A review on non-relativistic, fully numerical electronic

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Received: 4 February 2019 Revised: 19 April 2019 Accepted: 25 April 2019

DOI: 10.1002/qua.25968

REVIEW

A review on non-relativistic, fully numerical electronic


structure calculations on atoms and diatomic molecules

Susi Lehtola

Department of Chemistry, University of


Helsinki, Helsinki, Finland Abstract
The need for accurate calculations on atoms and diatomic molecules is motivated by
Correspondence
Susi Lehtola, Department of Chemistry, the opportunities and challenges of such studies. The most commonly used approach
University of Helsinki, P.O. Box for all-electron electronic structure calculations in general—the linear combination of
55 (A. I. Virtasen aukio 1), FI-00014 Helsinki,
Finland. atomic orbitals (LCAO) method—is discussed in combination with Gaussian, Slater
Email: susi.lehtola@alumni.helsinki.fi a.k.a. exponential, and numerical radial functions. Even though LCAO calculations
Funding information have major benefits, their shortcomings motivate the need for fully numerical
The Academy of Finland (Suomen Akatemia), approaches based on, for example, finite differences, finite elements, or the discrete
Luonnontieteiden ja Tekniikan Tutkimuksen
Toimikunta, Grant/Award Number: 311149 variable representation, which are also briefly introduced. Applications of fully
numerical approaches for general molecules are briefly reviewed, and their challenges
are discussed. It is pointed out that the high level of symmetry present in atoms and
diatomic molecules can be exploited to fashion more efficient fully numerical
approaches for these special cases, after which it is possible to routinely perform all-
electron Hartree-Fock and density functional calculations directly at the basis set
limit on such systems. Applications of fully numerical approaches to calculations on
atoms as well as diatomic molecules are reviewed. Finally, a summary and outlook
is given.

KEYWORDS
atomic calculations, density functional theory, diatomic calculations, fully numerical electronic
structure theory, Hartree-Fock

1 | I N T RO D UC T I O N

Thanks to decades of development in approximate density functionals and numerical algorithms, density functional theory[1,2] (DFT) is now one of
the cornerstones of computational chemistry and solid-state physics.[3–5] Since atoms are the basic building block of molecules, a number of popu-
lar density functionals[6–10] have been parametrized using ab initio data on noble gas atoms; these functionals have been used in turn as the
starting point for the development of other functionals. A comparison of the results of density-functional calculations to accurate ab initio data
on atoms has, however, recently sparked controversy on the accuracy of a number of recently published, commonly used density
functionals.[11–21] This underlines that there is still room for improvement in DFT even for the relatively simple case of atomic studies.
The development and characterization of new density functionals require the ability to perform accurate atomic calculations. Because atoms
are the simplest chemical systems, atomic calculations may seem simple; however, they are in fact sometimes surprisingly challenging. For
instance, a long-standing discrepancy between the theoretical and experimental electron affinity and ionization potential of gold has been
resolved only very recently with high-level ab initio calculations.[22] Atomic calculations can also be used to formulate efficient starting guesses
for molecular electronic structure calculations via either the atomic density[23,24] or the atomic potential as discussed in Ref. [25].

Int J Quantum Chem. 2019;119:e25968. http://q-chem.org © 2019 Wiley Periodicals, Inc. 1 of 31


https://doi.org/10.1002/qua.25968
2 of 31 LEHTOLA

Diatomic molecules, in turn, again despite their apparent simplicity offer a profound richness of chemistry; ranging from the simple covalent
bond in H2 to the sextuple bonds in Cr2, Mo2, and W2;[26] and from the strong ionic bond in Na+Cl− to the weak dispersion interactions in the
noble gases such as Ne  Ar. Even the deceivingly simple Be2 molecule is a challenge to many theoretical methods, due to its richness of static
and dynamic correlation effects.[27] Counterintuitive nonnuclear electron density maxima can also be found in diatomic molecules, both in the
homonuclear and heteronuclear case,[28] again highlighting the surprising richness in the electronic structure of diatomic molecules. Going further
down in the periodic table, diatomic molecules can be used to study relativistic effects to chemical bonding[29,30] as well as the accuracy of various
relativistic approaches.[31–52] Although much of the present discussion applies straightforwardly to relativistic methodologies as well, they will not
be discussed further in this article. Instead, in the present work, I will review basic approaches for modeling the non-relativistic electronic struc-
ture of atoms and diatomic molecules to high accuracy, that is, close to the complete basis set (CBS) limit.
The organization of the article is the following. In Section 2, I will give an overview of commonly used linear combination of atomic orbitals
(LCAO) approaches. I will discuss both the advantages as well as the deficiencies of LCAO methods, the latter of which motivate the need for fully
numerical alternatives, which are then summarized in Section 3. Section 4 discusses applications of fully numerical approaches, starting out with a
summary review on approaches for general molecules in Section 4.1, and continuing with a detailed review on the special approaches for atoms
and diatomic molecules in Sections 4.2 and 4.3, respectively. A summary and outlook is given in Section 5. Atomic units are used throughout
the work.

2 | LCAO APPROACHES

In order to perform an electronic structure calculation, one must first specify the degrees of freedom that are allowed for the one-particle states
a.k.a. orbitals of the many-electron wave function; that is, a basis set must be introduced. An overwhelming majority of the all-electron quantum
chemical calculations that have been, and still are being reported in the literature employ basis sets constructed of atomic orbitals (AOs) through
the LCAO approach.[53,54] As the name suggests, in LCAO the electronic single-particle states, that is, molecular orbitals ψ σi of spin σ are expanded
in terms of AO basis functions as

X
ψ σi ðrÞ = Cσαi χ nα lα mα ðrα Þ, ð1Þ
α

where the index α runs over all shells of all atoms in the system, rα denotes the distance rα = r − Rα, and Rα is the center of the αth basis function
that typically coincides with a nucleus. The matrix Cσ contains the molecular orbital coefficients, its rows and columns corresponding to the basis
function and molecular orbital indices, respectively. The quantities n, l, and m in Equation (1) are the principal, azimuthal, and magnetic quantum
numbers, respectively, of the AO basis functions

χ nlm ðrÞ = Rnl ðr ÞY m ^


l ðrÞ, ð2Þ

^
l ðrÞ are spherical harmonics that are usually chosen
where Rnl(r) are radial functions, choices for which are discussed below in Section 2.2, and Y m
in the real form.
The variational solution of the Hartree-Fock (HF) or Kohn-Sham (KS)[2] equations within the LCAO basis set leads to the Roothaan-Hall or
Pople-Nesbet equations[55–57] for the molecular orbitals

F σ C σ = SC σ Eσ , ð3Þ

where Fσ = Fσ (Cα, Cβ) is the (Kohn-Sham-)Fock matrix, S is the overlap matrix, and Eσ is a diagonal matrix containing the orbital energies. Due to
the interdependence of the Fock matrix and the molecular orbitals, Equation (3) has to be solved self-consistently, leading to the self-consistent
field (SCF) approach. The iterative solution of Equation (3) starts from an initial guess for the orbitals, choices for which I have reviewed recently
in Ref. [25].

2.1 | Benefits of LCAO approaches


LCAO calculations allow all electrons to be explicitly included in the calculation—even if it includes heavy atoms—without significant additional
computational cost, as the necessary number of basis functions scales only in proportion to the number of electrons; the prefactor furthermore
being very small. Moreover, since the AO basis functions tend to look a lot like real AOs, LCAO basis set truncation errors are often systematic,
that is, the error is similar at different geometries or spin states, for example. In such a case, even if the errors themselves are large, the relative
LEHTOLA 3 of 31

error between different geometries or electronic states is usually orders of magnitude smaller, meaning that qualitatively correct estimates can
typically be achieved with just a few basis functions per electron.
Due to the systematic truncation error, LCAO basis sets often allow good compromises between computational speed and accuracy, as basis
sets of various sizes and accuracies are available. Single-ζ basis sets may be used for preliminary studies where a qualitative level of accuracy suf-
fices; for instance, for determining the occupation numbers in different orbital symmetry blocks, or the spin multiplicity of the ground state of a
well-behaved molecule. Polarized double-ζ or triple-ζ basis sets are already useful for quantitative studies within DFT: thanks to the exponential
basis set convergence of SCF approaches and to the fortunate cancellation of errors mentioned above, these basis sets often afford results that
are sufficiently close to the CBS limit. Quadruple-ζ and higher basis sets are typically employed for benchmark DFT calculations, for the calcula-
tion of molecular properties that are more challenging to describe than the relative energy, as well as for post-HF calculations, which require large
basis sets to reproduce reliable results.
Again allowed by the small number of basis functions, post-HF methods such as Møller-Plesset perturbation theory,[58] configuration interac-
tion (CI) theory, and coupled-cluster (CC) theory[59] can often be applied in the full space of molecular orbitals in LCAO calculations. Because of
this, the use of post-HF methods becomes unambiguous, allowing the definition of trustworthy model chemistries[60] and enabling their routine
applications to systems of chemical interest.
Finally, once again facilitated by the small number of basis functions, several numerically robust methods are tractable for the solution of the
SCF equations of Equation (3) within LCAO approaches. As the direct iterative solution of Equation (3) typically leads to oscillatory divergence,
the solution of the SCF problem is traditionally stabilized with techniques such as damping,[61] level shifts,[62–64] or Fock matrix extrapolation[65,66]
in the initial iterations. Once the procedure starts to approach convergence, the solution can be accelerated with, for example, the direct inver-
sion in the iterative subspace (DIIS) method,[67,68] or approached with true[69–78] or approximate[79–82] second-order orbital rotation techniques
that converge even in cases where DIIS methods oscillate between solutions. For more information on recently suggested Fock matrix extrapola-
tion techniques, I refer to Refs. [83–85] and references therein.

2.2 | Variants of LCAO


Three types of LCAO basis sets are commonly used. First, consider the Slater-type orbitals[86–88] (STOs), which are also commonly known as
exponential type orbitals

Rnl ðrÞ / rn−1 expð −ζ nl r Þ: ð4Þ

STOs were especially popular in the early days of quantum chemistry, since they represent analytical solutions to the hydrogenic Schrödinger
equation. The STO atomic wave functions tabulated by Clementi and Roetti[89] should especially be mentioned here due to their wealth of appli-
cations: for instance, their wave functions are still sometimes used for non-self-consistent benchmarks of density functionals.[9,90,91]
STOs have the correct asymptotic form both at the nucleus—the Kato cusp condition[92] can be satisfied—as well as far away, where the
orbitals should exhibit exponential decay.[93–95] However, STOs are tricky for integral evaluation in molecular calculations, typically forcing one to
resort to numerical quadrature.[96–99] Note that even though STO basis sets with non-integer n values[100–103] have been shown to yield signifi-
cantly more accurate absolute energies than analogous basis sets that employ integer values for n,[104–109] they have not become widely used.
Closely related to STOs, Coulomb Sturmians[110,111] have recently gained renewed attention but have not yet become mainstream, see for exam-
ple, Refs. [112] and [113] and references therein.
Second, Gaussian-type orbitals (GTOs)

 
Rnl ðr Þ / rl exp −αnl r2 ð5Þ

enjoy overwhelming popularity, thanks to the Gaussian product theorem[114] that facilitates the computation of matrix elements, which is espe-
cially beneficial for the two-electron integrals,[115–117] the calculation of which is typically the rate determining step in SCF calculations. The
importance of efficient, analytical evaluation of molecular integrals can hardly be overstated. Efficient integral evaluation makes calculations
affordable: it has long been possible to use large atomic basis sets and/or study large systems with Gaussian basis sets. Analytical integral evalua-
tion guarantees accuracy: it is straightforward to calculate, for example, geometric derivatives and reliably estimate force constants and vibrational
frequencies, because the matrix elements are numerically stable.
Another reason for the popularity of GTOs was the recognition early on that STOs can in principle be expanded exactly in GTOs;[118] thus, for
instance, the hydrogenic atom can be solved with GTO expansions with controllable accuracy, highlighting the flexibility of Gaussian basis
sets.[119] Indeed, most commonly used quantum chemistry programs, such as Refs. [120–130] to name a few, employ Gaussian basis sets such as
the Pople 6-31G[131–140] and 6-311G series,[141,142] the Karlsruhe def[143,144] and def2 series,[145,146] the correlation consistent cc-pVXZ series by
4 of 31 LEHTOLA

Dunning, Peterson, and coworkers,[147–158] the polarization consistent pc-n and pcseg-n series by Jensen,[159–166] the nZaPa sets by Petersson
and coworkers,[167,168] as well as the atomic natural orbital (ANO) basis sets by Almlöf, Roos, and coworkers;[169–180] see, for example, Refs.
[181–183] for reviews on Gaussian basis sets.
Third, numerical atomic orbitals (NAOs)

unl ðr Þ
Rnl ðrÞ = ð6Þ
r

are the most recent addition to the field.[184–187] Unlike GTOs and STOs, NAOs have no analytical form; instead, NAOs are typically pretabulated
on a radial grid. (Note that adaptive approaches employing flexible NAOs have also been suggested but have not become commonly used;[188–192]
the atomic radial grid approach should also be mentioned in this context.[193,194]).
The NAO form is by definition superior to GTOs or STOs: a suitably parametrized NAO basis set is able to represent GTOs and STOs exactly,
while the opposite does not hold. Thanks to this great flexibility, unlike GTOs or STOs, a minimal NAO basis set can be an exact solution to the
noninteracting atom at the SCF level of theory: a single NAO can simultaneously satisfy both the Kato cusp condition at the nucleus and have the
correct exponential asymptotic behavior.
In further contrast to primitive, that is, uncontracted STO and GTO basis sets, single-center NAOs can be defined at no additional cost to be
orthonormal, which leads to better-conditioned equations even for the molecular case. These features make NAOs the most accurate LCAO basis
set, as has been recently demonstrated in multiple studies both at the SCF and post-HF levels of theory.[195–198] Alike STOs, NAOs require quad-
rature for molecular integral evaluation, but this does not appear to be a problem on present-day computers.[195,199] NAOs are typically con-
structed from SCF calculations on atoms, which will be reviewed in the next Section 4.2.

2.3 | Case study: Kr atom; contracted basis sets


A comparison of the three types of basis functions for an exchange-only local density calculation on the krypton atom is shown in Figure 1 for the
atomic 1s, 2p, and 3d orbitals. Here, the NAOs are chosen as the exact 1s, 2p, and 3d orbitals, whereas the STO and GTO functions have been
optimized for best overlap with the exact orbital.
The STO form clearly allows an accurate expansion of the 1s orbital. The 1s, however, is a special case, as it experiences the least screening by
the other electrons. These screening effects deform the exact orbitals from the hydrogenic STO form, and become more visible for the 2p and 3d
orbitals: for 3d, the STO orbital is already quite dissimilar from the exact one. But, as was discussed above in Section 2.2, relaxing the restrictions
in the STO form from n = 2 and n = 3 for the 2p and 3d orbitals, respectively, to also allow fractional n leads to significant improvements in the
accuracy of the STO orbitals, as can be seen in Figure 1.
There are large differences between the GTO and the exact orbitals, which are especially clear for the 1s orbital, which should have a cusp at
the origin. However, the accuracy of the GTO basis set can be radically improved by adding more functions. Figure 1 also shows results for con-
tracted GTO (cGTO) basis functions, which have been formed from two GTO primitive functions by optimizing the primitives and their expansion
coefficients for the best overlap with the exact orbital. The cGTO basis set can be made more and more accurate by adding more and more primi-
tives to the expansion.

(A) 450 (B) 30 (C) 4.5


NAO (exact) NAO (exact) NAO (exact)
400 STO STO 4.0 STO
25
350 GTO optSTO 3.5 optSTO
cGTO GTO GTO
20
300 cGTO 3.0 cGTO
250 2.5
15
R (r)

R (r)

R (r)

200 2.0
10
150 1.5

100 5 1.0

50 0.5
0
0 0.0

-50 -5 -0.5
0.0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
r/a 0 r/a 0 r/a 0

F I G U R E 1 Comparison of NAOs, STOs, GTO, and contracted GTOs (cGTOs) for the 1s, 2p, and 3d radial functions of krypton. STOs with an
optimized fractional value for n are also shown (optSTO)
LEHTOLA 5 of 31

However, it is worth pointing out here that in a sense, cGTOs are also a form of a NAO basis. Employing general contractions,[200] the single-
center basis functions can be chosen to be strictly orthonormal, and noninteracting atoms can be solved to high accuracy in a minimal basis set
that is assembled from a large number of primitive Gaussians. Alike NAOs, efficient integral evaluation in generally contracted GTO (gcGTO) basis
sets is significantly more complicated than in primitive or segmented Gaussian basis sets, since a primitive integral can contribute to several con-
tracted integrals.
gcGTO basis sets are typically formed from ANOs, and they are well-known to yield faster convergence to the basis set limit with respect to
the number of basis functions than uncontracted basis sets or basis sets employing segmented contractions.[169–180,201–203] Even though gcGTOs
are similar to NAOs, the number of degrees of freedom in any Gaussian basis set expansion is much smaller than what is possible with a true
NAO basis set: large Gaussian basis sets for heavy atoms typically contain up to 30 or 40 primitives per angular momentum shell, whereas NAOs
are often tabulated with hundreds to thousands of radial points. As NAOs allow for a more accurate description of AOs than Gaussian expansions
do, while ANOs yield faster convergence in post-HF calculations than basis sets constructed from HF orbitals do, it is clear that it should be possi-
ble to form extremely efficient numerical ANO (NANO) basis sets for post-HF studies. The generation of NANO basis sets could follow pre-
established procedures to generate ANOs from Gaussian expansions; suitable fully numerical implementations of the multiconfigurational
(MC) approaches that are typically used to construct ANOs have been described in the literature, as is discussed below in Section 4.2.

2.4 | Deficiencies of LCAO approaches


In the case of atomic calculations, it is relatively straightforward to obtain near-CBS limit energies even with LCAO basis sets: since only a single
expansion center is involved, the basis set is relatively well conditioned even if large basis set expansions are used. However, the basis set require-
ments depend not only on the level of theory and property; they may also depend on the studied state.
For some properties, it is possible to explicitly identify the regions of the basis set that require extra focus: for instance, the Fermi contact
term for a nucleus A is proportional to δ(rA), where rA is the distance from nucleus, so an accurate reproduction of the electronic structure at the
nucleus is necessary. Because only s orbitals are nonzero at the nucleus, the accuracy of the description of the contact term is mostly dependent
on the tight s functions.
Other examples include the polarization induced by weak electric and/or magnetic fields, for which analytic expressions can be derived for
the perturbative change in the AOs. Equipped with this knowledge, one can parametrize perturbation-including basis sets both in the case of
STOs[204–206] and GTOs,[205,207–220] which achieve faster convergence to the basis set limit than the non-perturbation-including basis sets. Simi-
larly, knowledge on the asymptotic long-range behavior of the wave function[94,95] can be used to construct GTO basis sets with improved asymp-
totic behavior.[221]
Completeness-optimization[222,223] is a general approach that may be used to parametrize basis sets that yield near-CBS values for the energy
and/or whatever molecular properties. The approach does not make any assumptions on the level of theory or specific properties that are calcu-
lated; instead, the only assumption is that the property can be converged with a sufficiently large atomic basis set. The approach relies on an itera-
tive approach where new basis functions are added one by one, until the property has converged. Completeness-optimization has been
demonstrated so far for electric and magnetic properties,[222–234] electron momentum densities,[235,236] as well as NAO simulations of
nanoplasmonics.[237]
However, getting to the CBS limit can still be painstaking, as several sets of tight and/or diffuse functions may be necessary to achieve con-
verged results.[150,238–241] For instance, Gaussian primitives with exponents as small as 6.9 × 10−9 have been found necessary to make F− prop-
erly bound at the DFT level.[242] Basis sets with a fixed analytic form also have intrinsic limits to their accuracy because of their incorrect
asymptotic behavior. This has been shown in the case of Gaussian basis sets, for example, for the moments of the electron momentum
density,[235,236,243] the electron density,[244–247] as well as nuclear magnetic properties.[248] Reaching the CBS limit for such properties may be
untractable unless more flexible basis sets are employed.
Molecular calculations are even more difficult than atomic ones. Although it is straightforward to establish a hierarchy between NAO, STO,
and GTO basis sets in the case of atoms and optimize basis sets for them, atomic symmetries are broken in a molecular environment, requiring
the addition of polarization functions into the atomic basis set. It is generally acknowledged that the optimal form of polarization functions is not
known, and that the ranking of GTOs vs. STOs vs. NAOs is not clear in the case of polarization functions. In fact, NAO calculations typically
employ polarization functions of GTO or STO form.[195,249]
Another complication is that the set of AO basis functions placed on the various nuclei may develop linear dependencies. Linear dependencies
are especially problematic for diffuse basis functions, whereas tight functions are better behaved since they are highly localized. Thus, it is espe-
cially difficult to accurately reproduce molecular properties that are sensitive to the diffuse part of the wave function with LCAO approaches; fur-
ther complications may also be caused by the limitations in the form of the AO basis functions discussed above.[235,236,243–248]
Any linear dependencies in the basis set are typically removed via the canonical orthonormalization procedure[250] in order to make the basis
unambiguous. In such a case, a calculation with a smaller basis set can reproduce the same or a lower energy than the one predicted by the calcu-
lation with the linearly dependent basis.
6 of 31 LEHTOLA

There are workarounds to the problem of significant linear dependencies for some cases. Instead of using a large basis set on each of the
nuclei, one can employ bond functions: use a smaller basis on the nuclei while placing new basis functions in the middle of chemical bonds; such a
procedure can yield faster convergence to the basis set limit.[251] Moreover, one can replace diffuse functions placed on every atom with ones
placed only on the molecular center of symmetry. However, in either case, the optimal division and placement of functions to recover the wanted
property yet keep linear dependencies at bay may be highly nontrivial for a general system, and is clearly dependent on the molecular geometry.
A general approach for avoiding the problems with linear dependencies that arise from diffuse functions in a molecular setting has, however,
been suggested. In this approach, diffuse functions are modeled by a large number of non-diffuse functions placed all around the system.[252,253]
The use of such lattices of Gaussian functions is an old idea, originally proposed as the “Gaussian cell model” by Murrell and coworkers.[254] The
Gaussian cell model was later revisited by Wilson and coworkers, who showed that much better results can be obtained by supplementing the lat-
tice basis set with AO functions on the nuclei.[255–257] The authors of Refs. 253 and 254 appear to have been unaware of the work by Murrell,
Wilson, and coworkers. Nonetheless, the Gaussian cell model is but an imitation of proper real-space methods: the use of a grid of spherically
symmetric Gaussians clearly cannot yield the same amount of accuracy or convenience as a true real-space approach that employs a set of sys-
tematically convergent basis functions with finite support. In one dimension, however, arrays of Gaussian functions called gausslets have been
recently shown to yield useful completeness properties.[258] But, it is not yet clear how well the gausslet approach can be generalized to atoms
heavier than hydrogen, or to general three-dimensional molecular structures.
The accurate calculation of, for example, potential energy surfaces VAB = E(A  B) − E(A) − E(B) is also complicated in the LCAO approach due
to the so-called basis set superposition error (BSSE). E(A  B) can be underestimated with respect to E(A) and E(B), because the locality of the elec-
trons onto the basis functions belonging to the fragments is not enforced. Although BSSE can be approximately removed with the Jansen-Ros-
Boys-Bernardi counterpoise correction[259,260] or developments thereof,[261–266] and although BSSE becomes smaller when going to larger atomic
basis sets, it is not always clear how accurate the obtained interaction energies are. The geometry-dependence of the molecular basis set also
complicates the calculation of molecular geometries and vibrational frequencies, due to the introduction of Pulay terms.[267]
Finally, it is difficult to improve the description of the wave function in an isolated region of space in the LCAO approach. The addition of a
new function to describe, for example, the wave function close to the nucleus requires changing the solution globally to maintain orbital
orthonormality, because the basis functions are non-orthogonal and have global support. Could these problems be circumvented somehow?

3 | REAL-SPACE METHODS

An alternative to the use of LCAO basis sets is to switch to real-space methods, where the basis functions no longer have a form motivated purely
by chemistry. Instead of expanding the orbitals in terms of AOs as in Equation (1), the expansion is generalized to

X σ
ψ σi ðrÞ = CIi ΞI ðr Þ, ð7Þ
I

where the real-space basis set ΞI(r) can be chosen based on convenience for the problem at hand. For illustration, a simple option is to form the
three-dimensional basis set ΞI(r) as a tensor product of one-dimensional basis functions χ i(r) as

ΞI ðr Þ = χ Ix ðxÞχ Iy ðyÞχ Iz ðzÞ, ð8Þ

where the χ i(r) can in turn be chosen as, for example, the hat functions shown in Figure 2. In many dimensions, also other kinds of choices for the
three-dimensional basis functions can be made instead of the rectangular elements of Equation (8): for instance, triangles in two dimensions, or
tetrahedra, hexahedra, or prisms in three dimensions; more information on these can be found in any standard finite element textbook such as
Ref. [268].
The present work focuses on fully numerical calculations on atoms and diatomic molecules. As will be reviewed below in Section 4, a one-
dimensional treatment is sufficient for atoms, whereas diatomic molecules require a two-dimensional solution. Almost all the diatomic approaches
reviewed below use a rectangular basis similar to Equation (8):

ΞI ðξ, ηÞ = χ Iξ ðξÞχ Iη ðηÞ, ð9Þ

where ξ and η are prolate spheroidal coordinates introduced below in Section 4.3. The only exception is the program by Heinemann, Fricke and
Kolb, which employs triangular elements as discussed in Ref. [269]; the present discussion can thus be restricted to the case of Equation (8) or
Equation (9). Now, as the basis factorizes in the coordinates, it is sufficient to restrict the discussion to one dimension, as the generalization to
many dimensions is obvious.
LEHTOLA 7 of 31

F I G U R E 2 Illustration of hat functions, which are nonzero only within a 1.0


finite interval
0.9

0.8

0.7

0.6

χ (x )
0.5

0.4

0.3

0.2

0.1

0.0
x0 x1 x2 x3 x4
x

Four kinds of numerical approaches are commonly used for the one-dimensional basis functions χ I(x): finite differences, finite elements, basis
splines (B-splines), and the discrete variable representation (DVR). The common feature in all the approaches is that a function f(x) is represented
in terms of the basis functions χ I(x) as
X
f ðxÞ ≈ cI χ I ðxÞ: ð10Þ
I

As will be seen below, the first three methods rely on a local piecewise polynomial representation for f(x), whereas the last one can also be for-
mulated in terms of non-polynomial basis functions χ I(x). All four approaches should yield the same value at the CBS limit—this is what makes fully
numerical approaches enticing in the first place. However, as will be discussed below, the speed of convergence to the CBS limit depends heavily
on the order of the used numerical approximation, with higher-order methods affording faster convergence in the number of basis functions than
lower-order methods. This means that comparing the accuracy of the various approaches is not trivial, as even the same approach can manifest
different convergence properties when implemented at different orders.
Furthermore, the speed of convergence also depends in each case on the system, the studied property, and the level of theory used in the cal-
culation: HF and DFT afford exponential converge in the basis set, whereas post-HF methods only exhibit polynomial convergence if the orbital
space is not fixed. For this reason, for numerical results I refer the reader to the individual references listed in Section 4.

3.1 | Finite differences


The finite difference method is arguably the oldest and the best-known method for numerically solving differential equations. The method is
based on a Taylor expansion of f(x) as

1  
f ðx  hÞ = f ðxÞ  f 0 ðxÞh + f 00 ðxÞh2 + O h3 , ð11Þ
2

where it is assumed that f(x) is tabulated on a grid xi = x0 + ih as fi = f(xi), where h is the grid spacing; that is, f(xi + h) = fi + 1. From the Taylor expan-
sion, Equation (11), one can obtain approximations for the derivatives of f(x), for instance, using central differences as

f ðx + hÞ− f ðx− hÞ  
f 0 ðxÞ≈ + O h2 , ð12Þ
2h

f ðx + hÞ −2f ðxÞ + f ðx − hÞ  
f 00 ðxÞ ≈ + O h2 ; ð13Þ
h2

these are known as the two-point and three-point stencils, respectively. The procedure can also be done at higher orders, leading to, for example,
the four-point and five-point rules
8 of 31 LEHTOLA

− f ðx + 2hÞ + 8f ðx + hÞ− 8f ðx − hÞ + f ðx− 2hÞ  


f 0 ðxÞ ≈ + O h4 , ð14Þ
12h

− f ðx + 2hÞ + 16f ðx + hÞ −30f ðxÞ + 16f ðx− hÞ −f ðx − 2hÞ  


f 00 ðxÞ ≈ + O h4 , ð15Þ
12h2

which carry a smaller error, as evidenced by the smaller remainder term. Finite difference formulas can be easily generated to arbitrary order by
recursion, even on nonuniform grids.[270]
An important feature of the finite difference method is that relaxation approaches can be formulated for Schrödinger-type equations. For
instance, applying Equation (12) onto

d2 ψ
− ðxÞ + V ðxÞψ ðxÞ = Eψ ðxÞ ð16Þ
dx2

one obtains

ψ i + 1 − 2ψ i + ψ i−1
− + V i ψ i = Eψ i ð17Þ
h2

which can be rearranged into the relaxation equation

ψ i− 1 + ψ i + 1 1
ψi = + ðVi − EÞψ i h2 ð18Þ
2 2

ð0Þ ð1Þ ðkÞ


that can be solved by iteration. Furthermore, the iteration ψ i ! ψi ! … ! ψ i ! … can be made faster by the use of the successive over-
relaxation method

ðkÞ ðkÞ ðk −1Þ


ψ i = ωψ i + ð1− ωÞψ i ð19Þ

where 0 < ω ≤ 1 corresponds to relaxation and ω > 1 to overrelaxation.


Although the relaxation approach is straightforward and readily applicable to high-performance computing, the drawback is that a good initial
guess is necessary for the solution ψ i as well as for the eigenvalue E. Another drawback is that, as can be seen from Equation (18), information is
passed only locally, and so the relaxation becomes extremely slow in fine meshes; this can be, however, circumvented with multigrid
approaches[271] that pass information simultaneously at multiple length scales, thus avoiding the slowdown.
It is also possible to avoid using relaxation approaches altogether: finite difference calculations can be written as matrix eigenvalue problems.
For example, Equation (17) can be rewritten as

Hψ = Eψ ð20Þ

and the finite-difference Hamiltonian is given by the banded-diagonal matrix

0 1
.. ..
B . . C
B. C
B . −2 C
B . −h 2h −2 + V i −1 − h −2 C
B C
H=B
B −h − 2
2h −2
+ V i − h −2 C,
C ð21Þ
B . C
B −2 −2 −2 . C
B − h 2h + V i+1 − h . C
@ A
.. ..
. .

where ψ and V are vectors of the values of the wave function and the potential, respectively; however, this is rarely done in practice, as the
B-spline and finite element methods (FEM) described below naturally lead to matrix equations.
Having obtained approximate derivatives of the function in the vicinity of the current point using, for example, Equations (12) and (13) or
Equations (14) and (15), Equation (11) expands the description of the function beyond the known values fi to how the function behaves in-between the
LEHTOLA 9 of 31

grid points. From the form of Equation (11), it is clear that the finite difference method in fact corresponds to a local, piecewise polynomial basis for f(x).
However, the Taylor rule of Equation (11) implies that a different polynomial is used for every grid point xi, which leads to a discontinuous representa-
tion for f(x). This is the puzzle of the finite difference method: it is not obvious how the approximated function behaves in-between grid points.
Applications to quantum mechanics require matrix elements that involve spatial integrals; thus, a quadrature rule is necessary, but due to the
aforementioned ambiguity, it is not possible to derive one from the finite difference formulation itself. Instead, the typical solution is to just use
quadrature formulas such as Newton-Cotes, which can be derived from Lagrange interpolating polynomials (LIPs), yielding, for example, the two-
point trapezoid rule, the three-point Simpson's rule, and the five-point Boole's rule. Although these quadrature formulas imply using the same
fitting polynomial for all points {xi} in the fitting integral—contrasted to using different polynomials at every point to get the derivatives according
to Equations (12)-(15)—there is a solid mathematical argument for using such a hybrid scheme: the error bounds. For instance, the trapezoid and
       
Simpson's rules have errors of O h3 and O h5 , respectively, which can be contrasted with the O h2 and O h4 errors for the derivatives in

Equations (12) and (13) or Equations (14) and (15), respectively. Because of this, the mismatch between the rules for derivation and integration is
not a problem in practice: numerical integration is more accurate than numerical differentiation, and an arbitrary level of accuracy is anyhow
achievable by using a finer grid h ! 0.
Still, getting to the CBS limit is complicated by the lack of variationality in finite difference calculations:[272] when the variational principle does
not hold, calculations with a smaller h can yield an increased energy, even though a better basis set is used. This error is caused by the underesti-
mation of the kinetic energy by the finite difference representation of the Laplace operator.[273,274]
But, finite difference electronic structure calculations often have also another, even larger source of non-variationality than the kinetic energy
approximation: aliasing error in the Coulomb and exchange potentials. This error arises whenever the electron density is expressed in the same
grid as the molecular orbitals

X X
ρn = ρðrn Þ = jψ i ðr n Þj2 = jψ ni j2 ð22Þ
i occupied i occupied

because an order-k polynomial wave function yields a density that has an order-2k polynomial expansion, which ends up being truncated. The
aliasing error is analogous to the error of resolution-of-the-identity approaches of LCAO methods:[275–279] it makes the classical Coulomb
interaction between the electrons less repulsive, and the exchange interaction less attractive. The latter error yields a slight increase of the
energy, but the former error makes the energy too negative, resulting in energies that are underestimated, that is, anti-variational until the
grid is made fine enough.
One more limitation in the finite difference method is that if an equispaced grid is used, higher-order rules, which are based on high-order
polynomials, become unstable due to Runge's phenomenon.[280] For this reason, for instance the X2DHF program[281–283] discussed below in
 
Section 4.3 employs eight-order, that is, 9-point central difference expressions for computing second derivatives with O h8 error, and a 7-point
 9
rule for quadrature with O h error.[282]

3.2 | Explicit basis set methods


Instead of invoking a local piecewise polynomial basis only implicitly as in the finite difference method, the FEM explicitly uses a piecewise polyno-
mial basis set to represent the function f(x) like in interpolation theory. Because of this, it can be remarked that finite differences approximate the
Schrödinger equation, while finite elements—as well as B-splines and the DVR—approximate its solution. The nice thing about these three
approaches is that the basis is fully defined: evaluating the function f(x) or its derivatives f(n)(x) at an arbitrary point x is completely unambiguous,
meaning, for example, that adaptive quadrature rules, or rules relying on points other than the discretization nodes can be easily employed. More-
over, because an unambiguous basis set has now been defined, instead of relying on an (over)relaxation approach as in the finite difference method,
one may use the same variational techniques as in LCAO approaches to solve for the wave functions, for example, the Roothaan–Hall equation of
Equation (3).
In the FEM, the domain for x is divided into line segments called elements x 2 [xa, xb], and a separate, local polynomial basis is defined within
each element. The basis functions within an element, commonly known also as shape functions, are determined by choosing n points xi called nodes
within the element, xi 2 [xa, xb].
In the most commonly used finite element variant, one demands that the value of f(x) at each node is given by a specific basis function

X
f ðxi Þ = cj χ j ðxi Þ = f i : ð23Þ
j
10 of 31 LEHTOLA

Employing a polynomial expansion for χ i, one sees that the shape functions for n nodes are n − 1 order polynomials. Especially, the condition
χ j(xi) = δij in Equation (23) is fulfilled by LIPs

nY
−1
x− xj
χ i ðxÞ = , ð24Þ
x
j = 0, j6¼i i
− xj

which are well-known to yield the error estimate

  Qn−1 ðx − x Þ  
   
f ðxÞ− ~f ðxÞ ≤ i = 0 max ξ2½xa , xb  f ðnÞ ðξÞ:
i
ð25Þ
n!

Nodes are typically placed also at the element boundaries. Because the node at the element boundary is shared by the rightmost function in
the element on the left, and the leftmost function in the element on the right, these two basis functions have to be identified with each other,
guaranteeing continuity of f(x) across the element boundary. For instance, the hat functions in Figure 2 correspond to two-node elements, where
both nodes are on element boundaries.
Because all shape functions are fully localized within a single element, it is very easy to take advantage of this locality in computer
implementations. Moreover, the correspondence of functions belonging to nodes at element boundaries can be easily handled in computer
implementations by index manipulations, that is, overlaying.
In this formulation, because all basis functions are smooth, f(x) will also be smooth within each element, and the function f(x) is guaranteed to
be continuous over element boundaries. However, while all derivatives are guaranteed to be continuous within a single element, discontinuities in
the first derivative f0 (x) may be seen at the element boundaries.
Although the continuity of the derivative(s) can be strictly enforced by straightforward extension of the approach in Equation (23), in which
the description of the values of f(x) and f0 (x) at the nodes are each identified to arise from a single shape function, this does not appear to be nec-
essary in typical applications to quantum mechanics. The reason for this is the variational principle: discontinuities in the derivative imply a higher
kinetic energy. The derivatives of the LIPs, Equation (24), do not vanish at the nodes, and this flexibility gives the variational principle room to try
to match the derivatives across element boundaries.
Although the Runge phenomenon[280] is also a problem for the FEM with uniformly spaced nodes within the element, the method can be
made numerically stable to high orders if one employs nonuniformly spaced nodes; suitable choices include Gauss-Lobatto nodes, or Chebyshev
nodes as in the spectral element method.[284] The benefit of high-order elements is apparent in Equation (25): the errors rapidly become small in
increasing element order.[285] If the same quadrature rule is used both to define the shape functions, and for calculating integrals, one obtains a
DVR which will be discussed in more detail below.
In addition to varying the location of the nodes within each element to make high-order rules numerically stable, it is possible to use elements
of different sizes and orders within a single calculation, granting optimal convergence to the basis set limit with the least possible number of basis
functions; this is the origin for the popularity of FEM across disciplines. For more discussion on the FEM, including other kinds of shape functions
than LIPs, I refer to my recent discussion in Ref. [285].
Like the FEM, B-splines are piecewise polynomials, with an order n spline corresponding to a polynomial function of order n. B-splines were
originally developed by Schoenberg,[286] and were introduced in the context of atomic calculations by Gilbert and Bentoncini,[287–290] based on
earlier work on their application to single-particle problems by Shore.[291–295] B-splines are defined by knots x0 ≤ x1 ≤ xn + 1, which need not be
uniformly spaced, with the Cox–de Boor recursion


1 xi ≤ x < xi + 1
Bi, 0 = , ð26Þ
0 otherwise

x − xi xi + n + 1 − x
Bi, n ðxÞ = Bi, n −1 ðxÞ + Bi + 1, n−1 ðxÞ; ð27Þ
xi + n − xi xi + n + 1 − xi + 1

thus, Bi,0 are piecewise constants, Bi,1 are hat functions (Figure 2), and so on. On each interval (xi, xi + 1), only n+1 B-splines are nonzero. This
means that computational implementations of B-splines and finite elements share similar aspects: matrix elements are computed in batches on
the individual knot intervals; only the indexing of the basis functions is different. B-splines and their applications to atomic and molecular physics
have been reviewed by Bachau in Ref. [296], to which I refer for further details.
The DVR,[297–303] in turn, is conventionally based on orthogonal polynomial bases and their corresponding quadrature rules,[299,300] under-
lining a strong connection to the finite element and B-spline approaches. Generalizations of DVR to non-polynomial bases have been suggested
by Szalay[304–307] and Littlejohn, Cargo, and coworkers.[308–313] The DVR community discerns between a variational basis representation, in which
LEHTOLA 11 of 31

all operator matrix elements, or integrals, are computed exactly, that is, analytically, a finite basis representation (FBR), in which a quadrature rule
is employed to approximate the integrals as in FEM, and the DVR itself in which a further unitary transformation of the FBR basis set, which is
typically delocal, is used to obtain an effectively localized basis which affords a diagonal approximation for matrix elements as

ðb
Oij = χ *i ðxÞOðxÞχ j ðxÞdx≈Oðxi Þδij ; ð28Þ
a

this is the famous “discrete δ property” of DVR. Although the DVR approach offers few benefits for one-dimensional problems over the full, varia-
tional treatment, since cost of the additional transformation has similar scaling to the savings, the approach becomes extremely powerful in many
dimensions due to the diagonality of potential operators, Equation (28).
The DVR is commonly used on top of a finite element treatment, yielding the FEM-DVR method, as discussed, for example, by Rescigno and
McCurdy.[314] Indeed, the DVR approaches discussed in the rest of the article all employ Lobatto elements, which use a basis of LIPs defined by a
set of Gauss-Lobatto quadrature nodes, as was discussed above in relation to FEM. It is easy to see that if the same Gauss-Lobatto quadrature
rule is used both to define the finite element basis and for evaluating matrix elements, then the property of the interpolating polynomials, Equa-
tion (23), directly leads to the diagonal approximation of matrix elements, Equation (28).
However, this is an approximation: reliable computation of the matrix elements Oij in Equation (28) may require using more than N quadrature
points, which in turn may introduce off-diagonal elements in the matrix representation of O. Because of this quadrature error, which is analogous
to the aliasing error discussed above in the case of finite-difference calculations, DVR calculations can be nonvariational; however, the error can
be made smaller and smaller by going to larger and larger DVR basis sets.

4 | APPLICATIONS

4.1 | Overview on general approaches for arbitrary molecules


With the ability to use varying levels of accuracy for different parts of the system in the real-space basis set, even challenging parts of the wave
function, such as at the nucleus or far away from it, can be straightforwardly converged to arbitrary accuracy. Real-space methods thereby grant
forthright access even to properties that are difficult to calculate accurately with, for example, Gaussian basis sets, including nuclear magnetic
properties, the electron density close to, or at the nucleus, as well as the electron momentum density. A further benefit of real-space methods is
that the locality of the real-space basis set allows for simple and efficient algorithms even on massively parallel architectures, whereas LCAO cal-
culations are much less trivial to parallelize to thousands of processes.
Achieving the CBS limit is simple in real-space methods. Due to the local support of the basis functions, the basis set never develops linear
dependencies, so the CBS limit can be achieved simply by increasing the accuracy of the real-space basis set until the result does not change any-
more. The drawback to the arbitrary accuracy is that the number of basis functions necessary for achieving even a qualitative level of accuracy in
real-space approaches is significantly larger than in the LCAO approach; the crossover typically happens only at high precision.
The number of basis functions in real-space methods is especially large in all-electron calculations, as disparate length scales are introduced by
the core orbitals. The description of the tightly bound core orbitals necessitates an extremely robust basis near the nuclei, whereas a much coarser
basis is sufficient for the valence electrons. One option to circumvent this problem is to forgo all-electron calculations and to approximate the
core electrons with pseudopotentials[315–321] or projector-augmented waves[322] (PAWs), making the problem accessible to a number of fully
numerical approaches; see, for example, Refs. [272,323–326]. However, as the present review focuses on all-electron calculations, the use of
PAWs and pseudopotentials will not be considered in the rest of the article.
In order to make all-electron a.k.a. full-potential calculations feasible, a suitable representation has to be picked for the description of the core
electrons. As was stated above, LCAO calculations have problems especially with diffuse parts of the wave function, whereas core orbitals do not
pose significant complications. Vice versa, real-space methods tend to struggle with core orbitals, but are well adapted to the description of dif-
fuse character. This has lead to approaches that hybridize aspects of LCAO and real-space calculations.[191,193,327–339] By representing the core
electrons using an atom-centric radial description, a significantly smaller real-space basis set may suffice, making the approaches scalable to large
systems while still maintaining a very high level of accuracy. In a somewhat similar spirit, the multidomain finite element muffin-tin and full-
potential linearized augmented plane-wave + local-orbital approaches have also been recently shown to afford high-accuracy all-electron calcula-
tions for molecules.[340,341]
However, several methods that are able to solve the all-electron problem without using an LCAO component exist as well. First, the FEM
allows for calculations at arbitrary precision even at complicated molecular geometries by employing spatially nonuniform basis functions. A num-
ber of FEM implementations for all-electron HF or DFT calculations on molecules or solids with arbitrary geometries have been reported in
the literature.[342–360] Another solution is to use grids with hierarchical precision[361,362] or adaptive multiresolution multiwavelet
approaches[198,248,323,363–367] that are under active development and which have recently achieved microhartree-accuracy in molecular
12 of 31 LEHTOLA

calculations, even at the post-HF level of theory.[368–371] Regularization of the nuclear cusp[372–374] and approaches employing adaptive coordi-
nate systems[375] also have been suggested.
The considerably larger number of basis functions in real-space approaches compared to LCAO calculations means not only that more compu-
tational resources—especially memory and disk space—are needed, but also that alternative approaches for, for example, the solution of the SCF
equations have to be used. For instance, employing 100 degrees of freedom such as grid points per Cartesian axis yields (100)3 = 106 basis func-
tions in total. The corresponding Fock matrix would formally have a dimension 106 × 106, requiring a whopping 8 TB of memory (assuming double
precision), and making its exact diagonalization untractable. Naturally, these challenges become even worse for molecules for which orders of
magnitude more points are necessary.
Furthermore, the calculation of properties may be complicated in real-space methods. If the numerical basis set is independent of geometry,
there is no BSSE as in LCAO calculations. Instead, one has to deal with the infamous “egg-box effect,”[376] which is caused by the use of a finite
spatial discretization that breaks the translational and rotational symmetries of the system. Then, the numerical accuracy depends on the relative
position of the nuclei to the discretization, which can manifest as, for example, spurious forces.[376] The egg-box effect can be made small by
going to larger and larger basis sets; other alternatives have been discussed in Ref. [377], but the best approach may depend on the form of the
numerical basis set.
Typical real-space approaches employ local basis functions such as the ones in Figure 2, making most matrices extremely sparse. Large-scale
computational implementations are possible only via the exploitation of this sparsity. For instance, the Coulomb interaction as well as local and
“semilocal” density functionals require only local information, namely, the Coulomb potential VC(r) at every grid point, as well as the exchange-
correlation potential VXC(r) that is built from the density (and possibly its gradient) at every grid point. In contrast, the exact HF exchange interac-
tion is nonlocal; however, it can be truncated to finite range in large systems, especially if exact exchange is only included within a short range, as
in the popular Heyd-Scuseria-Ernzerhof functional.[378,379] Alternatively, the evaluation of exact exchange can be sped up with projection opera-
tors;[380,381] or the Krieger-Li-Iafrate (KLI) approximation[382,383] can be used to build a fully local exchange potential, also allowing exact-exchange
calculations on large systems.[384–387] Further necessary steps in the implementation of large-scale real-space approaches involve replacing the
full matrix diagonalization in the conventional Roothaan orbital update of Equation (3) with an iterative approach such as the Davidson
method[388] that is used to solve only for the new occupied subspace, or avoiding diagonalization altogether by using, for example, Green's func-
tion methods to solve for the new occupied orbitals, as in the Helmholtz kernel method.[334,389–391]
So far, the discussion has been completely general. Next, I shall discuss calculations on atoms and diatomic molecules. Although it is possible
to obtain accurate results for atoms or diatomic molecules by employing general three-dimensional real-space approaches as the ones discussed
above—see, for example, Refs. [392] and [393] for tetrahedral-element calculations on H2+, and H− and He, respectively—they carry a higher com-
putational cost. A specialized implementation is able to treat some dimensions analytically or semianalytically, whereas these dimensions need to
be fully described in the general three-dimensional programs, necessitating orders of magnitude more degrees of freedom to achieve the same
level of accuracy as is possible in the specialized algorithms.
For instance, while a 1003 grid can be assumed to yield a decent HF or DFT energy for any atom, an atomic solver will likely give a much bet-
ter energy with just 1000 basis functions, as the angular expansion for s to f electrons requires but 16 spherical harmonics, while 60 radial func-
tions afford sub-microhartree accuracy even for xenon.[285] Furthermore, the smaller basis set afforded by the use of the symmetry inherent in
the problem allows (di)atomic calculations to use a diversity of powerful approaches—such as the use of Equation (3) with full diagonalization for
the orbital update—which are not tractable in the general three-dimensional approach.[285,394]

4.2 | Atoms
At the non-relativistic level of theory, the electronic Hamiltonian for an atom with nuclear charge Z

X XZ X 1
^el = − 1
H r2i − + ð29Þ
2 i i
r i i > j rij

^ As a result, electronic states are


commutes with the total angular momentum operator L^ , its z component L^z , as well as the total spin operator S.
2

identified with the term symbol 2S + 1L, where the total electronic angular momentum L is one of S, P, D, F, G…. In cases where the term symbol is
not nS, the wave function is in principle strongly correlated due to the degeneracy of partially filled shells. Unlike the molecular case, in which the
treatment of strong correlation effects is still under active research, for atoms these effects are much more docile due to the limited number of
accessible orbitals. For example, in the ground state configuration of boron the 2p shell only has a single electron, which should be distributed
equally between the px, py, and pz orbitals, yielding a MC wave function. A good symmetry-compatible ansatz for the wave function can be con-
structed with Russell-Saunders LS-coupling,[395] yielding, for example, the MC-HF method. Such calculations can also be formulated as an average
over the energies of the individual configurations.[396–400]
LEHTOLA 13 of 31

Fully numerical calculations are possible for atoms, because the two-electron interaction has a compact, factorizable representation in the
polar spherical coordinate system, as shown by the Legendre expansion


1 4π X ∞
1 r< L X
L  M *
= YM
L ðΩ1 Þ Y L ðΩ2 Þ : ð30Þ
r12 r > L = 0 2L + 1 r > M = −L

The numerical representation of the AOs was already given in Equation (6), which yields the boundary value unl(0) = 0 for all values of l. It is
seen that the two-electron integrals

ðð
χ i ðr Þχ *j ðr Þχ k ðr0 Þχ *l ðr0 Þ
ðijjklÞ = d3 rd3 r0 ð31Þ
jr − r 0 j

reduce to an angular part times a radial part in such a basis set. The angular part can be evaluated analytically using the algebra of spherical har-
monics, whereas the radial parts involve one-dimensional and two-dimensional integrals, which can be readily evaluated by quadrature.
Indeed, real-space calculations on atoms in the (MC-)HF approach have a long history, starting out with finite-difference approaches.[401–411]
By employing large grids, the atomic energies and wave functions can be straightforwardly converged to the CBS limit. The MC-HF energies for
atoms have been known to high precision for a long time;[412–418] reproduction of the exact ground state energy, however, is a hard problem even
for the Li atom.[419] Ivanov and Schmelcher have employed finite differences to study atoms in weak to strong magnetic fields.[420–426]
More recently, finite difference calculations have been complemented with FEM[427,428] as well as B-spline approaches.[287–290,429–440] In
addition to real-space methods, also momentum space approaches have been suggested,[441–449] but they have not become widely used. As was
discussed in Section 3, the finite element and B-spline methods are similar to each other: either one can be used for a systematic and smooth
approach to the CBS limit; also the implementations of the two approaches are similar. In addition to enabling variational calculations—
approaching the converged value strictly from above—and making sure orbital orthonormality is always honored, the use of an explicit radial basis
set instead of the implicit basis set of the finite-difference method also enables the straightforward use of post-HF methods.[450] For instance, Flores
et al. have calculated correlation energies with Møller-Plesset perturbation theory[58] truncated at the second order,[451–463] while Sundholm and
coworkers have relied on highly accurate multiconfigurational self-consistent field (MCSCF) methods[464] to study the ground state of the beryllium
atom,[465] electron affinities,[466–470] excitation energies and ionization potentials,[471–473] hyperfine structure,[474–476] nuclear quadrupole
moments,[476–491] the extended Koopmans' theorem,[492–494] and the Hiller–Sucher–Feinberg identity.[495] Braun[496] and Engel and coworkers[497–499]
have in turn reported finite-element calculations of atoms in strong magnetic fields. Approaches based on FEM-DVR have also been developed:
Hochstuhl and Bonitz have implemented a restricted active space procedure for modeling photoionization of many-electron atoms.[500] For more details
and references on MC calculations on atoms, I refer to the recent review by Froese Fischer and coworkers in Ref. [450].
As is obvious from the above references, multiple programs are available for atomic calculations at the HF and post-HF levels of theory. How-
ever, the situation is not as good for DFT. Note that unlike wave function based methods such as MC-HF, in theory a single-determinant
approach should suffice for DFT also in the case of significant strong correlation effects.[2] Unfortunately, this is true only for the (unknown) exact
exchange-correlation functional, while approximate functionals typically yield unreliable estimates for strongly correlated molecular systems;[3–5]
this is an active research topic within the DFT community.[501–508]
Before gaining acceptance among chemists, DFT was popular for a long time within the solid-state physics community. As was already stated
above, solid-state calculations typically employ pseudopotentials[315–321] or PAWs[322] to avoid explicit treatment of the highly localized and
chemically inactive core electrons.[509] Even though DFT calculations on atoms are the starting point for the construction of pseudopotentials and
PAW setups—meaning that a number of atomic DFT programs such as fhi98PP,[510] atompaw,[511] APE,[512] as well as one in GPAW[249] are com-
monly available—the scope of these programs appears to be limited, restricting their applicability for general calculations.
Especially, atomic DFT programs typically employ spherically averaged occupations

X X X ð2l + 1Þf nl unl ðrÞ 2


nðrÞ = f nl jψ nlm ðr Þj2 = , ð32Þ
nl m nl
4π r

which allows for huge performance benefits, since now calculations employing the local spin density approximation[2] (LDA) or the generalized
gradient approximation[513] (GGA) can be formulated as a set of coupled one-dimensional radial Schrödinger equations that can be solved in a
matter of seconds even in large grids. Using such an approach, for instance Andrae, Brodbeck, and Hinze have investigated closed-shell 1S states
for 2 ≤ Z ≤ 102 to study differences between HF and DFT approaches.[514]
Unfortunately, the restriction to spherical symmetry prevents the application of such programs to, for example, the prediction of electric properties,
as electric fields break spherical symmetry. Furthermore, due to the computational and theoretical challenges of using exact exchange in solid-state
14 of 31 LEHTOLA

calculations, support for hybrid exchange-correlation functionals is limited in the existing atomic DFT programs. To my best knowledge, only one atomic
program that supports both HF and DFT calculations has been described in the literature,[515] but it only supports calculations at LDA level, lacks hybrid
functionals, and employs cubic Hermite basis functions. Another program employing higher-order basis functions has been reported, but it is again
restricted to LDA calculations and also lacks support for HF.[516–519] However, as discussed in Section 5.1, I have recently solved this issue in Ref. [285].

4.3 | Diatomic molecules


At the nonrelativistic level of theory, the Hamiltonian for the electrons in a diatomic molecule composed of nuclei with charges ZA and ZB sepa-
rated by RAB is given by

X X ZA X ZB X 1 ZA ZB
^el = − 1
H r2i − − + + ð33Þ
2 i i
riA i
riB i > j r ij RAB

and the electronic states are identified in analogy to atomic states with the term symbol Λ, where the total electronic angular momentum
2S + 1

quantum number Λ, measured along the intermolecular axis, is one of Σ, Π, Δ, Φ, Γ, …, with strong correlation again arising via degenerate levels
especially when the term symbol is not Σ. Although diatomic molecules are not as straightforward for numerical treatment as atoms due to the
existence of two nuclei with two nuclear cusps, it was noticed early on by McCullough that a fully numerical treatment is tractable via the choice
of a suitable coordinate system.[520–522]
Analogously to atoms, where the orbitals can be written in the form of Equation (2), placing the nuclei along the z axis the orbitals for a
diatomic molecule become expressible as

ψ nm ðr Þ = χ nm ðξ, ηÞeimϕ ; ð34Þ

the same blocking by m holds for any linear molecule. The m quantum number describes the orbital character (see Appendix for a brief discussion),
and the prolate spheroidal coordinates are given by

rA + rB
ξ= , 1 ≤ ξ < ∞, ð35Þ
R

rA −r B
η= , − 1 ≤ η ≤ 1: ð36Þ
R

with ϕ measuring the angle around the bond axis. The coordinate system given by Equations (35) and (36) is illustrated in Figures 3–4, and 5. Even
though the ξ and η coordinates in the (ξ, η, ϕ) prolate spheroidal coordinate system do not have as clear roles as (r, θ, ϕ) in spherical polar coordi-
nates, as can be seen from Figures 4 and 5, the prolate spheroidal coordinate ξ can still be roughly identified as a “radial” coordinate, whereas η
can be identified as an “angular” coordinate that describes variation along the bond direction.
As in the spherical polar coordinate system for atoms, the nuclear attraction integrals have no singularities in the (ξ, η, ϕ) coordinate system in
the case of diatomic molecules, allowing for smooth and quick convergence to the CBS limit. Like in the case of atoms, where the volume element
r2dΩ cancels the r−1 divergence of the nuclear Coulomb attraction, the volume element of the prolate spheroidal coordinate system[523]

1  
dV = R3 ξ2 − η2 dξdηdϕ ð37Þ
8

F I G U R E 3 Illustration of the
r
A rB diatomic coordinate system defined
by nuclei A and B, depicted by the
large balls. The position of an electron
R (small ball) is parametrized in terms of
A B distances rA and rB from the nuclei A
and B. The ellipse with foci at A and B
describes an isosurface of ξ, which is
drawn by η 2 [−1, 1]
LEHTOLA 15 of 31

F I G U R E 4 Illustration of isosurfaces of ξ in terms of


distances parallel to the bond (along the molecular z axis)
and perpendicular to the bond (in the molecular (x, y) plane). 1.0
The isosurface for ξ = 1 is the straight line connecting the
nuclei, with larger values of ξ corresponding to larger and
larger ellipses. For ξ ! ∞ the isosurfaces become circles. In
the figure, $R_h = R/2$, the two nuclei are positioned at
$(−R/2,0)$ and at $(R/2,0)$ 0.5

r ⊥ /R h
0.0

-0.5

-1.0

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5


r /R h

1
r⊥ /R h

-1

F I G U R E 5 Illustration of isosurfaces of η in terms of -2


distances parallel to the bond (along the molecular z axis)
and perpendicular to the bond (in the molecular (x, y) plane).
The isosurfaces for η = −1 and η = +1 are the straight lines -3
emanating from the nuclei toward –∞ and +∞, respectively,
with intermediate values of η corresponding to the curved
isosurfaces. For η = 0 the isosurface is the plane (x, y, z = 0). -4
In the figure, $R_h = R/2$, the two nuclei are positioned at -4 -3 -2 -1 0 1 2 3 4
$(−R/2,0)$ and at $(R/2,0)$ r /R h

can be rewritten in the form

1
dV = R rA ðξ, ηÞrB ðξ, ηÞ dξdηdϕ ð38Þ
2

where the product rArB kills off the divergences of rA−1 and r B−1 in attraction integrals. Equally importantly, the two-electron integrals have a com-
pact representation in the prolate spheroidal coordinate system due to the Neumann expansion[523]
16 of 31 LEHTOLA

1 8π X∞ X L
ðL− jMjÞ! jMj jMj  M *
= ð − 1ÞM P ðξ < ÞQL ðξ > ÞY M
L ðΩ1 Þ Y L ðΩ2 Þ : ð39Þ
r12 R L = 0 M = −L ðL + jMjÞ! L

As is obvious from Equations (35) and (36), the coordinate system depends on the internuclear distance. This means that there are no egg-box
effects in the diatomic calculations; however, the accuracy of the (ξ, η) description now depends on the geometry of the system, as discussed in
Ref. [394].
In McCullough's pioneering approach, the orbitals are expanded as[520]

X

χ nm ðξ,ηÞ = f nml ðξÞPm
l ðηÞ, ð40Þ
l = jmj

where fnml(ξ) are unknown functions determined on a grid using finite differences, and Pm
l are normalized associated Legendre functions. Alterna-
tively, rearranging Equations (34) and (40) the expansion can be written as[521]

X

ψ nm ðr Þ = Xnml ðξÞY m
l ðη,ϕÞ, ð41Þ
l = jmj

pffiffiffiffiffiffi
where X nml ðξÞ = f nml ðξÞ= 4π . McCullough called the approach “the partial-wave self-consistent field method” (PW-SCF), as in practice only a finite
number of partial waves l can be included in the expansion. Pioneering work by McCullough and coworkers applied the PW-SCF method to the
study of parallel polarizabilities[524–526] and the accuracy of LCAO calculations.[527] The work on the approach continued with MCSCF calcula-
tions[528,529] and excited states,[530,531] quadrupole moments,[532] post-HF methods,[533,534] the extended Koopmans' theorem,[535] hyperfine
splitting constants,[536] the chromium[537] and copper dimers,[538] as well as magnetic hyperfine parameters.[539] Adamowicz and coworkers used
the program for post-HF methods[540–542] and electric polarizabilities,[543] and Chipman used it for calculating spin densities.[544]
Contemporaneously, motivated by McCullough's initial work, Laaksonen, Sundholm, and Pyykkö pursued an approach in which χ nm(ξ, η) in
Equation (34) is determined directly with a finite difference procedure.[545,546] At the same time, Becke developed a program for LDA calculations
on diatomic molecules,[547–551] proposing the use of a further coordinate transform

ξ = coshμ, 0 ≤ μ < ∞, ð42Þ

η = cosν, 0 ≤ ν ≤ π, ð43Þ

which eliminates nuclear cusps in the wave function. This can be easily seen by solving Equations (35) and (36) for the distances from the nuclei
and by substituting Taylor expansions of Equations (42) and (43) around (μ, ν) = (0, 0), and around (μ, ν) = (0, π) with Δν = ν − π:

1 1    
rA ≈ R 1 + μ2 − ν2 + O μ4 + O ν4 , ð44Þ
4 4

1 1    
rB ≈ R 1 + μ2 − ðΔνÞ2 + O μ4 + O ðΔνÞ4 : ð45Þ
4 4

Thus, while an exponential function exp (−ζrA) centered on atom A has a cusp in both ξ and η in the (ξ, η) coordinate system, there is no cusp
in the (μ, ν) coordinate system. Thanks to Equation (44), the exponential turns into a Gaussian function, which is much easier to represent numeri-
cally in both μ and ν.
The coordinate transform of Equations (42) and (43) was instantly adopted by Laaksonen, Sundholm, and Pyykkö for HF,[552–559] MC-
SCF,[560] and density functional calculations.[561,562] Laaksonen, Sundholm, Pyykkö, and others also presented applications of the method to the
calculation of electric field gradients,[563,564] nuclear quadrupole moments,[565–567] as well as repulsive interatomic potentials.[568] Note, however,
that Becke's own work employed a further transform from the (μ, ν) coordinates, see Ref. [547].
The development of the program originally written by Laaksonen, Sundholm, and Pyykkö was taken over by Kobus,[281,282,569] who proposed
an alternative relaxation approach for solving the SCF equations,[570] and extensively studied deficiencies of LCAO basis sets,[251,571–587] hydro-
gen halides,[587,588] multipole moments and parallel polarizabilities,[589,590] and molecular orbitals in a Xe-C ion-target system.[591]
The X2DHF program is open source and is still maintained, and publicly available on the internet.[283] X2DHF has been used, for example, by
Jensen in the development of polarization consistent basis sets,[159–161,163,592,593] which suggested that some numerical HF energies in the
LEHTOLA 17 of 31

literature were inaccurate due to an insufficient value for the practical infinity r∞.[594] Namely, as the Coulomb and exchange potentials are deter-
mined in X2DHF[281,282,556] by relaxation approaches that start from asymptotic multipole expansions,[552] the practical infinity r∞ may need to be
several hundred atomic units away to reach fully converged results, even if the electron density itself typically vanishes in a fraction of this
distance.
The X2DHF program was also used by Grabo and Gross for optimized effective potential calculations within the KLI approximation,[595,596] by
Karasiev and coworkers for density functional development,[597–605] by Halkier and Coriani for computing molecular electric quadrupole
moments,[606] Roy and Thakkar who studied MacLaurin expansions of the electron momentum density for 78 diatomic molecules,[607] Weigend,
Furche, and Ahlrichs for studying total energy and atomization energy basis set errors in quadruple-ζ basis sets,[608] Shahbazian and Zahedi who
studied basis set convergence patterns,[609] Pruneda, Artacho, and Kuzmin for the calculation of range parameters,[610–612] Madsen and Madsen
for modeling high-harmonic generation,[613] Williams et al. who reported numerical HF energies for transition metal diatomics,[614] Madsen and
coworkers who calculated structure factors for tunneling,[615–618] Kornev and Zon who studied anti-Stokes-enhanced tunneling ionization of polar
molecules,[619] and Endo and coworkers who studied laser tunneling ionization of NO.[620]
Based on the results of Laaksonen et al., Kolb and coworkers developed a FEM program for HF and density functional calculations employing
triangular basis functions, again in the (μ, ν) coordinate system.[269,621–629] A finite element variant employing rectangular elements was also
developed by Sundholm and coworkers for MCSCF calculations,[464,630] whereas a multigrid conjugate residual HF solver for diatomics has been
described by Davstad.[631] Makmal, Kümmel, and Kronik studied fully numerical all-electron solutions to the optimized effective potential equation
with and without the KLI approximation, employing a finite difference approach similarly to X2DHF.[632,633] Morrison et al. have also discussed vari-
ous numerical approaches for diatomic molecules.[634–636] Artemyev and coworkers have reported an implementation of the partial-wave
approach based on B-splines.[637]
Apparently unaware of the unanimous agreement between Becke, Laaksonen, Sundholm, Pyykkö, Kobus, Kolb, and others in the quantum
chemistry community on the supremacy of the (μ, ν) coordinate system, several works utilizing the original (ξ, η) coordinates defined by Equa-
tions (35) and (36) have been published in the molecular physics literature. For instance, a B-spline CI program has been reported by Vanne and
Saenz,[638] whereas FEM-DVR expansions have been employed by Tao and coworkers[639,640] and Guan and coworkers[641,642] for studying
dynamics of the H2 molecule, as well as by Tolstikhin, Morishita, and Madsen for studying HeH2+.[643] Haxton and coworkers have reported multi-
configuration time-dependent HF calculations of many-electron diatomic molecules with FEM-DVR basis sets,[644] whereas Larsson and
coworkers[645] and Yue and coworkers[646] have studied correlation effects in ionization of diatomic molecules employing restricted-active space
calculations also with FEM-DVR. A FEM-DVR-based program supporting HF calculations for diatomics has been published by Zhang and
coworkers.[647] All of these programs employ a product grid in ξ and η. Momentum space approaches have also been developed for diatomic
molecules[648–651] as well as larger linear molecules,[648,652,653] but they have not become widely used.

5 | S UM M ARY A N D O U T L O O K

I have reviewed approaches for non-relativistic all-electron electronic structure calculations based on the LCAO approach employing a variety of
radial functions, and discussed their benefits and drawbacks. The shortcomings of LCAO calculations were used to motivate a need for fully
numerical approaches, which were also briefly discussed along with their good and bad sides. I pointed out that fully numerical approaches can be
made more affordable for atoms and diatomic molecules by exploiting the symmetry inherent in these systems, and reviewed the extensive litera-
ture on fully numerical quantum chemical calculations on atoms and diatomic molecules.

5.1 | SCF approaches


The development of novel density functionals might especially benefit from access to affordable, fully numerical electronic structure approaches
for atoms and diatomic molecules. As evidenced by the existence of several numerically ill-behaved functionals,[654–658] it has not been common
procedure to assess the behavior of new functionals with fully numerical calculations. The flexibility of fully numerical basis sets guarantees that
any numerical instabilities in a functional are instantly seen in the results of a calculation, in contrast to calculations in small LCAO basis sets.
I have recently published a program for fully numerical electronic structure calculations on atoms[285] and diatomic molecules[394] for HF and
DFT called HELFEM.[659] In the atomic program,[285] finite elements are used for the radial unl(r) expansion, whereas the diatomic program[394]
employs the partial wave approach of McCullough[520,521] to represent the angular ν and ϕ directions using spherical harmonics, while one-
dimensional finite elements are again used in the radial μ direction as in the atomic case.[285]
The freely available, open-source HELFEM program introduces support for hybrid density functionals, as well as meta-GGA functionals, which
have not been publicly available in other fully numerical program packages for calculations on atoms or diatomic molecules; hundreds of func-
tionals are supported via an interface to the LIBXC library.[660] In addition to calculations in a finite electric field,[285,394] HELFEM has been recently
extended to calculations in finite magnetic fields that are especially challenging to model with popular LCAO approaches.[661] Due to the close
18 of 31 LEHTOLA

coupling of HELFEM to LIBXC, my hope is that the program will be adopted for the development of new density functionals, as both atoms and
diatomic molecules can be calculated easily and efficiently directly at the basis set limit.
Unlike previous programs relying on relaxation approaches to solve the Poisson and HF/KS equations, the approach used in HELFEM is fully
variational, guaranteeing a monotonic convergence from above to the ground state energy. Extremely fast convergence to the radial grid limit is
observed both for the atomic and diatomic calculations, with less than 100 radial functions typically guaranteeing full convergence even in the
case of heavy atoms.
The angular expansions are typically small in the case of atoms, even in the presence of finite electric fields that break spherical symmetry.[285]
However, in the case of diatomics, the angular expansions may converge less rapidly. Still, I have shown in Ref. [394] that a nonuniform expansion
length in m affords considerably cheaper calculations with no degradation in accuracy. Although Ref. [394] shows that the partial wave expansion
yields extremely compact expansions for a large number of molecules as compared to finite-difference calculations with X2DHF of the same quality,
going beyond fourth period elements may require further optimizations and parallelization in the HELFEM program.
The approach used in HELFEM is fast for light molecules that do not require going to high partial waves l, but the calculations slow down con-
siderably for heavy atoms due to two reasons. First, the 1s core orbitals become smaller and smaller going deeper in the periodic table, thus
requiring smaller length scales to be represented in the partial wave expansion. Second, heavier atoms are also bigger in size, which makes the
bond lengths larger, further decreasing the angular size of the core orbitals. Going to heavier atoms will also require the treatment of relativistic
effects, which have not been discussed in the present work.
While the Coulomb matrix can be formed in little time even in large calculations, the exchange contributions in both HF and DFT scale as the
cube of the number of basis functions, N3. However, as the program is still quite new, it is quite likely that these problems can be circumvented
via optimizations to the code and/or further parallelization.

5.2 | Post-HF approaches


Although fully numerical calculations afford arbitrary precision for atoms and diatomic molecules within HF or DFT, the accuracy of these
methods themselves is insufficient for many applications. More accurate estimates can be obtained by switching to a post-HF level of theory,
such as MC theory,[662,663] density matrix renormalization group theory,[664] or coupled-cluster theory.[59] However, going to the post-HF level
causes further complications.
As is commonly known from Gaussian-basis calculations, the basis set convergence of the correlation energy is very slow.[152,665–669] In fact,
the problem appears already for two-electron systems,[670] for example, in CI calculations on the 1s2 state of He as discovered in the late 1970s
in seminal work by Silverstone and coworkers.[431,432] Because of this, even atomic fully numerical calculations have to rely on extrapolation
approaches.[460,462,463] Although extrapolation techniques are commonly used even for SCF energies in the context of LCAO calculations, as the
SCF energy affords exponential convergence, such approaches are usually not necessary except for highly accurate benchmark studies. Regardless
of the used basis set, the key to obtaining accurate post-HF energetics, however, is the extrapolation to the CBS limit, which has been studied
extensively in the literature using a wealth of extrapolation formulas.[667,668,671–695]
Moreover, although it is extremely simple to obtain large sets of virtual orbitals for both atoms and diatomic molecules,[285,394] it is unlikely
that they can be fully used in correlation treatmentss: even calculating and storing the two-electron integrals, which is a rank-4 tensor, may rapidly
become challenging in large numerical basis sets. For instance, the number of operations in CI and coupled-cluster theory scales as O(onvn + 2),
where o and v are the number of occupied and virtual orbitals, respectively, and n is the maximum level of substitutions included in the calculation.
Taking the n = 2 case for doubles, the v4 scaling quickly makes calculations infeasible, meaning that only a small subset of virtual orbitals can be
included in the treatment. If all excitations are included, as in the full CI approach, the scaling becomes even more prohibitive: per common knowl-
edge, one is limited to studying problem sizes of at most 20 electrons in 20 orbitals due to the exponential scaling algorithm. While the density
matrix renormalization group[664] approach is a more efficient parameterization of the correlation problem, it too is exponentially scaling in gen-
eral, and is limited in such cases to problem sizes of about 50 electrons in 50 orbitals.[696]
Although post-HF calculations on atoms and diatomic molecules have been reported by several authors in the literature, as discussed above,
specialized algorithms for atoms and diatomic molecules have fewer advantages over general three-dimensional programs at post-HF levels of
theory, since a large set of virtual orbitals is anyhow necessary. It may then be less important to be able to treat the mean-field problem to arbi-
trary precision, when significant errors may still be made in the correlation treatment due the intrinsic limitations therein caused by the computa-
tional scaling.
Despite this slightly disconcerting analysis, there is still room for specialized calculations on atoms and diatomic molecules. In cases where HF
or DFT is sufficient, a specialized numerical approach is more cost-effective than a general one. Furthermore, the development and verification of
novel, general three-dimensional programs benefits from extremely accurate benchmark data, which is achievable through the specialized atomic
and diatomic approaches. The specialization of post-HF approaches to the case of atoms or diatomic molecules may allow for a more efficient use
of symmetries as well, enabling one to perform slightly larger calculations than within a completely general approach.
LEHTOLA 19 of 31

ACKNOWLEDGMENTS

I thank Dage Sundholm and Jacek Kobus for discussions, and Dage Sundholm, Lukas Wirz, Jacek Kobus, and Pekka Pyykkö for comments on the manuscript.

ORCID

Susi Lehtola https://orcid.org/0000-0001-6296-8103

RE FE R ENC E S

[1] P. Hohenberg, W. Kohn, Phys. Rev. 1964, 136, B864.


[2] W. Kohn, L. J. Sham, Phys. Rev. 1965, 140, A1133.
[3] A. D. Becke, J. Chem. Phys. 2014, 140, 18A301.
[4] R. O. Jones, Rev. Mod. Phys. 2015, 87, 897.
[5] N. Mardirossian, M. Head-Gordon, Mol. Phys. 2017, 115, 2315.
[6] A. D. Becke, J. Chem. Phys. 1986, 85, 7184.
[7] A. D. Becke, Phys. Rev. A 1988, 38, 3098.
[8] A. D. Becke, J. Chem. Phys. 1997, 107, 8554.
[9] J. Sun, J. P. Perdew, A. Ruzsinszky, Proc. Natl. Acad. Sci. USA 2015, 112, 685.
[10] J. Sun, A. Ruzsinszky, J. Perdew, Phys. Rev. Lett. 2015, 115, 036402.
[11] M. G. Medvedev, I. S. Bushmarinov, J. Sun, J. P. Perdew, K. A. Lyssenko, Science 2017, 355, 49.
[12] K. R. Brorsen, Y. Yang, M. V. Pak, S. Hammes-Schiffer, J. Phys. Chem. Lett. 2017, 8, 2076.
[13] K. P. Kepp, Science 2017, 356, 496.
[14] M. G. Medvedev, I. S. Bushmarinov, J. Sun, J. P. Perdew, K. A. Lyssenko, Science 2017, 356, 496.
[15] S. Hammes-Schiffer, Science 2017, 355, 28.
[16] M. Korth, Angew. Chem., Int. Ed. 2017, 56, 5396.
[17] T. Gould, J. Chem. Theory Comput. 2017, 13, 2373.
[18] P. D. Mezei, G. I. Csonka, M. Kállay, J. Chem. Theory Comput. 2017, 13, 4753.
[19] Y. Wang, X. Wang, D. G. Truhlar, X. He, J. Chem. Theory Comput. 2017, 13, 6068.
[20] K. P. Kepp, Phys. Chem. Chem. Phys. 2018, 20, 7538.
[21] N. Q. Su, Z. Zhu, X. Xu, Proc. Natl. Acad. Sci. USA 2018, 115(10), 2287.
[22] L. F. Pašteka, E. Eliav, A. Borschevsky, U. Kaldor, P. Schwerdtfeger, Phys. Rev. Lett. 2017, 118, 023002.
[23] J. Almlöf, K. Faegri, K. Korsell, J. Comput. Chem. 1982, 3, 385.
[24] J. H. Van Lenthe, R. Zwaans, H. J. J. Van Dam, M. F. Guest, J. Comput. Chem. 2006, 27, 926.
[25] S. Lehtola, J. Chem. Theory Comput. 2019, 15, 1593.
[26] B. O. Roos, A. C. Borin, L. Gagliardi, Angew. Chem., Int. Ed. 2007, 46, 1469.
[27] A. Karton, L. K. McKemmish, Aust. J. Chem. 2018, 71, 804.
[28] L. A. Terrabuio, T. Q. Teodoro, M. G. Rachid, R. L. A. Haiduke, J. Phys. Chem. A 2013, 117, 10489.
[29] P. Pyykkö, Chem. Rev. 2012, 112, 371.
[30] P. Pyykkö, Annu. Rev. Phys. Chem. 2012, 63, 45.
[31] J. P. Desclaux, P. Pyykkö, Chem. Phys. Lett. 1976, 39, 300.
[32] P. Pyykkö, J. Desclaux, Chem. Phys. Lett. 1976, 42, 545.
[33] P. Pyykkö, J. Chem. Soc., Faraday Trans. 2 1979, 75, 1256.
[34] W. D. Sepp, D. Kolb, W. Sengler, H. Hartung, B. Fricke, Phys. Rev. A 1986, 33, 3679.
[35] , D. Heinemann, W.-D. Sepp, D. Kolb, B. Fricke, Chem. Phys. Lett. 1993, 211, 119.
T. Baş tug
[36] P. Pyykkö, Phys. Scr. 1979, 20, 647.
[37] J. G. Snijders, P. Pyykkö, Chem. Phys. Lett. 1980, 75, 5.
[38] C. Park, J. E. Almlöf, Chem. Phys. Lett. 1994, 231, 269.
[39] E. van Lenthe, E. J. Baerends, J. G. Snijders, J. Chem. Phys. 1994, 101, 9783.
[40] L. Visscher, J. Styszyñski, W. C. Nieuwpoort, J. Chem. Phys. 1996, 105, 1987.
[41] , K. Rashid, W.-D. Sepp, D. Kolb, B. Fricke, Phys. Rev. A 1997, 55, 1760.
T. Baş tug
[42] W. Liu, G. Hong, D. Dai, L. Li, M. Dolg, Theor. Chem. Acc. Theory Comput. Model. (Theor. Chim. Acta) 1997, 96, 75.
[43] W. A. de Jong, J. Styszynski, L. Visscher, W. C. Nieuwpoort, J. Chem. Phys. 1998, 108, 5177.
[44] T. Nakajima, K. Hirao, Chem. Phys. Lett. 1999, 302, 383.
[45] T. Nakajima, T. Suzumura, K. Hirao, Chem. Phys. Lett. 1999, 304, 271.
[46] K. Fægri, T. Saue, J. Chem. Phys. 2001, 115, 2456.
[47] J. Anton, T. Jacob, B. Fricke, E. Engel, Phys. Rev. Lett. 2002, 89, 213001.
[48] H. Zhang, O. Kullie, D. Kolb, J. Phys. B At. Mol. Opt. Phys. 2004, 37, 905.
[49] B. O. Roos, P.-Å. Malmqvist, L. Gagliardi, J. Am. Chem. Soc. 2006, 128, 17000.
[50] S. Höfener, R. Ahlrichs, S. Knecht, L. Visscher, ChemPhysChem 2012, 13, 3952.
[51] Y.-L. Wang, H.-S. Hu, W.-L. Li, F. Wei, J. Li, J. Am. Chem. Soc. 2016, 138, 1126.
[52] S. Knecht, H. J. A. Jensen, T. Saue, Nat. Chem. 2019, 11, 40.
[53] W. Heitler, F. London, Z. Phys. 1927, 44, 455.
20 of 31 LEHTOLA

[54] J. E. Lennard-Jones, Trans. Faraday Soc. 1929, 25, 668.


[55] C. Roothaan, Rev. Mod. Phys. 1951, 23, 69.
[56] J. A. Pople, R. K. Nesbet, J. Chem. Phys. 1954, 22, 571.
[57] J. A. Pople, P. M. W. Gill, B. G. Johnson, Chem. Phys. Lett. 1992, 199, 557.
[58] C. Møller, M. S. M. Plesset, Phys. Rev. 1934, 46, 618.
[59] J. Čížek, J. Chem. Phys. 1966, 45, 4256.
[60] J. A. Pople, Rev. Mod. Phys. 1999, 71, 1267.
[61] E. Cancès, C. Le Bris, Int. J. Quantum Chem. 2000, 79, 82.
[62] V. R. Saunders, I. H. A. Hillier, Int. J. Quantum Chem. 1973, 7, 699.
[63] A. V. Mitin, J. Comput. Chem. 1988, 9, 107.
[64] G. Dömötör, M. Bán, Comput. Chem. 1989, 13, 53.
[65] K. N. Kudin, G. E. Scuseria, J. Chem. Phys. 2002, 116, 8255.
[66] X. Hu, W. Yang, J. Chem. Phys. 2010, 132, 054109.
[67] P. Pulay, Chem. Phys. Lett. 1980, 73, 393.
[68] P. Pulay, J. Comput. Chem. 1982, 3, 556.
[69] W. R. Wessel, J. Chem. Phys. 1967, 47, 3253.
[70] J. Douady, Y. Ellinger, R. Subra, B. Levy, Comput. Phys. Commun. 1979, 17, 23.
[71] J. Douady, Y. Ellinger, R. Subra, B. Levy, J. Chem. Phys. 1980, 72, 1452.
[72] G. B. Bacskay, Chem. Phys. 1981, 61, 385.
[73] G. B. Bacskay, Chem. Phys. 1982, 65, 383.
[74] M. Head-Gordon, J. A. Pople, M. J. Frisch, Int. J. Quantum Chem. 1989, 36, 291.
[75] G. Chaban, M. W. Schmidt, M. S. Gordon, Theor. Chem. Acc. Theory Comput. Model. (Theor. Chim. Acta) 1997, 97, 88.
[76] J. B. Francisco, J. M. Martínez, L. Martínez, J. Chem. Phys. 2004, 121, 10863.
[77] L. Thøgersen, J. Olsen, D. Yeager, P. Jørgensen, P. Sałek, T. Helgaker, J. Chem. Phys. 2004, 121, 16.
[78] L. Thøgersen, J. Olsen, A. Köhn, P. Jørgensen, P. Sałek, T. Helgaker, J. Chem. Phys. 2005, 123, 074103.
[79] F. Neese, Chem. Phys. Lett. 2000, 325, 93.
[80] T. Van Voorhis, M. Head-Gordon, Mol. Phys. 2002, 100, 1713.
[81] B. D. Dunietz, T. Van Voorhis, M. Head-Gordon, J. Theor. Comput. Chem. 2002, 01, 255.
[82] C. Yang, J. C. Meza, L.-W. Wang, SIAM J. Sci. Comput. 2007, 29, 1854.
[83] K. N. Kudin, G. E. Scuseria, ESAIM Math. Model. Numer. Anal. 2007, 41, 281.
[84] A. J. Garza, G. E. Scuseria, J. Chem. Phys. 2012, 137, 054110.
[85] A. J. Garza, G. E. Scuseria, J. Chem. Phys. 2015, 142, 164104.
[86] C. Zener, Phys. Rev. 1930, 36, 51.
[87] J. C. Slater, Phys. Rev. 1930, 36, 57.
[88] C. Eckart, Phys. Rev. 1930, 36, 878.
[89] E. Clementi, C. Roetti, At. Data Nucl. Data Tables 1974, 14, 177.
[90] D. Chakraborty, P. W. Ayers, J. Math. Chem. 2011, 49, 1810.
[91] S. Laricchia, L. A. Constantin, E. Fabiano, F. Della Sala, J. Chem. Theory Comput. 2014, 10, 164.
[92] T. Kato, Commun. Pure Appl. Math. 1957, 10, 151.
[93] J. Katriel, E. R. Davidson, Proc. Natl. Acad. Sci. USA 1980, 77, 4403.
[94] H. J. Silverstone, Phys. Rev. A 1981, 23, 1030.
[95] I. Mayer, Simple Theorems, Proofs and Derivations in Quantum Chemistry, Kluwer Academic/Plenum Publishers, New York 2003.
[96] P. M. Boerrigter, G. Te Velde, J. E. Baerends, Int. J. Quantum Chem. 1988, 33, 87.
[97] G. te Velde, E. J. Baerends, Phys. Rev. B 1991, 44, 7888.
[98] G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra, S. J. A. van Gisbergen, J. G. Snijders, T. Ziegler, J. Comput. Chem. 2001, 22, 931.
[99] M. Franchini, P. H. T. Philipsen, L. Visscher, J. Comput. Chem. 2013, 34, 1819.
[100] R. G. Parr, H. W. Joy, J. Chem. Phys. 1957, 26, 424.
[101] H. W. Joy, R. G. Parr, J. Chem. Phys. 1958, 28, 448.
[102] A. F. Saturno, R. G. Parr, J. Chem. Phys. 1958, 29, 490.
[103] A. F. Saturno, R. G. Parr, J. Chem. Phys. 1960, 33, 22.
[104] T. Koga, K. Kanayama, A. J. Thakkar, Int. J. Quantum Chem. 1997, 62, 1.
[105] T. Koga, K. Kanayama, Chem. Phys. Lett. 1997, 266, 123.
[106] T. Koga, K. Kanayama, J. Phys. B At. Mol. Opt. Phys. 1997, 30, 1623.
[107] T. Koga, J. M. García de la Vega, B. Miguel. Chem. Phys. Lett. 1998, 283, 97–101.
[108] T. Koga, T. Shimazaki, T. Satoh, J. Mol. Struct. THEOCHEM 2000, 496, 95.
[109] I. I. Guseinov, M. Ertürk, Int. J. Quantum Chem. 2009, 109, 176.
[110] H. Shull, P. Löwdin, J. Chem. Phys. 1959, 30, 617.
[111] M. Rotenberg, Ann. Phys. (N. Y). 1962, 19, 262.
[112] J. S. Avery, J. E. Avery, Mol. Phys. 2012, 110, 1593.
[113] M. F. Herbst, A. Dreuw, J. E. Avery, J. Chem. Phys. 2018, 149, 084106.
[114] S. F. Boys, Proc. R. Soc. A Math. Phys. Eng. Sci. 1950, 200, 542.
[115] L. E. McMurchie, E. R. Davidson, J. Comput. Phys. 1978, 26, 218.
[116] S. Obara, A. Saika, J. Chem. Phys. 1986, 84, 3963.
[117] P. M. W. Gill, J. A. Pople, Int. J. Quantum Chem. 1991, 40, 753.
[118] R. Kikuchi, J. Chem. Phys. 1954, 22, 148.
LEHTOLA 21 of 31

[119] W. J. Hehre, R. F. Stewart, J. A. Pople, J. Chem. Phys. 1969, 51, 2657.


[120] M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. Su, T. L. Windus,
M. Dupuis, J. A. Montgomery, J. Comput. Chem. 1993, 14, 1347.
[121] M. F. Guest, I. J. Bush, H. J. J. Van Dam, P. Sherwood, J. M. H. Thomas, J. H. Van Lenthe, R. W. A. Havenith, J. Kendrick, Mol. Phys. 2005, 103, 719.
[122] H.-J. Werner, P. J. Knowles, G. Knizia, F. R. Manby, M. Schütz, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 242.
[123] F. Neese, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73.
[124] K. Aidas, C. Angeli, K. L. Bak, V. Bakken, R. Bast, L. Boman, O. Christiansen, R. Cimiraglia, S. Coriani, P. Dahle, E. K. Dalskov, U. Ekström,
T. Enevoldsen, J. J. Eriksen, P. Ettenhuber, B. Fernández, L. Ferrighi, H. Fliegl, L. Frediani, K. Hald, A. Halkier, C. Hättig, H. Heiberg, T. Helgaker,
A. C. Hennum, H. Hettema, E. Hjertenaes, S. Høst, I. M. Høyvik, M. F. Iozzi, B. Jansík, H. J. A. Jensen, D. Jonsson, P. Jørgensen, J. Kauczor,
S. Kirpekar, T. Kjaergaard, W. Klopper, S. Knecht, R. Kobayashi, H. Koch, J. Kongsted, A. Krapp, K. Kristensen, A. Ligabue, O. B. Lutnaes, J. I. Melo,
K. V. Mikkelsen, R. H. Myhre, C. Neiss, C. B. Nielsen, P. Norman, J. Olsen, J. M. H. Olsen, A. Osted, M. J. Packer, F. Pawlowski, T. B. Pedersen,
P. F. Provasi, S. Reine, Z. Rinkevicius, T. A. Ruden, K. Ruud, V. V. Rybkin, P. Sałek, C. C. M. Samson, A. S. de Merás, T. Saue, S. P. A. Sauer,
B. Schimmelpfennig, K. Sneskov, A. H. Steindal, K. O. Sylvester-Hvid, P. R. Taylor, A. M. Teale, E. I. Tellgren, D. P. Tew, A. J. Thorvaldsen,
L. Thøgersen, O. Vahtras, M. A. Watson, D. J. D. Wilson, M. Ziolkowski, H. Ågren, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4, 269.
[125] Y. Shao, Z. Gan, E. Epifanovsky, A. T. B. Gilbert, M. Wormit, J. Kussmann, A. W. Lange, A. Behn, J. Deng, X. Feng, D. Ghosh, M. Goldey, P. R. Horn,
L. D. Jacobson, I. Kaliman, R. Z. Khaliullin, T. Kus, A. Landau, J. Liu, E. I. Proynov, Y. M. Rhee, R. M. Richard, M. A. Rohrdanz, R. P. Steele,
E. J. Sundstrom, H. L. Woodcock III., P. M. Zimmerman, D. Zuev, B. Albrecht, E. Alguire, B. Austin, G. J. O. Beran, Y. A. Bernard, E. Berquist,
K. Brandhorst, K. B. Bravaya, S. T. Brown, D. Casanova, C. M. Chang, Y. Chen, S. H. Chien, K. D. Closser, D. L. Crittenden, M. Diedenhofen,
R. A. DiStasio Jr., H. Do, A. D. Dutoi, R. G. Edgar, S. Fatehi, L. Fusti-Molnar, A. Ghysels, A. Golubeva-Zadorozhnaya, J. Gomes, M. W. D. Hanson-
Heine, P. H. P. Harbach, A. W. Hauser, E. G. Hohenstein, Z. C. Holden, T. C. Jagau, H. Ji, B. Kaduk, K. Khistyaev, J. Kim, J. Kim, R. A. King,
P. Klunzinger, D. Kosenkov, T. Kowalczyk, C. M. Krauter, K. U. Lao, A. D. Laurent, K. V. Lawler, S. V. Levchenko, C. Y. Lin, F. Liu, E. Livshits,
R. C. Lochan, A. Luenser, P. Manohar, S. F. Manzer, S. P. Mao, N. Mardirossian, A. V. Marenich, S. A. Maurer, N. J. Mayhall, E. Neuscamman,
C. M. Oana, R. Olivares-Amaya, D. P. O’Neill, J. A. Parkhill, T. M. Perrine, R. Peverati, A. Prociuk, D. R. Rehn, E. Rosta, N. J. Russ, S. M. Sharada,
S. Sharma, D. W. Small, A. Sodt, T. Stein, D. Stück, Y. C. Su, A. J. W. Thom, T. Tsuchimochi, V. Vanovschi, L. Vogt, O. Vydrov, T. Wang, M. A. Watson,
J. Wenzel, A. White, C. F. Williams, J. Yang, S. Yeganeh, S. R. Yost, Z. Q. You, I. Y. Zhang, X. Zhang, Y. Zhao, B. R. Brooks, G. K. L. Chan,
D. M. Chipman, C. J. Cramer, W. A. Goddard III., M. S. Gordon, W. J. Hehre, A. Klamt, H. F. Schaefer III., M. W. Schmidt, C. D. Sherrill, D. G. Truhlar,
A. Warshel, X. Xu, A. Aspuru-Guzik, R. Baer, A. T. Bell, N. A. Besley, J. D. Chai, A. Dreuw, B. D. Dunietz, T. R. Furlani, S. R. Gwaltney, C. P. Hsu,
Y. Jung, J. Kong, D. S. Lambrecht, W. Z. Liang, C. Ochsenfeld, V. A. Rassolov, L. V. Slipchenko, J. E. Subotnik, T. van Voorhis, J. M. Herbert,
A. I. Krylov, P. M. W. Gill, M. Head-Gordon, Mol. Phys. 2015, 113, 184.
[126] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li,
M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-
Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang,
M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr.,
J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari,
A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma,
O. Farkas, J. B. Foresman, D. J. Fox, et al., Gaussian 16 Revision B.01, Gaussian, Inc., Wallingford, CT 2016.
[127] F. Aquilante, J. Autschbach, R. K. Carlson, L. F. Chibotaru, M. G. Delcey, L. de Vico, I. Fdez. Galván, N. Ferré, L. M. Frutos, L. Gagliardi, M. Garavelli,
A. Giussani, C. E. Hoyer, G. Li Manni, H. Lischka, D. Ma, P. Å. Malmqvist, T. Müller, A. Nenov, M. Olivucci, T. B. Pedersen, D. Peng, F. Plasser,
B. Pritchard, M. Reiher, I. Rivalta, I. Schapiro, J. Segarra-Martí, M. Stenrup, D. G. Truhlar, L. Ungur, A. Valentini, S. Vancoillie, V. Veryazov,
V. P. Vysotskiy, O. Weingart, F. Zapata, R. Lindh, J. Comput. Chem. 2016, 37, 506.
[128] T. Saue, L. Visscher, H. J. A. Jensen, R. Bast, DIRAC, a Relativistic Ab Initio Electronic Structure Program, Release DIRAC17, http://www.
diracprogram.org (accessed: May 2019).
[129] R. M. Parrish, L. A. Burns, D. G. A. Smith, A. C. Simmonett, A. E. DePrince III., E. G. Hohenstein, U. Bozkaya, A. Y. Sokolov, R. di Remigio,
R. M. Richard, J. F. Gonthier, A. M. James, H. R. McAlexander, A. Kumar, M. Saitow, X. Wang, B. P. Pritchard, P. Verma, H. F. Schaefer III.,
K. Patkowski, R. A. King, E. F. Valeev, F. A. Evangelista, J. M. Turney, T. D. Crawford, C. D. Sherrill, J. Chem. Theory Comput. 2017, 13, 3185.
[130] Q. Sun, T. C. Berkelbach, N. S. Blunt, G. H. Booth, S. Guo, Z. Li, J. Liu, J. D. McClain, E. R. Sayfutyarova, S. Sharma, S. Wouters, G. K.-L. Chan, Wiley
Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, e1340.
[131] R. Ditchfield, W. J. Hehre, J. A. Pople, J. Chem. Phys. 1971, 54, 724.
[132] W. J. Hehre, R. Ditchfield, J. A. Pople, J. Chem. Phys. 1972, 56, 2257.
[133] P. C. Hariharan, J. A. Pople, Theor. Chim. Acta 1973, 28, 213.
[134] P. C. Hariharan, J. A. Pople, Mol. Phys. 1974, 27, 209.
[135] M. S. Gordon, Chem. Phys. Lett. 1980, 76, 163.
[136] M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, M. S. Gordon, D. J. DeFrees, J. A. Pople, J. Chem. Phys. 1982, 77, 3654.
[137] R. C. Binning, L. A. Curtiss, J. Comput. Chem. 1990, 11, 1206.
[138] J.-P. Blaudeau, M. P. McGrath, L. A. Curtiss, L. Radom, J. Chem. Phys. 1997, 107, 5016.
[139] V. A. Rassolov, J. A. Pople, M. A. Ratner, T. L. Windus, J. Chem. Phys. 1998, 109, 1223.
[140] V. A. Rassolov, M. A. Ratner, J. A. Pople, P. C. Redfern, L. A. Curtiss, J. Comput. Chem. 2001, 22, 976.
[141] A. D. McLean, G. S. Chandler, J. Chem. Phys. 1980, 72, 5639.
[142] R. Krishnan, J. S. Binkley, R. Seeger, J. A. Pople, J. Chem. Phys. 1980, 72, 650.
[143] A. Schäfer, H. Horn, R. Ahlrichs, J. Chem. Phys. 1992, 97, 2571.
[144] A. Schäfer, C. Huber, R. Ahlrichs, J. Chem. Phys. 1994, 100, 5829.
[145] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297.
[146] D. Rappoport, F. Furche, J. Chem. Phys. 2010, 133, 134105.
[147] T. H. Dunning, J. Chem. Phys. 1989, 90, 1007.
[148] R. A. Kendall, T. H. Dunning, R. J. Harrison, J. Chem. Phys. 1992, 96, 6796.
22 of 31 LEHTOLA

[149] D. E. Woon, T. H. Dunning, J. Chem. Phys. 1993, 98, 1358.


[150] D. E. Woon, T. H. Dunning, J. Chem. Phys. 1994, 100, 2975.
[151] D. E. Woon, T. H. Dunning, J. Chem. Phys. 1995, 103, 4572.
[152] A. K. Wilson, T. van Mourik, T. H. Dunning, J. Mol. Struct. THEOCHEM 1996, 388, 339.
[153] B. P. Prascher, D. E. Woon, K. A. Peterson, T. H. Dunning, A. K. Wilson, Theor. Chem. Acc. 2011, 128, 69.
[154] T. Van Mourik, T. H. Dunning, Int. J. Quantum Chem. 2000, 76, 205.
[155] A. K. Wilson, D. E. Woon, K. A. Peterson, T. H. Dunning, J. Chem. Phys. 1999, 110, 7667.
[156] T. H. Dunning, K. A. Peterson, A. K. Wilson, J. Chem. Phys. 2001, 114, 9244.
[157] K. A. Peterson, T. H. Dunning, J. Chem. Phys. 2002, 117, 10548.
[158] J. G. Hill, K. A. Peterson, J. Chem. Phys. 2017, 147, 244106.
[159] F. Jensen, J. Chem. Phys. 2001, 115, 9113.
[160] F. Jensen, J. Chem. Phys. 2002, 116, 3502.
[161] F. Jensen, J. Chem. Phys. 2002, 117, 9234.
[162] F. Jensen, J. Phys. Chem. A 2007, 111, 11198.
[163] F. Jensen, T. Helgaker, J. Chem. Phys. 2004, 121, 3463.
[164] F. Jensen, J. Chem. Phys. 2012, 136, 114107.
[165] F. Jensen, J. Chem. Phys. 2013, 138, 014107.
[166] F. Jensen, J. Chem. Theory Comput. 2014, 10, 1074.
[167] D. S. Ranasinghe, G. A. Petersson, J. Chem. Phys. 2013, 138, 144104.
[168] D. S. Ranasinghe, M. J. Frisch, G. A. Petersson, J. Chem. Phys. 2015, 143, 214110.
[169] J. Almlöf, P. R. Taylor, J. Chem. Phys. 1987, 86, 4070.
[170] J. Almlöf, T. Helgaker, P. R. Taylor, J. Phys. Chem. 1988, 92, 3029.
[171] J. Almlöf, P. R. Taylor, J. Chem. Phys. 1990, 92, 551.
[172] P.-O. Widmark, P.-Å. Malmqvist, B. O. Roos, Theor. Chim. Acta 1990, 77, 291.
[173] P.-O. Widmark, B. J. Persson, B. O. Roos, Theor. Chim. Acta 1991, 79, 419.
[174] K. Pierloot, B. Dumez, P.-O. Widmark, B. O. Roos, Theor. Chim. Acta 1995, 90, 87.
[175] R. Pou-Amérigo, M. Merchán, I. Nebot-Gil, P.-O. Widmark, B. O. Roos, Theor. Chim. Acta 1995, 92, 149.
[176] B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark, J. Phys. Chem. A 2004, 108, 2851.
[177] B. O. Roos, V. Veryazov, P.-O. Widmark, Theor. Chem. Acc. Theory Comput. Model. (Theor. Chim. Acta) 2004, 111, 345.
[178] B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark, J. Phys. Chem. A 2005, 109, 6575.
[179] B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark, Chem. Phys. Lett. 2005, 409, 295.
[180] B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark, A. C. Borin, J. Phys. Chem. A 2008, 112, 11431.
[181] K. A. Peterson, Encyclopedia of Inorganic and Bioinorganic Chemistry, John Wiley & Sons, Ltd, Chichester, UK 2011.
[182] J. G. Hill, Int. J. Quantum Chem. 2013, 113, 21.
[183] F. Jensen, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2013, 3, 273.
[184] F. W. Averill, J. Chem. Phys. 1973, 59, 6412.
[185] B. Delley, D. E. Ellis, J. Chem. Phys. 1982, 76, 1949.
[186] B. Delley, J. Chem. Phys. 1990, 92, 508.
[187] J. D. Talman, Int. J. Quantum Chem. 1993, 48, 321.
[188] J. D. Talman, Phys. Rev. Lett. 2000, 84, 855.
[189] D. Andrae, Mol. Phys. 2001, 99, 327.
[190] J. D. Talman, Int. J. Quantum Chem. 2003, 95, 442.
[191] J. D. Talman, Phys. Rev. A 2010, 82, 052518.
[192] J. D. Talman, Mol. Phys. 2010, 108, 3289.
[193] A. D. Becke, R. M. Dickson, J. Chem. Phys. 1990, 92, 3610.
[194] T. Shiozaki, S. Hirata, Phys. Rev. A 2007, 76, 040503.
[195] V. Blum, R. Gehrke, F. Hanke, P. Havu, V. Havu, X. Ren, K. Reuter, M. Scheffler, Comput. Phys. Commun. 2009, 180, 2175.
[196] I. Y. Zhang, X. Ren, P. Rinke, V. Blum, M. Scheffler, New J. Phys. 2013, 15, 123033.
[197] K. Lejaeghere, G. Bihlmayer, T. Bjorkman, P. Blaha, S. Blugel, V. Blum, D. Caliste, I. E. Castelli, S. J. Clark, A. Dal Corso, S. de Gironcoli, T. Deutsch,
J. K. Dewhurst, I. di Marco, C. Draxl, M. du ak, O. Eriksson, J. A. Flores-Livas, K. F. Garrity, L. Genovese, P. Giannozzi, M. Giantomassi, S. Goedecker, X. Gonze,
O. Granas, E. K. U. Gross, A. Gulans, F. Gygi, D. R. Hamann, P. J. Hasnip, N. A. W. Holzwarth, D. Iu an, D. B. Jochym, F. Jollet, D. Jones, G. Kresse, K. Koepernik,
E. Kucukbenli, Y. O. Kvashnin, I. L. M. Locht, S. Lubeck, M. Marsman, N. Marzari, U. Nitzsche, L. Nordstrom, T. Ozaki, L. Paulatto, C. J. Pickard, W. Poelmans,
M. I. J. Probert, K. Refson, M. Richter, G. M. Rignanese, S. Saha, M. Scheffler, M. Schlipf, K. Schwarz, S. Sharma, F. Tavazza, P. Thunstrom, A. Tkatchenko,
M. Torrent, D. Vanderbilt, M. J. van Setten, V. van Speybroeck, J. M. Wills, J. R. Yates, G. X. Zhang, S. Cottenier, Science 2016, 351, aad3000.
[198] S. R. Jensen, S. Saha, J. A. Flores-Livas, W. Huhn, V. Blum, S. Goedecker, L. Frediani, J. Phys. Chem. Lett. 2017, 8, 1449.
[199] B. Delley, J. Chem. Phys. 2000, 113, 7756.
[200] R. C. Raffenetti, J. Chem. Phys. 1973, 58, 4452.
[201] F. Illas, J. M. Ricart, J. Rubio, P. S. Bagus, J. Chem. Phys. 1990, 93, 4982.
[202] C. W. Bauschlicher, P. R. Taylor, Theor. Chim. Acta 1993, 86, 13.
[203] F. Neese, E. F. Valeev, J. Chem. Theory Comput. 2011, 7, 33.
[204] G. Zeiss, W. Scott, N. Suzuki, D. Chong, S. Langhoff, Mol. Phys. 1979, 37, 1543.
[205] D. P. Chong, M. Grüning, E. J. Baerends, J. Comput. Chem. 2003, 24, 1582.
[206] V. Rossikhin, E. Voronkov, S. Okovytyy, T. Sergeieva, K. Kapusta, J. Leszczynski, Int. J. Quantum Chem. 2014, 114, 689.
[207] A. J. Sadlej, Chem. Phys. Lett. 1977, 47, 50.
[208] K. Wolin ski, A. J. Sadlej, Chem. Phys. Lett. 1979, 64, 51.
LEHTOLA 23 of 31

[209] B. O. Roos, A. J. Sadlej, Chem. Phys. 1985, 94, 43.


[210] K. Wolinski, B. O. Roos, A. J. Sadlej, Theor. Chim. Acta 1985, 68, 431.
[211] A. J. Sadlej, Collect. Czechoslov. Chem. Commun. 1988, 53, 1995.
[212] A. J. Sadlej, Theor. Chim. Acta 1991, 79, 123.
[213] A. J. Sadlej, Theor. Chim. Acta 1991, 81, 45.
[214] A. J. Sadlej, M. Urban, J. Mol. Struct. THEOCHEM 1991, 234, 147.
[215] V. Kellö, A. J. Sadlej, Theor. Chim. Acta 1996, 94, 93.
[216] I. Miadoková, V. Kellö, A. J. Sadlej, Theor. Chem. Acc. Theory, Comput. Model. (Theor. Chim. Acta) 1997, 96, 166.
[217] I. Černušák, V. Kellö, A. J. Sadlej, Collect. Czechoslov. Chem. Commun. 2003, 68, 211.
[218] A. Baranowska, M. Siedlecka, A. J. Sadlej, Theor. Chem. Acc. 2007, 118, 959.
[219] A. Baranowska, A. J. Sadlej, J. Comput. Chem. 2010, 31, 552.
[220] E. Voronkov, V. Rossikhin, S. Okovytyy, A. Shatckih, V. Bolshakov, J. Leszczynski, Int. J. Quantum Chem. 2012, 112, 2444.
[221] T. Koga, A. J. Thakkar, Can. J. Chem. 1992, 70, 362.
[222] P. Manninen, J. Vaara, J. Comput. Chem. 2006, 27, 434.
[223] S. Lehtola, J. Comput. Chem. 2015, 36, 335.
[224] S. Ikäläinen, P. Lantto, P. Manninen, J. Vaara, J. Chem. Phys. 2008, 129, 124102.
[225] S. Ikäläinen, P. Lantto, P. Manninen, J. Vaara, Phys. Chem. Chem. Phys. 2009, 11, 11404.
[226] S. Ikäläinen, M. Romalis, P. Lantto, J. Vaara, Phys. Rev. Lett. 2010, 105, 153001.
[227] P. Lantto, K. Jackowski, W. Makulski, M. Olejniczak, M. Jaszun ski, J. Phys. Chem. A 2011, 115, 10617.
[228] S. Ikäläinen, P. Lantto, J. Vaara, J. Chem. Theory Comput. 2012, 8, 91.
[229] L.-J. Fu, J. Vaara, J. Chem. Phys. 2013, 138, 204110.
[230] J. Vaara, M. Hanni, J. Jokisaari, J. Chem. Phys. 2013, 138, 104313.
[231] N. Abuzaid, A. M. Kantola, J. Vaara, Mol. Phys. 2013, 111, 1390.
[232] J. Vähäkangas, S. Ikäläinen, P. Lantto, J. Vaara, Phys. Chem. Chem. Phys. 2013, 15, 4634.
[233] J. Vähäkangas, P. Lantto, J. Vaara, J. Phys. Chem. C 2014, 118, 23996.
[234] M. Hanni, P. Lantto, M. Repiský, J. Mareš, B. Saam, J. Vaara, Phys. Rev. A 2017, 95, 032509.
[235] J. Lehtola, P. Manninen, M. Hakala, K. Hämäläinen, J. Chem. Phys. 2012, 137, 104105.
[236] S. Lehtola, P. Manninen, M. Hakala, K. Hämäläinen, J. Chem. Phys. 2013, 138, 044109.
[237] T. P. Rossi, S. Lehtola, A. Sakko, M. J. Puska, R. M. Nieminen, J. Chem. Phys. 2015, 142, 094114.
[238] F. Sim, D. R. Salahub, S. Chin, Int. J. Quantum Chem. 1992, 43, 463.
[239] T. Helgaker, M. Jaszun ski, K. Ruud, A. Górska, Theor. Chem. Acc. Theory, Comput. Model. (Theor. Chim. Acta) 1998, 99, 175.
[240] F. Jensen, J. Chem. Theory Comput. 2006, 2, 1360.
[241] F. Jensen, J. Chem. Theory Comput. 2008, 4, 719.
[242] A. A. Jarȩcki, E. R. Davidson, Chem. Phys. Lett. 1999, 300, 44.
[243] A. M. Simas, A. J. Thakkar, V. H. Smith, Int. J. Quantum Chem. 1982, 21, 419.
[244] P. A. Christiansen, E. A. McCullough, Chem. Phys. Lett. 1978, 56, 484.
[245] A. M. Simas, A. J. Thakkar, V. H. Smith, Int. J. Quantum Chem. 1983, 24, 527.
[246] R. Mastalerz, R. Lindh, M. Reiher, Chem. Phys. Lett. 2008, 465, 157.
[247] R. Mastalerz, P.-O. Widmark, B. O. Roos, R. Lindh, M. Reiher, J. Chem. Phys. 2010, 133, 144111.
[248] S. R. Jensen, T. Flå, D. Jonsson, R. S. Monstad, K. Ruud, L. Frediani, Phys. Chem. Chem. Phys. 2016, 18, 21145.
[249] A. H. Larsen, M. Vanin, J. J. Mortensen, K. S. Thygesen, K. W. Jacobsen, Phys. Rev. B 2009, 80, 195112.
[250] P.-O. Löwdin, Adv. Phys. 1956, 5, 1.
[251] E. Matito, J. Kobus, J. Styszyn ski, Chem. Phys. 2006, 321, 277.
[252] M. Melichercík, M. Pitoňák, V. Kellö, P. Hobza, P. Neogrády, J. Chem. Theory Comput. 2013, 9, 5296.
[253] M. Melichercík, D. Suchá, P. Neogrády, M. Pitoňák, Int. J. Quantum Chem. 2018, 118, e25580.
[254] L. M. Haines, J. N. Murrell, B. J. Ralston, D. J. Woodnutt, J. Chem. Soc. Faraday Trans. 2 1974, 70, 1794.
[255] B. J. Ralston, S. Wilson, J. Mol. Struct. THEOCHEM 1995, 341, 115.
[256] S. Wilson, J. Mol. Struct. THEOCHEM 1995, 357, 37.
[257] S. Wilson, Int. J. Quantum Chem. 1996, 60, 47.
[258] S. R. White, J. Chem. Phys. 2017, 147, 244102.
[259] H. Jansen, P. Ros, Chem. Phys. Lett. 1969, 3, 140.
[260] S. Boys, F. Bernardi, Mol. Phys. 1970, 19, 553.
[261] B. H. Wells, S. Wilson, Chem. Phys. Lett. 1983, 101, 429.
[262] J. M. L. Martin, J. P. François, R. Gijbels, Theor. Chim. Acta 1989, 76, 195.
[263] J. C. White, E. R. Davidson, J. Chem. Phys. 1990, 93, 8029.
[264] L. Turi, J. J. Dannenberg, J. Phys. Chem. 1993, 97, 2488.
[265] P. Valiron, I. Mayer, Chem. Phys. Lett. 1997, 275, 46.
[266] K. Mierzwicki, Z. Latajka, Chem. Phys. Lett. 2003, 380, 654.
[267] P. Pulay, Mol. Phys. 1969, 17, 197.
[268] O. C. Zienkiewicz, R. L. Taylor, J. Z. Zhu, The Finite Element Method: Its Basis and Fundamentals, Elsevier Butterworth-Heinemann, Oxford, UK 2013.
[269] D. Heinemann, B. Fricke, D. Kolb, Phys. Rev. A 1988, 38, 4994.
[270] B. Fornberg, Math. Comput. 1988, 51, 699.
[271] W. Hackbusch, Multi-Grid Methods and Applications. Springer Series in Computational Mathematics, Vol. 4, Springer, Berlin, Heidelberg 1985.
[272] T. Beck, Rev. Mod. Phys. 2000, 72, 1041.
[273] P. Maragakis, J. Soler, E. Kaxiras, Phys. Rev. B 2001, 64, 193101.
24 of 31 LEHTOLA

[274] C.-K. Skylaris, O. Diéguez, P. D. Haynes, M. C. Payne, Phys. Rev. B 2002, 66, 073103.
[275] E. J. Baerends, Chem. Phys. 1973, 2, 41.
[276] H. Sambe, R. H. Felton, J. Chem. Phys. 1975, 62, 1122.
[277] J. Mintmire, B. Dunlap, Phys. Rev. A 1982, 25, 88.
[278] O. Vahtras, J. Almlöf, M. Feyereisen, Chem. Phys. Lett. 1993, 213, 514.
[279] M. Feyereisen, G. Fitzgerald, A. Komornicki, Chem. Phys. Lett. 1993, 208, 359.
[280] C. Runge, Z. Math. Phys. 1901, 46, 224.
[281] J. Kobus, L. Laaksonen, D. Sundholm, Comput. Phys. Commun. 1996, 98, 346.
[282] J. Kobus, Comput. Phys. Commun. 2013, 184, 799.
[283] J. Kobus, L. Laaksonen, D. Sundholm, S. Lehtola, x2dhf – Two-Dimensional Finite Difference Hartree-Fock Program, http://github.com/susilehtola/-
x2dhf (accessed: May 2019).
[284] A. T. Patera, J. Comput. Phys. 1984, 54, 468.
[285] S. Lehtola, Int. J. Quantum Chem. 2019, e25945.
[286] I. J. Schoenberg, Q. Appl. Math. 1946, 4, 45.
[287] T. Gilbert, P. Bertoncini, Chem. Phys. Lett. 1974, 29, 569.
[288] T. L. Gilbert, P. J. Bertoncini, J. Chem. Phys. 1974, 61, 3026.
[289] T. L. Gilbert, J. Chem. Phys. 1975, 62, 1289.
[290] T. L. Gilbert, P. J. Bertoncini, J. Chem. Phys. 1975, 62, 4245.
[291] B. W. Shore, J. Phys. B At. Mol. Phys. 1973, 6, 1923.
[292] B. W. Shore, J. Chem. Phys. 1973, 58, 3855.
[293] B. W. Shore, J. Chem. Phys. 1973, 59, 6450.
[294] B. W. Shore, J. Phys. B At. Mol. Phys. 1974, 7, 2502.
[295] B. W. Shore, J. Chem. Phys. 1975, 63, 3835.
[296] H. Bachau, E. Cormier, P. Decleva, J. E. Hansen, F. Martín, Rep. Prog. Phys. 2001, 64, 1815.
[297] D. O. Harris, G. G. Engerholm, W. D. Gwinn, J. Chem. Phys. 1965, 43, 1515.
[298] A. S. Dickinson, P. R. Certain, J. Chem. Phys. 1968, 49, 4209.
[299] J. V. Lill, G. A. Parker, J. C. Light, Chem. Phys. Lett. 1982, 89, 483.
[300] J. C. Light, I. P. Hamilton, J. V. Lill, J. Chem. Phys. 1985, 82, 1400.
[301] D. Baye, P. H. Heenen, J. Phys. A Math. Gen. 1986, 19, 2041.
[302] J. V. Lill, G. A. Parker, J. C. Light, J. Chem. Phys. 1986, 85, 900.
[303] J. C. Light, T. Carrington, Adv. Chem. Phys. 2000, 114, 263.
[304] V. Szalay, J. Chem. Phys. 1996, 105, 6940.
[305] V. Szalay, G. Czakó, A.  Nagy, T. Furtenbacher, A. G. Császár, J. Chem. Phys. 2003, 119, 10512.
[306] V. Szalay, J. Chem. Phys. 2006, 125, 154115.
[307] V. Szalay, J. Phys. Chem. A 2013, 117, 7075.
[308] M. Cargo, R. G. Littlejohn, J. Chem. Phys. 2002, 117, 59.
[309] M. C. Cargo, R. G. Littlejohn, Phys. Rev. E 2002, 65, 026703.
[310] R. G. Littlejohn, M. Cargo, T. Carrington, K. A. Mitchell, B. Poirier, J. Chem. Phys. 2002, 116, 8691.
[311] R. G. Littlejohn, M. Cargo, J. Chem. Phys. 2002, 117, 37.
[312] R. G. Littlejohn, M. Cargo, J. Chem. Phys. 2002, 116, 7350.
[313] R. G. Littlejohn, M. Cargo, J. Chem. Phys. 2002, 117, 27.
[314] T. N. Rescigno, C. W. McCurdy, Phys. Rev. A 2000, 62, 032706.
[315] D. R. Hamann, M. Schlüter, C. Chiang, Phys. Rev. Lett. 1979, 43, 1494.
[316] L. Kleinman, D. M. Bylander, Phys. Rev. Lett. 1982, 48, 1425.
[317] D. Hamann, Phys. Rev. B 1989, 40, 2980.
[318] W. E. Pickett, Comput. Phys. Rep. 1989, 9, 115.
[319] D. Vanderbilt, Phys. Rev. B 1990, 41, 7892.
[320] K. Laasonen, A. Pasquarello, R. Car, C. Lee, D. Vanderbilt, Phys. Rev. B 1993, 47, 10142.
[321] M. Dolg, X. Cao, Chem. Rev. 2012, 112, 403.
[322] P. E. Blöchl, Phys. Rev. B 1994, 50, 17953.
[323] T. A. Arias, Rev. Mod. Phys. 1999, 71, 267.
[324] T. Torsti, T. Eirola, J. Enkovaara, T. Hakala, P. Havu, V. Havu, T. Höynälänmaa, J. Ignatius, M. Lyly, I. Makkonen, T. T. Rantala, J. Ruokolainen,
K. Ruotsalainen, E. Räsänen, H. Saarikoski, M. J. Puska, Phys. Status Solidi 2006, 243, 1016.
[325] Y. Saad, J. R. Chelikowsky, S. M. Shontz, SIAM Rev. 2010, 52, 3.
[326] L. Frediani, D. Sundholm, Phys. Chem. Chem. Phys. 2015, 17, 31357.
[327] F. A. Pahl, N. C. Handy, Mol. Phys. 2002, 100, 3199.
[328] S. Yamakawa, S.-A. Hyodo, Phys. Rev. B 2005, 71, 035113.
[329] T. N. Rescigno, D. A. Horner, F. L. Yip, C. W. McCurdy, Phys. Rev. A 2005, 72, 052709.
[330] F. L. Yip, C. W. McCurdy, T. N. Rescigno, Phys. Rev. A 2008, 78, 023405.
[331] S. A. Losilla, D. Sundholm, J. Chem. Phys. 2012, 136, 214104.
[332] F. L. Yip, C. W. McCurdy, T. N. Rescigno, Phys. Rev. A 2014, 90, 063421.
[333] J. L. Jerke, Y. Lee, C. J. Tymczak, J. Chem. Phys. 2015, 143, 064108.
[334] E. Solala, S. A. Losilla, D. Sundholm, W. Xu, P. Parkkinen, J. Chem. Phys. 2017, 146, 084102.
[335] P. Parkkinen, S. A. Losilla, E. Solala, E. A. Toivanen, W.-H. Xu, D. Sundholm, J. Chem. Theory Comput. 2017, 13, 654.
[336] P. Parkkinen, W.-H. Xu, E. Solala, D. Sundholm, J. Chem. Theory Comput. 2018, 14, 4237.
LEHTOLA 25 of 31

[337] B. Kanungo, V. Gavini, Phys. Rev. B 2017, 95, 035112.


[338] C. Marante, M. Klinker, I. Corral, J. González-Vázquez, L. Argenti, F. Martín, J. Chem. Theory Comput. 2017, 13, 499.
[339] E. Solala, P. Parkkinen, D. Sundholm, Comput. Phys. Commun. 2018, 232, 98.
[340] M. Braun, K. O. Obodo, Comput. Math. Appl. 2017, 74, 35.
[341] A. Gulans, A. Kozhevnikov, C. Draxl, Phys. Rev. B 2018, 97, 161105.
[342] S. R. White, J. W. Wilkins, M. P. Teter, Phys. Rev. B 1989, 39, 5819.
[343] J. E. Pask, B. M. Klein, C. Y. Fong, P. A. Sterne, Phys. Rev. B 1999, 59, 12352.
[344] J. Pask, B. Klein, P. Sterne, C. Fong, Comput. Phys. Commun. 2001, 135, 1.
[345] J. E. Pask, P. A. Sterne, Model. Simul. Mater. Sci. Eng. 2005, 13, R71.
[346] E. Tsuchida, M. Tsukada, Phys. Rev. B 1995, 52, 5573.
[347] E. Tsuchida, M. Tsukada, Solid State Commun. 1995, 94, 5.
[348] E. Tsuchida, M. Tsukada, Phys. Rev. B 1996, 54, 7602.
[349] E. Tsuchida, M. Tsukada, J. Phys. Soc. Jpn. 1998, 67, 3844.
[350] E. J. Bylaska, M. Holst, J. H. Weare, J. Chem. Theory Comput. 2009, 5, 937.
[351] L. Lehtovaara, V. Havu, M. Puska, J. Chem. Phys. 2009, 131, 054103.
[352] R. Alizadegan, K. J. Hsia, T. J. Martinez, J. Chem. Phys. 2010, 132, 034101.
[353] P. Suryanarayana, V. Gavini, T. Blesgen, K. Bhattacharya, M. Ortiz, J. Mech. Phys. Solids 2010, 58, 256.
[354] L. Lehtovaara, V. Havu, M. Puska, J. Chem. Phys. 2011, 135, 154104.
[355] P. Motamarri, M. Nowak, K. Leiter, J. Knap, V. Gavini, J. Comput. Phys. 2013, 253, 308.
[356] V. Schauer, C. Linder, J. Comput. Phys. 2013, 250, 644.
[357] Y. Maday, h−P Finite Element Approximation for Full-Potential Electronic Structure Calculations. in Partial Differential Equations: Theory, Control and
Approximation, Springer, Berlin, Heidelberg 2014, p. 349.
[358] E. Tsuchida, Y.-K. Choe, T. Ohkubo, Phys. Chem. Chem. Phys. 2015, 17, 31444.
[359] V. Schauer, C. Linder, Adv. Comput. Math. 2015, 41, 1035.
[360] D. Davydov, T. D. Young, P. Steinmann, Int. J. Numer. Methods Eng. 2016, 106, 863.
[361] J. Wang, T. L. Beck, J. Chem. Phys. 2000, 112, 9223.
[362] O. Cohen, L. Kronik, A. Brandt, J. Chem. Theory Comput. 2013, 9, 4744.
[363] T. Yanai, G. I. Fann, Z. Gan, R. J. Harrison, G. Beylkin, J. Chem. Phys. 2004, 121, 6680.
[364] T. Yanai, G. I. Fann, Z. Gan, R. J. Harrison, G. Beylkin, J. Chem. Phys. 2004, 121, 2866.
[365] T. Yanai, R. J. Harrison, N. C. Handy, Mol. Phys. 2005, 103, 413.
[366] H. Sekino, Y. Maeda, T. Yanai, R. J. Harrison, J. Chem. Phys. 2008, 129, 034111.
[367] H. Sekino, Y. Yokoi, R. J. Harrison, J. Phys. Conf. Ser. 2012, 352, 012014.
[368] F. A. Bischoff, R. J. Harrison, E. F. Valeev, J. Chem. Phys. 2012, 137, 104103.
[369] F. A. Bischoff, E. F. Valeev, J. Chem. Phys. 2013, 139, 114106.
[370] J. S. Kottmann, S. Höfener, F. A. Bischoff, Phys. Chem. Chem. Phys. 2015, 17, 31453.
[371] T. Yanai, G. I. Fann, G. Beylkin, R. J. Harrison, Phys. Chem. Chem. Phys. 2015, 17, 31405.
[372] M. Braun, J. Comput. Appl. Math. 2012, 236, 4840.
[373] F. A. Bischoff, J. Chem. Phys. 2014, 141, 184105.
[374] F. A. Bischoff, J. Chem. Phys. 2014, 141, 184106.
[375] N. A. Modine, G. Zumbach, E. Kaxiras, Phys. Rev. B 1997, 55, 10289.
[376] E. Artacho, E. Anglada, O. Diéguez, J. D. Gale, A. García, J. Junquera, R. M. Martin, P. Ordejón, J. M. Pruneda, D. Sánchez-Portal, J. M. Soler, J. Phys.
Condens. Matter 2008, 20, 064208.
[377] S. Ryu, S. Choi, K. Hong, W. Y. Kim, J. Chem. Phys. 2016, 144, 094101.
[378] J. Heyd, G. E. Scuseria, M. Ernzerhof, J. Chem. Phys. 2003, 118, 8207.
[379] J. Heyd, G. E. Scuseria, M. Ernzerhof, J. Chem. Phys. 2006, 124, 219906.
[380] I. Duchemin, F. Gygi, Comput. Phys. Commun. 2010, 181, 855.
[381] N. M. Boffi, M. Jain, A. Natan, J. Chem. Theory Comput. 2016, 12, 3614.
[382] J. B. Krieger, Y. Li, G. J. Iafrate, Phys. Lett. A 1990, 146, 256.
[383] J. B. Krieger, Y. Li, G. J. Iafrate, Phys. Rev. A 1992, 45, 101.
[384] J. Kim, K. Hong, S. Choi, W. Y. Kim, Bull. Korean Chem. Soc. 2015, 36, 998.
[385] J. Lim, S. Choi, J. Kim, W. Y. Kim, J. Chem. Phys. 2016, 145, 224309.
[386] J. Kim, K. Hong, S.-Y. Hwang, S. Ryu, S. Choi, W. Y. Kim, Phys. Chem. Chem. Phys. 2017, 19, 10177.
[387] J. Kim, S. Kang, J. Lim, S.-Y. Hwang, W. Y. Kim, Comput. Phys. Commun. 2018, 230, 21.
[388] E. R. Davidson, J. Comput. Phys. 1975, 17, 87.
[389] R. J. Harrison, G. I. Fann, T. Yanai, Z. Gan, G. Beylkin, J. Chem. Phys. 2004, 121, 11587.
[390] L. Frediani, E. Fossgaard, T. Flå, K. Ruud, Mol. Phys. 2013, 111, 1143.
[391] S. Cools, B. Reps, W. Vanroose, SIAM J. Sci. Comput. 2014, 36, B367.
[392] J. Ackermann, B. Erdmann, R. Roitzsch, J. Chem. Phys. 1994, 101, 7643.
[393] J. Ackermann, Phys. Rev. A 1995, 52, 1968.
[394] S. Lehtola, Int. J. Quantum Chem. 2019, e25944.
[395] L. Sturesson, C. F. Fischer, Comput. Phys. Commun. 1993, 74, 432.
[396] C. Roothaan, Rev. Mod. Phys. 1960, 32, 179.
[397] S. Huzinaga, Phys. Rev. 1960, 120, 866.
[398] S. Huzinaga, Phys. Rev. 1961, 122, 131.
[399] S. Huzinaga, J. Chem. Phys. 1969, 51, 3971.
26 of 31 LEHTOLA

[400] E. R. Davidson, Chem. Phys. Lett. 1973, 21, 565.


[401] C. F. Fischer, Can. J. Phys. 1963, 41, 1895.
[402] C. F. Fischer, Comput. Phys. Commun. 1970, 1, 151.
[403] C. F. Fischer, Comput. Phys. Commun. 1972, 4, 107.
[404] L. Chernysheva, N. Cherepkov, V. Radojevic, Comput. Phys. Commun. 1976, 11, 57.
[405] C. F. Fischer, J. Comput. Phys. 1978, 27, 221.
[406] F. Biegler-König, J. Hinze, J. Comput. Phys. 1986, 67, 290.
[407] C. Froese Fischer, Comput. Phys. Commun. 1987, 43, 355.
[408] C. Froese Fischer, Comput. Phys. Commun. 1991, 64, 431.
[409] D. Andrae, J. Hinze, Int. J. Quantum Chem. 1997, 63, 65.
[410] C. Froese Fischer, Comput. Phys. Commun. 2000, 128, 635.
[411] C. Froese Fischer, G. Tachiev, G. Gaigalas, M. R. Godefroid, Comput. Phys. Commun. 2007, 176, 559.
[412] C. Froese Fischer, Can. J. Phys. 1968, 46, 2336.
[413] C. Froese Fischer, At. Data Nucl. Data Tables 1972, 4, 301.
[414] C. Froese Fischer, At. Data Nucl. Data Tables 1973, 12, 87.
[415] T. Koga, E. Shibata, A. J. Thakkar, Theor. Chim. Acta 1995, 91, 47.
[416] T. Koga, S. Watanabe, A. J. Thakkar, Int. J. Quantum Chem. 1995, 54, 261.
[417] T. Koga, A. J. Thakkar, J. Phys. B At. Mol. Opt. Phys. 1996, 29, 2973.
[418] S. L. Saito, At. Data Nucl. Data Tables 2009, 95, 836.
[419] F. W. King, J. Mol. Struct. THEOCHEM 1997, 400, 7.
[420] M. V. Ivanov, J. Phys. B At. Mol. Opt. Phys. 1994, 27, 4513.
[421] M. V. Ivanov, P. Schmelcher, Phys. Rev. A 1998, 57, 3793.
[422] M. V. Ivanov, P. Schmelcher, Phys. Rev. A 1999, 60, 3558.
[423] M. V. Ivanov, P. Schmelcher, Phys. Rev. A 2000, 61, 022505.
[424] M. V. Ivanov, P. Schmelcher, Eur. Phys. J. D 2001, 14, 279.
[425] M. V. Ivanov, P. Schmelcher, J. Phys. B At. Mol. Opt. Phys. 2001, 34, 2031.
[426] M. V. Ivanov, P. Schmelcher, Adv. Quantum Chem. 2001, 40, 361.
[427] J. R. Flores, E. Clementi, V. Sonnad, J. Chem. Phys. 1989, 91, 7030.
[428] J. R. Flores, E. Clementi, V. Sonnad, Chem. Phys. Lett. 1989, 163, 198.
[429] A. Altenberger-Siczek, T. L. Gilbert, J. Chem. Phys. 1976, 64, 432.
[430] J. L. Gázquez, H. J. Silverstone, J. Chem. Phys. 1977, 67, 1887.
[431] H. J. Silverstone, D. P. Carroll, D. M. Silver, J. Chem. Phys. 1978, 68, 616.
[432] D. P. Carroll, H. J. Silverstone, R. M. Metzger, J. Chem. Phys. 1979, 71, 4142.
[433] C. F. Fischer, W. Guo, J. Comput. Phys. 1990, 90, 486.
[434] C. Froese Fischer, W. Guo, Z. Shen, Int. J. Quantum Chem. 1992, 42, 849.
[435] J. E. Hansen, M. Bentley, H. W. van der Hart, M. Landtman, G. M. S. Lister, Y.-T. Shen, N. Vaeck, Phys. Scr. 1993, T47, 7.
[436] C. Froese Fischer, T. Brage, AIP Conf. Proc. 1995, 347, 139.
[437] Y. Qiu, C. F. Fischer, J. Comput. Phys. 1999, 156, 257.
[438] S. L. Saito, Theor. Chem. Acc Theory Comput. Model. (Theor Chim. Acta) 2003, 109, 326.
[439] C. Froese Fischer, Adv. At. Mol. Opt. Phys. 2008, 55, 235.
[440] C. F. Fischer, Comput. Phys. Commun. 2011, 182, 1315.
[441] J. Delhalle, M. Defranceschi, Int. J. Quantum Chem. 1987, 32, 425.
[442] M. Defranceschi, A. Lahmam-Bennani, J. Electron Spectrosc. Relat. Phenom. 1989, 48, 1.
[443] M. Defranceschi, J. Delhalle, Eur. J. Phys. 1990, 11, 172.
[444] M. Defranceschi, L. De Windt, J. G. Fripiat, J. Delhalle, J. Mol. Struct. THEOCHEM 1992, 258, 179.
[445] L. De Windt, J. G. Fripiat, J. Delhalle, M. Defranceschi, J. Mol. Struct. THEOCHEM 1992, 254, 145.
[446] L. De Windt, M. Defranceschi, J. Dehalle, Int. J. Quantum Chem. 1993, 45, 609.
[447] L. Windt, M. Defranceschi, J. Delhalle, Theor. Chim. Acta 1993, 86, 487.
[448] L. DeWindt, P. Fischer, M. Defranceschi, J. Delhalle, J. G. Fripiat, J. Comput. Phys. 1994, 111, 266.
[449] P. Fischer, M. Defranceschi, Appl. Comput. Harmon. Anal. 1994, 1, 232.
[450] C. F. Fischer, M. Godefroid, T. Brage, P. Jönsson, G. Gaigalas, J. Phys. B At. Mol. Opt. Phys. 2016, 49, 182004.
[451] J. R. Flores, Phys. Rev. A 1992, 46, 6063.
[452] J. R. Flores, Chem. Phys. Lett. 1992, 195, 377.
[453] J. R. Flores, J. Chem. Phys. 1993, 98, 5642.
[454] J. R. Flores, P. Redondo, Int. J. Quantum Chem. 1993, 45, 563.
[455] J. R. Flores, P. Redondo, J. Phys. B At. Mol. Opt. Phys. 1993, 26, 2251.
[456] J. R. Flores, P. Redondo, J. Comput. Chem. 1994, 15, 782.
[457] J. Flores, Chem. Phys. Lett. 1997, 270, 427.
[458] J. R. Flores, D. Kolb, J. Phys. B At. Mol. Opt. Phys. 1999, 32, 779.
[459] J. R. Flores, K. Jankowski, R. Słupski, Collect. Czechoslov. Chem. Commun. 2003, 68, 240.
[460] J. R. Flores, R. Słupski, K. Jankowski, P. Malinowski, J. Chem. Phys. 2004, 121, 12334.
[461] K. Jankowski, K. Nowakowski, R. Słupski, J. R. Flores, Int. J. Quantum Chem. 2004, 99, 277.
[462] J. R. Flores, R. Słupski, K. Jankowski, J. Chem. Phys. 2006, 124, 104107.
[463] J. R. Flores, Int. J. Quantum Chem. 2008, 108, 2172.
LEHTOLA 27 of 31

[464] D. Sundholm, J. Olsen, P.-Å. Malmqvist, B. O. Roos, Numerical MCSCF in One and Two Dimensions. in Numerical Determination of the Electronic
Structure of Atoms, Diatomic and Polyatomic Molecules, Springer Netherlands, Dordrecht 1989, p. 329.
[465] A.-M. Mårtensson-Pendrill, S. A. Alexander, L. Adamowicz, N. Oliphant, J. Olsen, P. Öster, H. M. Quiney, S. Salomonson, D. Sundholm, Phys. Rev. A
1991, 43, 3355.
[466] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1990, 171, 53.
[467] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1994, 217, 451.
[468] J. Olsen, L. G. M. Pettersson, D. Sundholm, J. Phys. B At. Mol. Opt. Phys. 1994, 27, 5575.
[469] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1995, 233, 115.
[470] D. Sundholm, J. Phys. B At. Mol. Opt. Phys. 1995, 28, L399.
[471] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1991, 182, 497.
[472] D. Sundholm, J. Olsen, S. A. Alexander, J. Chem. Phys. 1992, 96, 5229.
[473] D. Sundholm, J. Olsen, M. Godefroid, G. Van Meulebeke, Phys. Rev. A 1993, 48, 3606.
[474] D. Sundholm, J. Olsen, Phys. Rev. A 1990, 42, 2614.
[475] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1991, 177, 91.
[476] D. Sundholm, J. Olsen, J. Chem. Phys. 1991, 94, 5051.
[477] D. Sundholm, J. Olsen, Phys. Rev. A 1990, 42, 1160.
[478] D. Sundholm, J. Olsen, Nucl. Phys. A 1991, 534, 360.
[479] D. Sundholm, J. Olsen, J. Phys. Chem. 1992, 96, 627.
[480] D. Sundholm, J. Olsen, Chem. Phys. Lett. 1992, 198, 526.
[481] D. Sundholm, J. Olsen, Phys. Rev. Lett. 1992, 68, 927.
[482] D. Sundholm, J. Olsen, J. Chem. Phys. 1993, 98, 7152.
[483] D. Sundholm, J. Olsen, Phys. Rev. A 1993, 47, 2672.
[484] D. Sundholm, J. Olsen, Phys. Rev. A 1994, 49, 3453.
[485] J. Olsen, D. Sundholm, Chem. Phys. Lett. 1994, 226, 17.
[486] M. Tokman, D. Sundholm, P. Pyykkö, J. Olsen, Chem. Phys. Lett. 1997, 265, 60.
[487] M. Tokman, D. Sundholm, P. Pyykkö, Chem. Phys. Lett. 1998, 291, 414.
[488] D. Sundholm, Phys. Rev. A 1999, 59, 3355.
[489] V. Kellö, A. J. Sadlej, P. Pyykkö, D. Sundholm, M. Tokman, Chem. Phys. Lett. 1999, 304, 414.
[490] J. Bieron , P. Pyykkö, D. Sundholm, V. Kellö, A. J. Sadlej, Phys. Rev. A 2001, 64, 052507.
[491] D. Sundholm, P. Pyykkö, Mol. Phys. 2018, 116, 1682.
[492] D. Sundholm, J. Olsen, J. Chem. Phys. 1993, 98, 3999.
[493] D. Sundholm, J. Olsen, J. Chem. Phys. 1993, 99, 6222.
[494] J. Olsen, D. Sundholm, Chem. Phys. Lett. 1998, 288, 282.
[495] D. Sundholm, J. Chem. Phys. 1995, 102, 4895.
[496] M. Braun, Phys. Rev. A 2002, 65, 033415.
[497] D. Engel, G. Wunner, Phys. Rev. A 2008, 78, 032515.
[498] D. Engel, M. Klews, G. Wunner, Comput. Phys. Commun. 2009, 180, 302.
[499] C. Schimeczek, D. Engel, G. Wunner, Comput. Phys. Commun. 2014, 185, 1498.
[500] D. Hochstuhl, M. Bonitz, Phys. Rev. A 2012, 86, 053424.
[501] F. Malet, P. Gori-Giorgi, Phys. Rev. Lett. 2012, 109, 246402.
[502] A. D. Becke, J. Chem. Phys. 2013, 138, 074109.
[503] A. D. Becke, J. Chem. Phys. 2013, 138, 161101.
[504] E. R. Johnson, J. Chem. Phys. 2013, 139, 074110.
[505] L. Gagliardi, D. G. Truhlar, G. Li Manni, R. K. Carlson, C. E. Hoyer, J. L. Bao, Acc. Chem. Res. 2017, 50, 66.
[506] L. Li, T. E. Baker, S. R. White, K. Burke, Phys. Rev. B 2016, 94, 245129.
[507] E. R. Johnson, A. D. Becke, J. Chem. Phys. 2017, 146, 211105.
[508] N. Q. Su, C. Li, W. Yang, Proc. Natl. Acad. Sci. USA 2018, 115(39), 9678.
[509] P. Schwerdtfeger, ChemPhysChem 2011, 12, 3143.
[510] M. Fuchs, M. Scheffler, Comput. Phys. Commun. 1999, 119, 67.
[511] N. Holzwarth, A. Tackett, G. Matthews, Comput. Phys. Commun. 2001, 135, 329.
[512] M. J. Oliveira, F. Nogueira, Comput. Phys. Commun. 2008, 178, 524.
[513] D. Langreth, J. Perdew, Phys. Rev. B 1980, 21, 5469.
[514] D. Andrae, R. Brodbeck, J. Hinze, Int. J. Quantum Chem. 2001, 82, 227.
[515] T. Ozaki, M. Toyoda, Comput. Phys. Commun. 2011, 182, 1245.
[516] Z. Romanowski, Model. Simul. Mater. Sci. Eng. 2008, 16, 015003.
[517] Z. Romanowski, Mol. Phys. 2009, 107, 1339.
[518] Z. Romanowski, Int. J. Quantum Chem. 2009, 110, 1793.
[519] Z. Romanowski, Model. Simul. Mater. Sci. Eng. 2009, 17, 045001.
[520] E. A. McCullough, Chem. Phys. Lett. 1974, 24, 55.
[521] E. A. McCullough, J. Chem. Phys. 1975, 62, 3991.
[522] E. A. McCullough, Comput. Phys. Rep. 1986, 4, 265.
[523] K. Rüdenberg, J. Chem. Phys. 1951, 19, 1459.
[524] E. A. McCullough, J. Chem. Phys. 1975, 63, 5050.
[525] P. A. Christiansen, E. A. McCullough, Chem. Phys. Lett. 1977, 51, 468.
[526] P. A. Christiansen, E. A. McCullough, Chem. Phys. Lett. 1979, 63, 570.
28 of 31 LEHTOLA

[527] P. A. Christiansen, E. A. McCullough, J. Chem. Phys. 1977, 67, 1877.


[528] L. Adamowicz, E. A. McCullough, J. Chem. Phys. 1981, 75, 2475.
[529] E. A. McCullough, J. Phys. Chem. 1982, 86, 2178.
[530] E. A. McCullough, J. Chem. Phys. 1981, 75, 1579.
[531] L. Adamowicz, E. A. McCullough, J. Phys. Chem. 1984, 88, 2045.
[532] E. A. McCullough, Mol. Phys. 1981, 42, 943.
[533] E. A. McCullough, J. Morrison, K. W. Richman, Faraday Symp. Chem. Soc. 1984, 19, 165.
[534] L. Adamowicz, R. J. Bartlett, E. A. McCullough, Phys. Rev. Lett. 1985, 54, 426.
[535] L. Adamowicz, J. C. Ellenbogen, E. A. McCullough, Int. J. Quantum Chem. 1986, 30, 617.
[536] K. W. Richman, Z.-g. Shi, E. A. McCullough, Chem. Phys. Lett. 1987, 141, 186.
[537] K. W. Richman, E. A. McCullough, J. Chem. Phys. 1987, 87, 5050.
[538] H. Partridge, K. W. Richman, E. A. McCullough, Theor. Chim. Acta 1988, 74, 151.
[539] S. N. Beck, E. A. McCullough, D. Feller, Chem. Phys. Lett. 1990, 175, 629.
[540] L. Adamowicz, R. J. Bartlett, Int. J. Quantum Chem. 1984, 26, 213.
[541] L. Adamowicz, R. J. Bartlett, Phys. Rev. A 1988, 37, 1.
[542] L. Adamowicz, Very Accurate Calculations for Diatomic, Neutral and Anionic Systems with Numerical Orbitals. in Numerical Determination of the Elec-
tronic Structure of Atoms, Diatomic and Polyatomic Molecules, Springer Netherlands, Dordrecht 1989, p. 177.
[543] L. Adamowicz, Mol. Phys. 1988, 65, 1047.
[544] D. M. Chipman, J. Chem. Phys. 1989, 91, 5455.
[545] L. Laaksonen, P. Pyykkö, D. Sundholm, Int. J. Quantum Chem. 1983, 23, 309.
[546] L. Laaksonen, P. Pyykkö, D. Sundholm, Int. J. Quantum Chem. 1983, 23, 319.
[547] A. D. Becke, J. Chem. Phys. 1982, 76, 6037.
[548] A. D. Becke, J. Chem. Phys. 1983, 78, 4787.
[549] A. D. Becke, Int. J. Quantum Chem. 1985, 27, 585.
[550] A. D. Becke, J. Chem. Phys. 1986, 84, 4524.
[551] A. D. Becke, Phys. Rev. A 1986, 33, 2786.
[552] L. Laaksonen, P. Pyykkö, D. Sundholm, Chem. Phys. Lett. 1983, 96, 1.
[553] L. Laaksonen, F. Müller-Plathe, G. H. F. Diercksen, J. Chem. Phys. 1988, 89, 4903.
[554] F. Müller-Plathe, L. Laaksonen, Chem. Phys. Lett. 1989, 160, 175.
[555] F. Müller-Plathe, G. H. F. Diercksen, L. Laaksonen, Application of the Two-dimensional Fully-Numerical RHF Method to Open-Shell Hydrides. in
Numerical Determination of the Electronic Structure of Atoms, Diatomic and Polyatomic Molecules, Springer Netherlands, Dordrecht 1989, p. 311.
[556] L. Laaksonen, P. Pyykkö, D. Sundholm, Comput. Phys. Rep. 1986, 4, 313.
[557] P. Pyykkö, D. Sundholm, L. Laaksonen, Mol. Phys. 1987, 60, 597.
[558] P. Pyykkö, G. H. F. Diercksen, F. Müller-Plathe, L. Laaksonen, Chem. Phys. Lett. 1987, 134, 575.
[559] D. Sundholm, P. Pyykkö, L. Laaksonen, Mol. Phys. 1985, 56, 1411.
[560] L. Laaksonen, D. Sundholm, P. Pyykkö, Chem. Phys. Lett. 1984, 105, 573.
[561] L. Laaksonen, D. Sundholm, P. Pyykkö, Int. J. Quantum Chem. 1985, 27, 601.
[562] D. Sundholm, P. Pyykkö, L. Laaksonen, Finnish Chem. Lett. 1985, 1985, 51.
[563] D. Sundholm, P. Pyykkö, L. Laaksonen, Mol. Phys. 1985, 55, 627.
[564] E. Baerends, P. Vernooijs, A. Rozendaal, P. Boerrigter, M. Krijn, D. Feil, D. Sundholm, J. Mol. Struct. THEOCHEM 1985, 133, 147.
[565] D. Sundholm, P. Pyykkö, L. Laaksonen, A. J. Sadlej, Chem. Phys. Lett. 1984, 112, 1.
[566] D. Sundholm, P. Pyykkö, L. Laaksonen, A. J. Sadlej, Chem. Phys. 1986, 101, 219.
[567] G. H. Diercksen, A. J. Sadlej, D. Sundholm, P. Pyykkö, Chem. Phys. Lett. 1988, 143, 163.
[568] K. Nordlund, N. Runeberg, D. Sundholm, Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 1997, 132, 45.
[569] J. Kobus, AIP Conf. Proc. 2012, 1504, 189.
[570] J. Kobus, Comput. Phys. Commun. 1994, 78, 247.
[571] J. Kobus, D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 1994, 27, 2867.
[572] J. Kobus, D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 1994, 27, 5139.
[573] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 1995, 86, 1315.
[574] D. Moncrieff, J. Kobus, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 1995, 28, 4555.
[575] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 1997, 92, 1015.
[576] D. Moncrieff, J. Kobus, S. Wilson, Mol. Phys. 1998, 93, 713.
[577] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 1999, 96, 1559.
[578] J. Kobus, D. Moncrieff, S. Wilson, Phys. Rev. A 2000, 62, 062503.
[579] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 2000, 98, 401.
[580] J. Kobus, H. M. Quiney, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 2001, 34, 2045.
[581] J. Kobus, D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 2001, 34, 5127.
[582] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 2001, 99, 315.
[583] J. Kobus, D. Moncrieff, S. Wilson, Mol. Phys. 2002, 100, 499.
[584] J. Kobus, D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 2004, 37, 571.
[585] J. Kobus, D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 2007, 40, 877.
[586] V. N. Glushkov, J. Kobus, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 2008, 41, 205102.
[587] J. Styszynski, J. Kobus, Chem. Phys. Lett. 2003, 369, 441.
[588] J. Styszynski, Chem. Phys. Lett. 2000, 317, 351.
[589] J. Kobus, Comput. Lett. 2007, 3, 71.
LEHTOLA 29 of 31

[590] J. Kobus, Phys. Rev. A 2015, 91, 022501.


[591] R. A. Wilhelm, E. Gruber, J. Schwestka, R. Kozubek, T. I. Madeira, J. P. Marques, J. Kobus, A. V. Krasheninnikov, M. Schleberger, F. Aumayr, Phys. Rev.
Lett. 2017, 119, 103401.
[592] F. Jensen, J. Chem. Phys. 2002, 116, 7372.
[593] F. Jensen, J. Chem. Phys. 2003, 118, 2459.
[594] F. Jensen, Theor. Chem. Acc. 2005, 113, 187.
[595] T. Grabo, E. K. U. Gross, Int. J. Quantum Chem. 1997, 64, 95.
[596] T. Grabo, T. Kreibich, E. K. U. Gross, Mol. Eng. 1997, 7, 27.
[597] S. Liu, R. G. Parr, Phys. Rev. A 1996, 53, 2211.
[598] S. Liu, V. Karasiev, R. López-Boada, F. De Proft, Int. J. Quantum Chem. 1998, 69, 513.
[599] E. V. Ludeña, V. Karasiev, R. López-Boada, E. Valderrama, J. Maldonado, J. Comput. Chem. 1999, 20, 155.
[600] V. Karasiev, J. Mol. Struct. THEOCHEM 1999, 493, 21.
[601] V. V. Karasiev, E. V. Ludeña, A. N. Artemyev, Phys. Rev. A 2000, 62, 062510.
[602] V. V. Karasiev, E. V. Ludeña, Phys. Rev. A 2002, 65, 062510.
[603] V. V. Karasiev, E. V. Ludeña, Phys. Rev. A 2002, 65, 032515.
[604] V. V. Karasiev, J. Chem. Phys. 2003, 118, 8576.
[605] V. Karasiev, S. Trickey, F. E. Harris, Chem. Phys. 2006, 330, 216.
[606] A. Halkier, S. Coriani, Chem. Phys. Lett. 2001, 346, 329.
[607] A. K. Roy, A. J. Thakkar, Chem. Phys. Lett. 2002, 362, 428.
[608] F. Weigend, F. Furche, R. Ahlrichs, J. Chem. Phys. 2003, 119, 12753.
[609] S. Shahbazian, M. Zahedi, Theor. Chem. Acc. 2005, 113, 152.
[610] J. Pruneda, E. Artacho, Phys. Rev. B 2004, 70, 035106.
[611] V. Kuzmin, Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2006, 249, 13.
[612] V. Kuzmin, Surf. Coat. Technol. 2007, 201, 8388.
[613] C. B. Madsen, L. B. Madsen, Phys. Rev. A 2006, 74, 023403.
[614] T. G. Williams, N. J. DeYonker, A. K. Wilson, J. Chem. Phys. 2008, 128, 044101.
[615] L. B. Madsen, O. I. Tolstikhin, T. Morishita, Phys. Rev. A 2012, 85, 053404.
[616] L. B. Madsen, F. Jensen, O. I. Tolstikhin, T. Morishita, Phys. Rev. A 2013, 87, 013406.
[617] R. Saito, O. I. Tolstikhin, L. B. Madsen, T. Morishita, At. Data Nucl. Data Tables 2015, 103-104, 4.
[618] L. B. Madsen, F. Jensen, A. I. Dnestryan, O. I. Tolstikhin, Phys. Rev. A 2017, 96, 013423.
[619] A. S. Kornev, B. A. Zon, Laser Phys. 2014, 24, 115302.
[620] T. Endo, A. Matsuda, M. Fushitani, T. Yasuike, O. I. Tolstikhin, T. Morishita, A. Hishikawa, Phys. Rev. Lett. 2016, 116, 163002.
[621] W. Schulze, D. Kolb, Chem. Phys. Lett. 1985, 122, 271.
[622] D. Heinemann, D. Kolb, B. Fricke, Chem. Phys. Lett. 1987, 137, 180.
[623] D. Heinemann, B. Fricke, D. Kolb, Chem. Phys. Lett. 1988, 145, 125.
[624] D. Heinemann, B. Fricke, D. Kolb, Finite Element Method for the Accurate Solution of Diatomic Molecules. in Numerical Determination of the Elec-
tronic Structure of Atoms, Diatomic and Polyatomic Molecules, Springer Netherlands, Dordrecht 1989, p. 275.
[625] D. Heinemann, A. Rosén, B. Fricke, Phys. Scr. 1990, 42, 692.
[626] D. Heinemann, A. Rosén, B. Fricke, Chem. Phys. Lett. 1990, 166, 627.
[627] D. Heinemann, A. Rosén, Theor. Chim. Acta 1993, 85, 249.
[628] A. v. Kopylow, D. Heinemann, D. Kolb, J. Phys. B At. Mol. Opt. Phys. 1998, 31, 4743.
[629] A. v. Kopylow, D. Kolb, Chem. Phys. Lett. 1998, 295, 439.
[630] D. Sundholm, Comput. Phys. Commun. 1988, 49, 409.
[631] K. Davstad, J. Comput. Phys. 1992, 99, 33.
[632] A. Makmal, S. Kümmel, L. Kronik, J. Chem. Theory Comput. 2009, 5, 1731.
[633] A. Makmal, S. Kümmel, L. Kronik, J. Chem. Theory Comput. 2011, 7, 2665.
[634] J. C. Morrison, T. Wolf, B. Bialecki, G. Fairweather, L. Larson, Mol. Phys. 2000, 98, 1175.
[635] J. C. Morrison, K. Steffen, B. Pantoja, A. Nagaiya, J. Kobus, T. Ericsson, Commun. Comput. Phys. 2016, 19, 632.
[636] Morrison, J. C.; Kobus, J. Advances in Quantum Chemistry, 1st; Elsevier Inc., 2018; 76; 103–116.
[637] A. N. Artemyev, E. V. Ludeña, V. V. Karasiev, A. J. Hernández, J. Comput. Chem. 2004, 25, 368.
[638] Y. V. Vanne, A. Saenz, J. Phys. B At. Mol. Opt. Phys. 2004, 37, 4101.
[639] L. Tao, C. W. McCurdy, T. N. Rescigno, Phys. Rev. A 2009, 79, 012719.
[640] L. Tao, C. W. McCurdy, T. N. Rescigno, Phys. Rev. A 2010, 82, 023423.
[641] X. Guan, K. Bartschat, B. I. Schneider, Phys. Rev. A 2010, 82, 041404.
[642] X. Guan, K. Bartschat, B. I. Schneider, Phys. Rev. A 2011, 83, 043403.
[643] O. I. Tolstikhin, T. Morishita, L. B. Madsen, Phys. Rev. A 2011, 84, 053423.
[644] D. J. Haxton, K. V. Lawler, C. W. McCurdy, Phys. Rev. A 2011, 83, 063416.
[645] H. R. Larsson, S. Bauch, L. K. Sørensen, M. Bonitz, Phys. Rev. A 2016, 93, 013426.
[646] L. Yue, S. Bauch, L. B. Madsen, Phys. Rev. A 2017, 96, 043408.
[647] B. Zhang, J. Yuan, Z. Zhao, Comput. Phys. Commun. 2015, 194, 84.
[648] G. Berthier, M. Defranceschi, J. Navaza, M. Suard, G. Tsoucaris, J. Mol. Struct. THEOCHEM 1985, 120, 343.
[649] P. Fischer, M. Defranceschi, J. Delhalle, Numer. Math. 1992, 63, 67.
[650] P. Fischer, L. De Windt, M. Defranceschi, J. Delhalle, J. Chem. Phys. 1993, 99, 7888.
[651] G. Berthier, M. Defranceschi, J. Navaza, G. Tsoucaris, Int. J. Quantum Chem. 1997, 63, 451.
[652] M. Defranceschi, M. Suard, G. Berthier, Int. J. Quantum Chem. 1984, 25, 863.
30 of 31 LEHTOLA

[653] M. Defranceschi, J. Delhalle, Phys. Rev. B 1986, 34, 5862.


[654] J. Gräfenstein, D. Izotov, D. Cremer, J. Chem. Phys. 2007, 127, 214103.
[655] E. R. Johnson, A. D. Becke, C. D. Sherrill, G. A. DiLabio, J. Chem. Phys. 2009, 131, 034111.
[656] S. E. Wheeler, K. N. Houk, J. Chem. Theory Comput. 2010, 6, 395.
[657] N. Mardirossian, M. Head-Gordon, J. Chem. Theory Comput. 2013, 9, 4453.
[658] M. P. Bircher, P. Lopez-Tarifa, U. Rothlisberger, J. Chem. Theory Comput. 2019, 15(1), 557.
[659] S. Lehtola, HelFEM – Finite Element Methods for Electronic Structure Calculations on Small Systems, http://github.com/susilehtola/HelFEM
(accessed: May 2019).
[660] S. Lehtola, C. Steigemann, M. J. Oliveira, M. A. Marques, SoftwareX 2018, 7, 1.
[661] S. Lehtola, M. Dimitrova, D. Sundholm, Mol. Phys. 2019, 1.
[662] J. Hinze, C. C. J. Roothaan, Prog. Theor. Phys. Suppl. 1967, 40, 37.
[663] J. Hinze, J. Chem. Phys. 1973, 59, 6424.
[664] S. R. White, Phys. Rev. Lett. 1992, 69, 2863.
[665] S. Wilson, D. Moncrieff, J. Chem. Phys. 1996, 105, 3336.
[666] T. Helgaker, W. Klopper, H. Koch, J. Noga, J. Chem. Phys. 1997, 106, 9639.
[667] A. K. Wilson, T. H. Dunning Jr. , J. Chem. Phys. 1997, 106, 8718.
[668] A. Halkier, T. Helgaker, P. Jørgensen, W. Klopper, H. Koch, J. Olsen, A. K. Wilson, Chem. Phys. Lett. 1998, 286, 243.
[669] D. Moncrieff, S. Wilson, J. Phys. B At. Mol. Opt. Phys. 1996, 29, 6009.
[670] C. Schwartz, Phys. Rev. 1962, 126, 1015.
[671] M. R. Nyden, G. A. Petersson, J. Chem. Phys. 1981, 75, 1843.
[672] R. N. Hill, J. Chem. Phys. 1985, 83, 1173.
[673] D. Feller, J. Chem. Phys. 1992, 96, 6104.
[674] D. Feller, J. Chem. Phys. 1993, 98, 7059.
[675] K. A. Peterson, D. E. Woon, T. H. Dunning, J. Chem. Phys. 1994, 100, 7410.
[676] J. M. Martin, Chem. Phys. Lett. 1996, 259, 669.
[677] A. Halkier, T. Helgaker, P. Jørgensen, W. Klopper, J. Olsen, Chem. Phys. Lett. 1999, 302, 437.
[678] N. B. Balabanov, K. A. Peterson, J. Phys. Chem. A 2003, 107, 7465.
[679] L. Bytautas, K. Ruedenberg, J. Chem. Phys. 2004, 121, 10905.
[680] F. Jensen, Theor. Chem. Acc. 2005, 113, 267.
[681] D. A. Dixon, W. A. de Jong, K. A. Peterson, T. B. McMahon, J. Phys. Chem. A 2005, 109, 4073.
[682] K. A. Peterson, C. Puzzarini, Theor. Chem. Acc. 2005, 114, 283.
[683] D. W. Schwenke, J. Chem. Phys. 2005, 122, 014107.
[684] A. Karton, J. M. L. Martin, Theor. Chem. Acc. 2006, 115, 330.
[685] A. Karton, E. Rabinovich, J. M. L. Martin, B. Ruscic, J. Chem. Phys. 2006, 125, 144108.
[686] D. Bakowies, J. Chem. Phys. 2007, 127, 084105.
[687] E. C. Lee, D. Kim, P. Jurecka, P. Tarakeshwar, P. Hobza, K. S. Kim, J. Phys. Chem. A 2007, 111, 3446.
[688] A. J. C. Varandas, J. Chem. Phys. 2007, 126, 244105.
[689] J. G. Hill, K. A. Peterson, G. Knizia, H.-J. Werner, J. Chem. Phys. 2009, 131, 194105.
[690] D. Feller, K. A. Peterson, J. Grant Hill, J. Chem. Phys. 2011, 135, 044102.
[691] K. A. Peterson, D. Feller, D. A. Dixon, Theor. Chem. Acc. 2012, 131, 1079.
[692] D. Feller, J. Chem. Phys. 2013, 138, 074103.
[693] M. Okoshi, T. Atsumi, H. Nakai, J. Comput. Chem. 2015, 36, 1075.
[694] P. D. Mezei, G. I. Csonka, A. Ruzsinszky, J. Chem. Theory Comput. 2015, 11, 3961.
[695] J. S. Boschen, D. Theis, K. Ruedenberg, T. L. Windus, J. Phys. Chem. A 2017, 121, 836.
[696] R. Olivares-Amaya, W. Hu, N. Nakatani, S. Sharma, J. Yang, G. K.-L. Chan, J. Chem. Phys. 2015, 142, 034102.

AUTHOR BIOGRAPHY

Susi Lehtola completed his PhD in theoretical physics in 2013 at the University of Helsinki. After a one-year postdoc with
Hannes Jónsson at Aalto University and a three-year postdoc with Martin Head-Gordon at University of California,
Berkeley, he was awarded the Löwdin postdoctoral award at the Sanibel meeting 2017, as well as a three-year Academy
of Finland Postdoctoral Researcher scholarship in 2017 for methods development at the University of Helsinki.

How to cite this article: Lehtola S. A review on non-relativistic, fully numerical electronic structure calculations on atoms and diatomic
molecules. Int J Quantum Chem. 2019;119:e25968. https://doi.org/10.1002/qua.25968
LEHTOLA 31 of 31

A P P E N D I X : O R B I T A L S I N L I N E A R S Y M M E T RY

To ensure general readability of the article, I remind here that orbitals in linear molecules with m = 0, m = ±1, m = ±2, and m = ±3 are commonly
known as σ, π, δ, and φ orbitals, respectively. Orbitals beyond φ are not necessary for SCF calculations in the known periodic table, although they
may be necessary for post-HF approaches. A fully filled σ orbital can fit two electrons, whereas fully filled π, δ, and φ orbitals can fit four electrons,
as both the m = | m | and m = − | m | subshells fit two electrons.
It is worthwhile to point out here that a diatomic σ orbital includes the m = 0 components from all AOs: in addition to the ns orbitals, contribu-
tions may also arise from atomic npz, ndz2, nfz3, and so forth orbitals which describe polarization effects in LCAO calculations; similar remarks can
be also made about the π, δ, and φ orbitals.
The other way around, an atomic ns orbital yields a σ orbital, whereas an atomic np orbital yields one σ and two π orbitals, corresponding to
the m = 0 and m = ±1 components of the function. Similarly, atomic nd orbitals yield one σ, two π, and two δ functions, whereas atomic nf orbitals
yield one σ, two π, two δ, and two φ orbitals.

You might also like