floating ECG integrals are made calculable using a simple and robust Gaussian approximation to one of the 1/r factors. The computational approach is demonstrated to be numerically robust and generally applicable for two-, three-, and four-electron atomic, diatomic, and polyatomic systems and along potential energy curves (and potentially, surfaces). The numerical drachmannization approach developed in this work opens the route to the automated computation of high-precision relativistic corrections to potential energy curves and surfaces in the future.

The supplementary material includes (1) details of the integral calculation and implementation; (2) convergence tables; and (3) basis set parameterizations.

Financial support of the European Research Council through a Starting Grant (No. 851421) is gratefully acknowledged. We thank Péter Jeszenszki and Robbie Ireland for discussions and joint work over the past years.

The authors have no conflicts to disclose.

Balázs Rácsai: Investigation (equal); Methodology (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal). Dávid Ferenc: Investigation (equal); Methodology (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal). Ádám Margócsy: Investigation (equal); Methodology (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal). Edit Mátyus: Investigation (equal); Methodology (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article and its supplementary material.

1.
H. A.
Bethe
and
E. E.
Salpeter
,
Quantum Mechanics of One- and Two-Electron Atoms
(
Springer
,
Berlin
,
1957
).
2.
R. J.
Drachman
, “
A new global operator for two-particle delta functions
,”
J. Phys. B: At. Mol. Phys.
14
,
2733
(
1981
).
3.
G.
Drake
, “
High precision calculations for helium
,” in
Springer Handbook of Atomic, Molecular, and Optical Physics
,
Springer Handbooks
, edited by
G.
Drake
(
Springer
,
New York
,
2006
), pp.
199
219
.
4.
Z.-C.
Yan
and
G. W. F.
Drake
, “
Lithium isotope shifts as a measure of nuclear size
,”
Phys. Rev. A
61
,
022504
(
2000
).
5.
M.
Puchalski
,
D.
Kedziera
, and
K.
Pachucki
, “
Ionization potential for excited S states of the lithium atom
,”
Phys. Rev. A
82
,
062509
(
2010
).
6.
K.
Pachucki
,
W.
Cencek
, and
J.
Komasa
, “
On the acceleration of the convergence of singular operators in Gaussian basis sets
,”
J. Chem. Phys.
122
,
184101
(
2005
).
7.
M.
Puchalski
,
J.
Komasa
, and
K.
Pachucki
, “
Testing quantum electrodynamics in the lowest singlet states of the beryllium atom
,”
Phys. Rev. A
87
,
030502
(
2013
).
8.
M.
Puchalski
,
K.
Pachucki
, and
J.
Komasa
, “
Isotope shift in a beryllium atom
,”
Phys. Rev. A
89
,
012506
(
2014
).
9.
W.
Cencek
,
J.
Komasa
,
K.
Pachucki
, and
K.
Szalewicz
, “
Relativistic correction to the helium dimer interaction energy
,”
Phys. Rev. Lett.
95
,
233004
(
2005
).
10.
M.
Puchalski
,
J.
Komasa
, and
K.
Pachucki
, “
Relativistic corrections for the ground electronic state of molecular hydrogen
,”
Phys. Rev. A
95
,
052506
(
2017
).
11.
P.
Jeszenszki
,
D.
Ferenc
, and
E.
Mátyus
, “
All-order explicitly correlated relativistic computations for atoms and molecules
,”
J. Chem. Phys.
154
,
224110
(
2021
).
12.
P.
Jeszenszki
,
R. T.
Ireland
,
D.
Ferenc
, and
E.
Mátyus
, “
On the inclusion of cusp effects in expectation values with explicitly correlated Gaussians
,”
Int. J. Quantum Chem.
122
,
e26819
(
2022
).
13.
D.
Ferenc
,
V. I.
Korobov
, and
E.
Mátyus
, “
Nonadiabatic, relativistic, and leading-order QED corrections for rovibrational intervals of 4He2+(X2Σu+)
,”
Phys. Rev. Lett.
125
,
213001
(
2020
).
14.
E.
Saly
,
D.
Ferenc
, and
E.
Mátyus
, “
Pre-Born–Oppenheimer energies, leading-order relativistic and QED corrections for electronically excited states of molecular hydrogen
,”
Mol. Phys.
121
,
e2163714
(
2023
).
15.
P.
Jeszenszki
,
D.
Ferenc
, and
E.
Mátyus
, “
Variational Dirac–Coulomb explicitly correlated computations for atoms and molecules
,”
J. Chem. Phys.
156
,
084111
(
2022
).
16.
D.
Ferenc
,
P.
Jeszenszki
, and
E.
Mátyus
, “
On the Breit interaction in an explicitly correlated variational Dirac–Coulomb framework
,”
J. Chem. Phys.
156
,
084110
(
2022
).
17.
D.
Ferenc
,
P.
Jeszenszki
, and
E.
Matyus
, “
Variational vs perturbative relativistic energies for small and light atomic and molecular systems
,”
J. Chem. Phys.
157
,
094113
(
2022
).
18.
P.
Jeszenszki
and
E.
Mátyus
, “
Relativistic two-electron atomic and molecular energies using LS coupling and double groups: Role of the triplet contributions to singlet states
,”
J. Chem. Phys.
158
,
054104
(
2023
).
19.
D.
Ferenc
and
E.
Mátyus
, “
Pre–Born-Oppenheimer Dirac-Coulomb-Breit computations for two-body systems
,”
Phys. Rev. A
107
,
052803
(
2023
).
20.
E.
Mátyus
,
D.
Ferenc
,
P.
Jeszenszki
, and
A.
Margócsy
, “
The Bethe–Salpeter QED wave equation for bound-state computations of atoms and molecules
,”
ACS Phys. Chem. Au
3
,
222
(
2023
).
21.
A.
Margócsy
and
E.
Mátyus
, “
QED corrections to the correlated relativistic energy: One-photon processes
,”
J. Chem. Phys.
(
2024
).
22.
A.
Nonn
,
A.
Margócsy
, and
E.
Mátyus
, “
Bound-state relativistic quantum electrodynamics: A perspective for precision physics with atoms and molecules
J. Chem. Theory Comput.
(in press) (
2024
).
23.
Explicitly Correlated Wave Functions in Chemistry and Physics
, edited by
J.
Rychlewski
(
Kluwer Academic Publishers
,
Dordrecht
,
2003
).
24.
R.
Martinazzo
and
E.
Pollak
, “
Lower bounds to eigenvalues of the Schrödinger equation by solution of a 90-y challenge
,”
Proc. Natl. Acad. Sci. U. S. A.
117
,
16181
(
2020
).
25.
E.
Pollak
and
R.
Martinazzo
, “
Lower bounds for Coulombic systems
,”
J. Chem. Theory Comput.
17
,
1535
(
2021
).
26.
R.
Ireland
,
P.
Jeszenszki
,
E.
Mátyus
,
R.
Martinazzo
,
M.
Ronto
, and
E.
Pollak
, “
Lower bounds for nonrelativistic atomic energies
,”
ACS Phys. Chem. Au
2
,
23
37
(
2021
).
27.
M.
Ronto
,
P.
Jeszenszki
,
E.
Mátyus
, and
E.
Pollak
, “
Lower bounds on par with upper bounds for few-electron atomic energies
,”
Phys. Rev. A
107
,
012204
(
2023
).
28.
L.
Adamowicz
,
S.
Kvaal
,
C.
Lasser
, and
T. B.
Pedersen
, “
Laser-induced dynamic alignment of the HD molecule without the Born-Oppenheimer approximation
,”
J. Chem. Phys.
157
,
144302
(
2022
).
29.
M.
Beutel
,
A.
Ahrens
,
C.
Huang
,
Y.
Suzuki
, and
K.
Varga
, “
Deformed explicitly correlated Gaussians
,”
J. Chem. Phys.
155
,
214103
(
2021
).
30.
G.
Beylkin
and
L.
Monzón
, “
On approximation of functions by exponential sums
,”
Appl. Comput. Harmon. Anal.
19
,
17
(
2005
).
31.
R.
Paunz
,
Spin Eigenfunctions
(
Plenum Press
,
New York
,
1979
).
32.
E.
Mátyus
and
M.
Reiher
, “
Molecular structure calculations: A unified quantum mechanical description of electrons and nuclei using explicitly correlated Gaussian functions and the global vector representation
,”
J. Chem. Phys.
137
,
024104
(
2012
).
33.
R. J.
Harrison
,
G. I.
Fann
,
T.
Yanai
, and
G.
Beylkin
, “
Multiresolution quantum chemistry in multiwavelet bases
,” in
Lecture Notes in Computer Science Vol. 2660
(
Springer
,
Berlin, Heidelberg
,
2003
), pp.
103
110
.
34.
M. H.
Howells
and
R. A.
Kennedy
, “
Relativistic corrections for the ground and first excited states of H+2, HD+ and D+2
,”
J. Chem. Soc., Faraday Trans.
86
,
3495
(
1990
).
35.
I.
Hornyák
,
L.
Adamowicz
, and
S.
Bubin
, “
Ground and excited 1S states of the beryllium atom
,”
Phys. Rev. A
100
,
032504
(
2019
).
36.
M.
Beyer
and
F.
Merkt
, “
Structure and dynamics of H2+ near the dissociation threshold: A combined experimental and computational investigation
,”
J. Mol. Spectrosc.
330
,
147
(
2016
).