PyFLOSIC: Python-based Fermi–Löwdin orbital self-interaction correction Sebastian Schwalbe,1, a) Lenz Fiedler,1 Jakob Kraus,1 Jens Kortus,1, b) Kai Trepte,2 and Susi Lehtola3, c) 1)Institute of Theoretical Physics, TU Bergakademie Freiberg, Leipziger Str. 23, D-09599 Freiberg, Germany 2)Department of Physics, Central Michigan University, Mount Pleasant, MI 48859, USA 3)Department of Chemistry, University of Helsinki, P.O. Box 55 (A. I. Virtasen aukio 1), FI-00014 University of Helsinki, Finland We present PyFLOSIC, an open-source, general-purpose Python implementation of the Fermi–Löwdin orbital self-interaction correction (FLO-SIC), which is based on the Python simulation of chemistry frame- work (PySCF) electronic structure and code. Thanks to PySCF, PyFLOSIC can be used with any kind of -type basis set, various kinds of radial and angular quadrature grids, and all exchange-correlation functionals within the local density approximation (LDA), generalized-gradient approx- imation (GGA), and meta-GGA provided in the Libxc and XCFun libraries. A central aspect of FLO-SIC are Fermi-orbital descriptors, which are used to estimate the self-interaction correction. Importantly, they can be initialized automatically within PyFLOSIC; they can also be optimized with an interface to the atomic simulation environment, a Python library that provides a variety of powerful gradient-based algorithms for geometry optimization. Although PyFLOSIC has already facilitated applications of FLO-SIC to chemical studies, it offers an excellent starting point for further developments in FLO-SIC approaches, thanks to its use of a high-level programming language and pronounced modularity.

I. INTRODUCTION code written primarily in Python. As a general-purpose quantum chemistry program PySCF already contains a 7,14 The continuously growing availability of open-source vast number of methods. For example, self-consistent software has given rise to an ongoing paradigm shift in field (SCF) approaches like Hartree–Fock and density 15,16 quantum chemical software development. At variance to functional theory (DFT) are available with various the traditional model of monolithic programs, where new choices for spin treatment, density fitting techniques, as algorithms need to be re-implemented separately from well as numerical quadrature methods. Second-order or- 17 scratch in each code, present-day computational science bital optimization is also available in PySCF. is proceeding in leaps and bounds thanks to the thriving Since PySCF is highly modular, and its parts can be ecosystems of small projects dedicated to solving well- imported like any other Python module, it is rather defined sub-problems1–6 that can be easily combined to easy to add new functionality to it or to combine it with form a code that is greater than the sum of its parts; see existing software,14 as has already been demonstrated e.g. refs. 7 and 8 and references therein for further dis- by various authors.2,5,6 This is also the strategy that cussion. The efficiency of the new development models was adopted for PyFLOSIC. Before PyFLOSIC, the comes from not having to "re-invent the wheel" within FLO-SIC method has only been available in the refer- every program; instead, standardized, modular tools de- ence implementation within the Naval Research Labora- signed to solve specific problems can be reused across the tory Molecular Orbital Library (NRLMOL).18–26 There board. This kind of cleaner code design and efficient code is also a FLO-SIC implementation in a developer branch reuse has enabled fast development, and streamlined the of the NWChem program,27,28 but to the best of our adoption of new methodologies, since features added to knowledge this implementation is not currently publicly any one component become straightforwardly available available. in everything built on top of them. Similarly to NRLMOL and NWChem, PySCF— This work describes PyFLOSIC, an open-source, and therefore PyFLOSIC—uses a Gaussian-type or- Python-based implementation of the Fermi–Löwdin bital (GTO) basis set.7,14 However, in contrast to 9–12 orbital self-interaction correction (FLO-SIC). The the NRLMOL program, NWChem and PySCF (and core routines of PyFLOSIC were developed during PyFLOSIC) are able to routinely handle basis functions arXiv:1905.02631v4 [physics.comp-ph] 18 Jul 2020 Lenz Fiedler’s master’s thesis,13 and the code is freely with high angular momentum. The most-commonly used available on GitHub (https://github.com/pyflosic/ all-electron as well as effective core potential GTO ba- pyflosic). PyFLOSIC builds on the Python simula- sis sets are already available within PySCF, and addi- tion of chemistry framework (PySCF),7,14 which is an tional ones can be downloaded from, e.g., the Basis Set open-source electronic structure and quantum chemistry Exchange29 and parsed in from several formats. The importance of the tractability of calculations within large basis sets is underlined by recent fully nu- a)Electronic mail: pyfl[email protected] merical benchmark studies that have shown that the b)Electronic mail: [email protected] reproduction of atomization energies even within DFT c)Electronic mail: [email protected].fi may require several shells of polarization functions.30–33 2

PySCF contains fast routines for the evaluation of Fermi–Löwdin orbitals (Subsection II D). The schemes molecular integrals via the Libcint library,1 which are used in PyFLOSIC for optimizing the FLO-SIC density also used in PyFLOSIC. Moreover, PyFLOSIC inherits are discussed in Subsection II E. The code is showcased OpenMP parallelization from PySCF, enabling the effi- in Section III, starting with example inputs for running a cient use of multi-core computation architectures.7 By calculation on tetracyanoethylene. The structure of the enabling the use of large basis sets in FLO-SIC calcula- tetracyanoethylene example follows that of the calcula- tions thanks to the fast elementary routines in PySCF, tion: first, the FODs are initialized (Subsection III A), PyFLOSIC enables reliable computational studies of and then the electron density and the FODs are opti- even molecules for which very large and flexible basis mized (Subsection III B). A discussion on repeated cal- 30,31,33 sets are required, such as SO2 and SF6, as will culations follows in Subsection III C. As a practical ap- also be demonstrated later in this work. Recently de- plication of the code, in Subsection III D we perform an veloped powerful approaches to handle significant linear in-depth basis set convergence study of the atomization 34,35 dependencies in the underlying molecular basis set energies of SO2 and SF6 that have been found to be chal- are also available in PyFLOSIC through PySCF, en- lenging cases in the literature, as was already discussed abling accurate FLO-SIC calculations even in pathologi- above. The article concludes in a summary and outlook cally over-complete basis sets.33–35 in Section IV. Atomic units are used throughout the pa- Equally importantly, again at variance to the pre- per, if not stated otherwise. vious implementations of FLO-SIC that feature in- house implementations of only a handful of density functionals, PyFLOSIC contains (again via PySCF) II. FERMI–LÖWDIN ORBITAL SELF-INTERACTION interfaces to the Libxc3 and XCFun36 libraries of CORRECTION exchange-correlation functionals, which grant access to a wide range of hundreds of local density ap- We will adopt the notation introduced by Lehtola proximations (LDAs), generalized-gradient approxima- and Jónsson 38 in the following. However, in contrast tions (GGAs), meta-GGAs, hybrid functionals, non-local to ref. 38 vectors and matrices are distinguished in the correlation functionals, and range-separated hybrids.7 At presently used notation: all vectors are expressed in bold, the moment, all LDAs, GGAs, and meta-GGAs pro- non-italic letters (a), and all matrices are given in bold, vided through PySCF can be used in PyFLOSIC; italic letters (R). for a thorough list of the functionals implemented in Libxc and XCFun, we refer to the libraries’ re- spective documentations.37 Custom linear combinations A. Self-interaction correction in density functional theory of exchange-correlation functionals can also be used within PySCF, further expanding the capabilities of DFT has become one of the standard methods in com- PyFLOSIC. putational materials science, condensed matter physics, As an add-on, PyFLOSIC inherits all of the important as well as chemistry thanks to its combination of reason- features of PySCF. Because PyFLOSIC is implemented able accuracy with computational efficiency.39,40 How- as a collection of Python modules like PySCF,14 those ever, currently available density functional approxima- already familiar with PySCF do not have to learn a new tions (DFAs) are well-known to fail in a number of package-specific input format to run FLO-SIC calcula- situations.40–42 Although the description of systems ex- tions with PyFLOSIC. Users of PyFLOSIC can also hibiting significant static correlation remains an open rely on the full power of Python—one of today’s most problem, posing limitations on studies of many transition commonly used and taught programming languages that metal complexes and molecules with stretched bonds, has a massive and vibrant community—to fulfill their challenges exist also in the absence of static correlation. every need. The availability of all Python language For instance, localized states and negatively charged tools within the input script allows for elaborate work species such as F– are incorrectly described by pure (i.e., schemes, e.g., the combination of calculation, evaluation, local or semi-local) density functionals. and plotting routines.14 Calculations can also be done These shortcomings come from the fact that DFAs in- interactively in the Python interpreter shell.14 clude spurious interactions of electrons with themselves, Having introduced and motivated PyFLOSIC, we will known as the self-interaction error, which causes the elec- next discuss the theory of FLO-SIC in detail in Sec- tron density to delocalize. Delocalization error has been tion II. We will start out by motivating the use of self- long identified as a key issue in the application of DFT interaction corrections in general (Subsection II A), con- onto the study of chemical systems; for instance, the bar- tinue by presenting Perdew and Zunger’s approach in rier height of the simplest hydrogen abstraction reaction specific (Subsection II B), introduce Fermi–Löwdin or- H + H2 H2 + H already poses a challenge, as many bitals to undertake the self-interaction correction follow- functionals←−→ overestimate the stability of the intermediate 43 ing Perdew and Zunger’s prescription (Subsection II C), H3 state. The error can often be significantly reduced and discuss the mandatory initialization and optimiza- by including a fraction of exact exchange as in (possibly tion of the Fermi-orbital descriptors that parametrize the range-separated) hybrid functionals; however, this does 3 not fully remove the problems with self-interaction. B. Perdew–Zunger self-interaction correction

The total energy in the KS-DFT formalism is given by An exchange-correlation functional that fulfills three constraints may be free of self-interaction for many-body α β α β α β EKS[n , n ] = Ts[n , n ]+V [n]+J[n]+K[n , n ] , (1) systems. It needs to (i) be free of self-interactions for all α β α β one-electron densities, (ii) provide the correct piecewise where EKS[n , n ] is the total KS energy, Ts[n , n ] is linearity (PWL),44 and (iii) recover the correct asymp- the kinetic energy of the non-interacting system, V [n] is totic limit of the potential. As no DFAs that fulfill the external potential energy, J[n] is the Coulomb energy, all of these constraints completely are available, self- K[nα, nβ] is the exchange-correlation energy, and nσ is interaction corrections (SICs) aim to rectify some of the the electron density for spin σ, with α and β representing aforementioned issues for the available DFAs. Since the spin up and spin down, respectively. early formulation of SIC,45 SICs have been shown to be The KS functional can be minimized with a SCF able to significantly reduce the errors of the underlying procedure,89 in which the KS-Fock matrix exchange-correlation functional for the first two of the σ σ aforementioned issues, namely the delocalization error FKS = Hcore + J + K , (2) (i) and the lack of piecewise linearity (ii).11,38,46–55 The lack of piecewise linearity is the reason the highest oc- is iteratively diagonalized, using appropriate measures to cupied molecular orbital (HOMO) energies from Kohn– ensure the procedure becomes convergent. The first term Sham (KS) DFAs yield poor estimates for the ionization of equation (2), Hcore, is the core Hamiltonian that arises potential as IP = εHOMO; in contrast, the so-called from the first two terms of equation (1) as Koopmans-compliant− (KC) functionals50,56 can deliver nuclei especially accurate results for IPs. 1 2 X ZA Hcore = , (3) −2∇ − r rA A | − | Among the multitude of current implementations of where Z is the nuclear charge of atom A at position r . SIC, most employ real-valued orbitals (RSIC).57–64 How- A A The last two terms in equation (2), J and Kσ, are the ever, it has been shown that SIC based on complex- Coulomb and exchange-correlation matrix, respectively. valued orbitals (CSIC) has several advantages;48,53,65–67 As only the two first terms in equation (1)—or equiv- in fact, Lehtola, Head-Gordon, and Jónsson 68 showed alently, the first term of equation (2)—are necessary some time ago that the use of complex-valued orbitals is for exactness for one-electron systems,45 the exact KS actually mandatory to properly minimize the SIC func- exchange-correlation functional must perfectly cancel out tional. (Implementations of RSIC and CSIC that sup- the Coulomb term for any one-electron density nσ with port arbitrary basis sets and exchange-correlation func- 1 spin σ tionals similarly to PyFLOSIC are freely available in the 69,70 ERKALE program. ) K[nσ, 0] = J[nσ]. (4) 1 − 1 DFAs violate this condition, leading to a spurious self- Another variant of SIC is the FLO-SIC approach men- interaction (SI) energy tioned in Section I, which has been the subject of var- ious studies during recent years,9–12,27,28,55,71–86 and is σ σ σ ESI[n ] = K[n , 0] + J[n ]. (5) also the topic of the present work; it will be discussed in 1 1 1 depth below in Subsections II C, II D, and II E. It is first The Perdew–Zunger self-interaction correction45 (PZ- worth mentioning, however, that in spite of the many SIC) enforces the correct behavior of the DFA by explic- successes of SIC (some of which were referenced above), itly removing the SI energy orbital by orbital, leading to there are also several results that show SIC degrad- a corrected total energy functional ing the performance of higher-rung exchange-correlation 50,62,87 α β functionals. This is the so-called "paradox of self- EPZ = EKS[n , n ] + ESIC interaction correction" discussed in the comprehensive N σ summary of Perdew et al. 52 , which remains a challenge α β X X σ = EKS[n , n ] ESI[n ] , (6) − i for the future. For instance, it has been pointed out that σ i the Perdew–Zunger approach45 does not completely elim- inate one-electron self-interaction;88 however, the path where N σ is the number of electrons with spin σ. for rectifying the remaining error is still unclear. Despite Because of this correction, the PZ-SIC Hamiltonian de- these theoretical paradoxes, self-interaction corrections pends not only on the total spin densities, but also on the have been shown to be useful in many cases, which is individual orbital densities, at variance to standard KS- why we will not consider this problem further in this DFT. Due to this, the unitary invariance of KS-DFT89 work and will instead carry on with the Perdew–Zunger is broken, and the functional is no longer invariant to approach. orbital rotations within the occupied space. 4

The optimal PZ-SIC orbitals minimize EPZ. It has C. Fermi–Löwdin orbitals been shown that localized orbitals generally deliver lower self-interaction corrected total energies E than delo- PZ At variance to the direct minimization of EPZ with 45,90,91 calized ones. Hence, the best way to minimize orbital rotations, Pederson and coworkers proposed us- equation (6) is to start out with localized orbitals (see ing Fermi–Löwdin orbitals (FLOs) together with PZ-SIC, ref. 92 and references therein for discussion on this giving rise to the FLO-SIC method.9–12 FLOs were orig- 93 topic) produced, e.g., by the Foster–Boys, Edmiston– inally introduced by Luken and coworkers in a series 94 95 Ruedenberg, or Pipek–Mezey method or generaliza- of papers that dealt with the exchange or Fermi hole 96 tions thereof, and then iteratively rotate the orbitals in the context of many-electron wave functions.97–100 In until the derivative of the energy, the occupied-occupied the FLO approach, one starts by building Fermi orbitals orbital gradient38 σ (FOs) φFO,i via σ σ σ σ σ T σ σ σ cFO = C R , (9) κij = (cj ) (fi fj )ci (7) − σ σ where cFO and C contain the FO and KS orbital coef- σ vanishes; κσ = 0 is known as the Pederson condition. ficients, respectively. The transformation matrix R is σ σ σ defined as Here, fi = J(pi ) + K(pi ) is the orbital Fock matrix σ σ (without the one-electron part) that arises from the den- σ ψj ai σ Rji = ph | i . (10) sity matrix pi corresponding to the i:th occupied orbital nσ(aσ) of spin σ i σ where ai denotes a position eigenstate localized at the | i σ σ σ σ σ T so-called Fermi-orbital descriptor (FOD) ai , and ψj are pi = ci (ci ) , (8) KS orbitals. The FOs are normalized, but they do not form an orthonormal set in general. As a consequence, σ where ci are the coefficients for the i:th optimal orbital. Luken proposed using Löwdin’s method of symmetric or- In addition to the occupied-occupied rotations, the gra- thonormalization101 to end up with an orthonormalized dient for which was given in equation (7), the occupied- σ set of Fermi–Löwdin orbitals (FLOs) φk as virtual rotations also have to be optimized, as discussed σ σ σ σ −1/2 σ T in ref. 68. c = cFO[T (Q ) (T ) ], (11) Practical calculations based on orbital rotations, such where cσ holds the FLO coefficients, while T σ and Qσ as those with the ERKALE program, pursue minimiza- contain the eigenvectors and eigenvalues of the FO over- tion only to a finite numerical threshold; that is, until lap matrix, respectively. the orbital gradient is small but still non-zero. Still, the The usefulness of the FOs and FLOs arises from σ σ minimization of EPZ by orbital rotations is arduous due the property of the FO. It has the value a φ = h i | FO,ii to slow convergence of the orbital optimization. More- p σ σ n (ai ) at its descriptor, meaning that all the electron over, as was already mentioned in the Introduction, the density (for a given spin channel) at this point in space correct minimization of equation (6) has been shown to comes from a single orbital. Thus, the FOs are localized require complex-valued orbitals and to exhibit a plethora around their corresponding FODs, in contrast to the typ- 68 of local minima. ically delocalized KS orbitals. As the orthonormalization The asymptotic scaling of the orbital rotation approach mixes the FOs, the FLOs are slightly less localized than is determined by the calculation of the orbital gradient the original non-orthogonal FOs. in equation (7). Assuming K basis functions, cσ and Another key feature of the FLO approach is that since σ N σ σ fi i=1 are all K K matrices, with K N . Disre- the FLOs are uniquely determined by equations (10) and { } × σ ≥ σ garding the cost to compute the N orbital Fock matri- (11) for a given set of FODs ai and a set of occupied ces (which should overall scale linearly with system size orbitals Cσ, the FLO-SIC approach turns out to restore in an optimal implementation due to the localized nature the unitary invariance of KS-DFT: rotations of the oc- of the optimal orbitals), the evaluation of κ carries an it- cupied orbitals in Cσ are countermanded by an inverse erative (N σK3) cost, while orbital optimization using rotation occurring in Rσ in equation (10). This feature the approachO of ref. 68 carries an iterative (K3) cost like means that the optimization of the FODs and that of the conventional Kohn–Sham density functionalO theory. Still, density matrix can be decoupled. the bottleneck in practical calculations is typically in the evaluation of the orbital Fock matrices, and alternative approaches that avoid the irrelevant virtual-virtual block D. Initialization and optimization of Fermi-orbital descriptors can be pursued in the case K N σ. However, due to the presence of many local minima and saddle point solu- σ tions, the approach of ref. 68 recommends the use of sta- The FODs ai formally introduced in equation (10) bility analysis; since there are ((N σ)2) rotation angles, are a key feature of the FLO-SIC approach. In FLO- stability analysis employing iterativeO diagonalization car- SIC, the optimization of the N σ(N σ 1)/2 orbital ro- ries a ((N σ)4) cost. tation angles within the occupied space− (twice that if O 5 imaginary rotations are also included as in the CSIC ref. 38. approach) is replaced by the optimization of 3N σ FOD Surprisingly, the optimization of this small number of coordinates.9–12,27 Although full minimization of the PZ- FOD coordinates turns out to be at least as challenging SIC energy functional used in FLO-SIC requires opti- as the optimization of the far more numerous possible mization of the FODs, the interesting part about the orbital rotations. As previously shown in refs. 10 and 11, model is that since plausible FODs can be generated in an the construction of the FOD derivatives from the orbital automatic fashion,84 qualitative calculations may be per- Fock matrices leads to ((N σ)4) scaling and ((N σ)3) formed without having to optimize the FODs, although data storage cost; thisO computational scalingO and data the predictive power of such calculations is limited in storage behavior is also realized within PyFLOSIC. The the same way as the use of, e.g., Foster–Boys orbitals gradient of the energy with respect to the FODs can be for evaluating the self-interaction correction criticized in expressed as10,11

σ N  σ   σ  ∂EPZ ∂ESIC X k,σ ∂φk σ σ ∂φk = = ε φ + φ , (12) ∂ aσ ∂ aσ kl ∂ aσ l l ∂ aσ i i k,l i i

k,σ where the elements of the Lagrange multiplier matrix εkl portant part of FLO-SIC calculations. FOD opti- are defined by mization is carried out automatically in PyFLOSIC through an interface to the atomic simulation environ- k,σ σ T σ σ 108 εij = (ci ) fk cj . (13) ment (ASE ), which has ample algorithms for geom- − etry optimization with forces, including conjugate gra- As always, the FOD optimization starts from some dients (CG), the (limited memory) Broyden–Fletcher– initial values for the FODs. As was discussed in Sub- Goldfarb–Shanno scheme ((L-)BFGS),109–116 and the section II B, minimization of the PZ functional by or- fast inertial relaxation engine (FIRE).117 In order to use σ bital rotation techniques is typically started from lo- ASE, the FOD gradient ∂ESIC/∂ ai is simply expressed σ σ calized orbitals. Now, since FLOs are highly localized in terms of a fictitious force f = ∂EPZ/∂ a acting on i − i around their FODs, this initialization can also be car- the FODs, and the optimization is continued until some ried over to FLO-SIC by retrieving initial FODs from the numerical threshold is reached for these FOD forces. centroids of the localized orbitals from the Foster–Boys, Edmiston–Ruedenberg, or (generalized) Pipek–Mezey.84 Such a procedure is implemented in the Python center of mass (PyCOM) module within PyFLOSIC, which E. Optimization of the density with FLO-SIC employs the second-order orbital localizer17 in PySCF. Several other ways to initialize the FODs have also As was discussed in Subsection II C, the optimization been recently discussed in ref. 84. They include an elec- of the energy functional used in FLO-SIC can be split tronic force field which yields quasi-classical positions into two sub-problems: optimization of the density ma- for the electrons to which the FODs can be assigned trix for a fixed set of FODs, and optimization of the FODs (PyEFF); using Lewis-like bonding information to place for a fixed density matrix, which was already discussed FODs accordingly (PyLEWIS); or a Thomson problem- in Subsection II D. Although the PZ-SIC Hamiltonian like procedure of obtaining the FODs from Monte-Carlo σ FPZ based on equation (6) implies that each orbital ex- minimization of a distribution of point charges under the periences a different potential, all the orbital-dependent restriction of a certain bond order (fodMC). The initial Hamiltonians can be cast in a common form by employing FODs from any of these alternative generators can be a unified Hamiltonian approach.38,90,118 Assuming such easily read in by PyFLOSIC. Please refer to ref. 84 for a unified Hamiltonian, the SCF equations that determine further details on automatic FOD initialization. the optimal occupied orbitals for PZ-SIC can be written Regardless of the procedure used to initialize the as FODs, the initial FOD geometries should always be vi- sualized; an example will be discussed below in Sec- σ σ = ( σ + σ ) σ = σ σ , (14) tion III. By visualizing the resulting electronic geome- FPZC FKS FSIC C SC E try, one should check whether it is in reasonable agree- ment with Lewis102 or Linnett double-quartet (LDQ) where S is the basis set’s overlap matrix, and Eσ is a theory,103–106 as good-natured FODs should generally diagonal matrix that holds the energy eigenvalues. correspond to these theories.84,107 There are at least two different ways to construct σ 38,90,118 The gradients of these initial FODs are usually non- a unified SIC Hamiltonian FSIC, and both negligible, which is why FOD optimization is an im- Hamiltonians presented here have been implemented in 6

PyFLOSIC. The first SIC Hamiltonian space of occupied orbitals, as indicated by the identi- fier OO. This is equivalent to ignoring the frequently N σ σ 1 X σ σ σ σ small off-diagonal Lagrange multipliers that ensure or- F = (f p S + Sp f ) (15) 38 SIC,OO −2 i i i i bital orthonormality. The second SIC Hamiltonian is i only contains operators that project to and from the

N σ X F σ = S (pσf σpσ + vσf σpσ + pσf σvσ)S, (16) SIC,OOOV − i i i i i i i i

where the virtual space projector is defined as only two lines of code are needed to generate a reason- able initial guess for the FODs, not counting the import virtual of the necessary Python modules. σ X σ σ T v = Ci (Ci ) . (17) Alternatively to the use of PyCOM, the FODs can i also be initialized in PyFLOSIC by reading in the out- σ put of any of the other methods described in ref. 84. The Therefore, FSIC,OOOV also allows for projections between necessary code for the read-in is shown in figure 4 that the occupied and virtual orbital spaces, leading to the will be discussed in more detail below. identifier OOOV; this Hamiltonian does not assume a Optimized FODs have been shown to carry bonding diagonal Lagrange multiplier matrix. The second SIC information in the context of Lewis/LDQ theory, which Hamiltonian can also be written as a block matrix in the is discussed in detail in ref. 84. For example, a simple occupied and virtual spaces as FOD bond order (BOFOD) can be evaluated via  σ σ  σ FSIC,OO FSIC,OV BO = m /n (19) FSIC,OOOV = σ σ FOD electrons centers FSIC,VO FSIC,VV  σ σ  just by counting the number of FODs m between the FSIC,OO FSIC,OV = σ , (18) bonded atoms, and then dividing by the number of atoms FSIC,VO 0 n that partake in this bond. The initial set of FODs as the operator includes no terms that couple virtual or- shown in figure 2 appears to be reasonable, as it is in ex- bitals together. cellent agreement with Lewis/LDQ theory; it can thus be expected that the optimization of these FODs will yield useful results as well. It is, however, important to note that since the exchange-correlation functional affects the III. EXAMPLE CALCULATIONS WITH PYFLOSIC electron density, the reasonableness of the PyCOM guess may depend on the system and the functional, which is The features of PyFLOSIC are first showcased using why we recommend to always visualize the initial FODs the tetracyanoethylene (C2(CN)4) molecule. In addition before running calculations. As the FODs change during to specifying the nuclear geometry for the molecule, as the optimization, the final geometries should be inspected is necessary for any electronic structure calculation, the as well. initial FODs have to be specified in order to perform a FLO-SIC calculation, as was discussed in Subsection II D; that is, an electronic geometry needs to be specified as B. FOD and density optimization well. Having established the necessary nuclear and elec- tronic geometries, the FLO-SIC calculation can begin. A. Automatic generation of electronic geometry As was discussed in Section II, FLO-SIC has two classes of degrees of freedom: the occupied orbitals, i.e., the elec- Besides the versatile core functionality inherited from tron density, and the FODs. Both should be optimized PySCF, PyFLOSIC has unique features specific to in order to minimize the PZ-SIC energy functional; the FLO-SIC calculations. The first of these is the automatic workflow for such FLO-SIC calculations is illustrated in generation of the electronic geometry, i.e., the FODs. figure 3. Note that the self-interaction correction is re- The necessary code to obtain initial FODs with the Py- evaluated at every SCF iteration with the up-to-date den- COM approach is presented in figure 1, and the resulting sity matrix, ensuring the self-consistency of the orbitals electronic geometry is visualized in figure 2. Note that and the correction. 7 from ase.io import read However, in some cases one may want to keep the den- from pycom import pycom_guess sity or the FODs fixed; this may be useful, e.g., for ex- ploratory FLO-SIC calculations. The default mode in # Load nuclear information PyFLOSIC is to optimize both the FODs and the den- ase_nuclei = read(’Conformer3D_CID_12635. sity, but fixed-density and fixed-FOD calculations are sdf’) also supported. # StartFOD generation pycom_guess(ase_nuclei=ase_nuclei, As was discussed in Section II, PyFLOSIC optimizes charge =0, the FODs via ASE, whereas the electron density is op- spin =0, timized through a unified Hamiltonian approach. Exam- xc=’pbesol’, ple code for the optimization of both the density and the basis =’pc-0’, FODs with PyFLOSIC is shown in figure 4. For this ex- method =’fb’) ample, the FODs generated with PyCOM (see figure 2) are used as starting points for the FOD optimization, and FIG. 1. Generation of initial FODs using PyCOM. After σ the default unified SIC Hamiltonian FSIC,OOOV is used importing the needed modules, the nuclear information is for the density optimization within the SCF approach. read from an .sdf file containing data for tetracyanoethylene Note that again, only two function calls are needed to (C (CN) ), which has been downloaded from the 2 4 PubChem run the calculation. FOD optimization can be analyzed database.119 Next, PyCOM generates an FOD guess for the molecule by running a conventional KS-DFT calculation with just like any other geometry optimization in ASE, e.g., the PBEsol functional120 and the pc-0 basis set,121,122 localiz- by visualizing the trajectory of the system generated by ing the resulting occupied orbitals with the Foster–Boys (’fb’) ASE. method, and calculating their centroids. The resulting FODs are stored in an .xyz file together with the nuclear informa- tion, and are shown for this example in figure 2.

C. Repeated calculations

Having the optimized FODs for a given level of the- ory (including, e.g., the exchange-correlation functional, quadrature grid, and the orbital basis set), calculations for other levels of theory can easily be performed by starting the calculations from the preoptimized FODs. Because the cost of the FLO-SIC calculation is heavily dependent on the size of the orbital basis set, it is rec- ommended to start out with preoptimized FODs from calculations using small basis sets before going to expen- sive calculations in large but more accurate basis sets.

Note, however, that if the level of theory is changed significantly, e.g., by going from an LDA to a GGA or a meta-GGA, the optimal FOD positions may also ex- perience large changes, decreasing the value of the pre- optimization. The changes in the optimal FODs are re- lated not only to the effect of the total density changing FIG. 2. The nuclear geometry for tetracyanoethylene with the exchange-correlation functional, but also to the (C2(CN)4) and the FODs arising from the PyCOM script shown in figure 1. Color code: C: brown, N: blue, FODs: FLO single electron densities having much sharper fea- 123 green. Note: The FODs for spin up and spin down elec- tures than the total density, which accentuates the trons are identical in this case. Solid lines between nuclei sensitivity to the DFA. FLO-SIC may thus be more sen- indicate bonds and tiny solid lines between the FODs indi- sitive to the exchange-correlation functional than KS- cate the valence electronic geometry, in this case spanned by DFT. This is also reflected in the functional require- tetrahedra. Face-sharing tetrahedra between C and N atoms ments for PZ-SIC, which differ from KS-DFT. For ex- indicate a triple bond, i.e., BOFOD = 3. Edge-sharing tetra- ample, the Perdew–Burke–Ernzerhof functional with ad- hedra between C and C atoms indicate a double bond, i.e., justments for accuracy for solids120 (PBEsol) has been BOFOD = 2. Corner-sharing tetrahedra between C atoms in- found to afford better accuracy for molecular PZ-SIC cal- dicate a single bond, i.e., BO = 1. Therefore, these initial FOD culations than the original functional,124 at variance to FODs agree well with Lewis/LDQ theory. calculations at the KS-DFT level.87,125 8

SIC start: FOD geometry D. Basis set convergence of the atomization energy of SO2 Initial FOD geometry, kopt := 0 and SF6 kSCF := 0

Create: σ σ Having exemplified the use of the novel code, we pro- σ DFT: FKS,EKS, ψj , nKS FLOs φi ceed with a quantitative application. As was already mentioned in the Introduction, reproducing the atom- ization energies of SO2 and SF6 is known to require Calculate: σ σ No: kSCF + 1 No: kopt + 1 FSIC,EPZ, fi large and flexible basis sets at the Kohn–Sham level of theory,30,31,33 making these molecules ideal candidates for a basis set convergence study of FLO-SIC. For sim- Update: Check: SCF convergence plicity, we chose to use fixed high-level ab initio ge- F σ ,E , ψσ, n ∆E < δ or k = kmax KS KS j KS KS SCF SCF ometries from the W4-17 database127 for the molecules, as fixed nuclear geometries suffice for the present pur- Yes poses of establishing convergence to the complete ba-

Check: FOD convergence sis set limit. We also chose the polarization-consistent f σ <  or k = kmax | i |max opt opt pc-n family of basis sets for this study, since these basis sets are designed for achieving optimal conver- gence to the basis set limit in DFT and Hartree–Fock SIC end calculations.121,122 We furthermore chose to study the PBEsol functional, as it has been found to yield good FIG. 3. FLO-SIC workflow for optimizing both the density accuracy in PZ-SIC calculations as was already men- and the FODs. nKS refers to the KS density. Initial FODs can tioned in the Introduction;87,125 however, basis set con- be generated with the built-in PyCOM routine. Here, kSCF vergence patterns are well-known to be similar for both and kopt refer to the number of SCF iterations and FOD opti- mization steps, respectively, whereas δ and  are user-defined Hartree–Fock and all density functionals in the lack of numerical thresholds for the energy and the FOD forces. The post-Hartree–Fock correlation contributions. inner loop optimizes the density as well as the SIC total en- However, as it is well-known that SIC methods require ergy EPZ, while the outer loop optimizes the FODs by using much larger quadrature grids than KS-DFT to reach sim- σ 38,62 the FOD forces fi as input for any of the gradient-based al- ilar levels of convergence, a (200,590) quadrature grid gorithms found in ASE. was used for all calculations, because preliminary tests re- vealed that such a grid was necessary to converge molec- ular FLO-SIC total energies of SF6 and SO2 to roughly from ase.io import read µEh accuracy. Unpruned grids are used in this work from ase_pyflosic_optimizer import because at variance to KS-DFT, grid pruning leads to flosic_optimize significant errors in FLO-SIC calculations: the orbital densities are not spherically symmetric near the nuclei in # Load combined nuclear/FOD information contrast to the total electron density used in KS-DFT. ase_atoms = read(’FB_GUESS_COM.xyz’) For comparison, we also include results for the de- # Start optimization of density and FODs fault basis sets used in , i.e., the DFO and flosic = flosic_optimize(mode=’flosic-scf’, NRLMOL atoms=ase_atoms, DFO+ basis sets for density functional calculations which charge =0, have been used in a number of FLO-SIC studies in the 9–12,55,71–83,85,86 spin =0, literature. The DFO basis set is recom- xc=’pbesol’, mended for general use,23 while the DFO+ basis set is ob- basis =’ccpvqz’, tained from the DFO basis by adding further polarization opt =’fire’, functions aimed to improve the accuracy of polarizabil- maxstep=0.1, ity calculations;23 naturally, the additional functions in fmax=0.001) DFO+ may also improve the convergence of other prop- erties. For consistency with previous literature, the DFO FIG. 4. Python script file for a FLO-SIC optimization of and DFO+ basis sets were extracted from NRLMOL for both the density and the FODs (indicated by mode=’flosic- this work; the converted sets are available on GitHub as scf’), using the ASE interface. After importing the needed part of PyFLOSIC. modules, the nuclear and FOD information are read from a Four sets of computational models will be considered, file, in this case the .xyz file that was created by the script the first one being Kohn–Sham DFT, and the next three in figure 1. Afterwards, the optimization is started using the nuclear and FOD information, the charge and spin of the being variants of FLO-SIC. In the guess-pc-0 scheme, system, the chosen exchange-correlation functional (PBEsol) the FODs are fixed to the PyCOM starting guess val- and basis set (cc-pVQZ126), the optimizer (FIRE) with its ues, computed in the pc-0 basis set. Next, in the opt- maximum step size in Å, and the numerical threshold for the pc-0 scheme, the FODs are frozen to the variationally maximum absolute FOD force in eV/Å as input. optimized pc-0 values. Last, in the opt-basis scheme, the FODs are fully optimized. All FOD optimizations 9 are carried out until a force convergence criterion of ingly) several times larger than the contraction error in σ −3 maxiσ f < 10 Eh/a0 is satisfied. the KS-DFT calculations. Uncontracting the basis set | i | The atomization energies resulting from the previously allows for an improved description of the core orbitals, described procedures are shown in table I. The data show and the unc-pc-4 results in table II are our best esti- remarkable variations of hundreds of kcal/mol when the mates for the FLO-SIC atomization energies of SO2 and size of the basis set is increased. Because larger basis sets SF6 with the PBEsol functional. The W4-17 atomiza- are successively better at describing polarization effects tion energies for SO2 and SF6 are 260.580 kcal/mol and in the molecules—thus decreasing the total energy of the 485.425 kcal/mol. As can be seen from table II, FLO- molecule—the fully variational atomization energies (i.e. SIC reduces the error in the PBEsol functional from 42.5 KS-DFT and the opt-basis model) increase with increas- kcal/mol and 83.0 kcal/mol for SO2 and SF6 at the KS- ing basis set size. An acceptable level of convergence (ba- DFT level of theory to -36.3 kcal/mol and -33.0 kcal/mol sis set truncation error smaller than 1 kcal/mol) is only at the FLO-SIC level of theory, respectively, analogously reached with the quadruple-ζ pc-3 basis set, highlighting to the findings for the PZ-SIC level of theory in ref. 87. the need to support large basis sets in FLO-SIC calcu- lations which is routinely available in PyFLOSIC. In contrast, the DFO and DFO+ basis sets produce results IV. SUMMARY AND OUTLOOK between the double-ζ pc-1 and the triple-ζ pc-2 basis sets, suggesting the DFO and DFO+ basis sets are merely We have presented the implementation of FLO-SIC in of polarized double-zeta quality. The DFO and DFO+ PyFLOSIC as an extension to the PySCF code, as well basis sets have a basis set truncation error of roughly as given instructive examples of code usage. We have 14 kcal/mol and 30 kcal/mol in KS-DFT calculations on also demonstrated the need to be able to use large and SO2 and SF6, respectively, while in FLO-SIC calculations flexible basis sets in FLO-SIC calculations with a basis set the truncation error for SO increases to 17 kcal/mol, the 2 convergence study of the atomization energies of SO2 and one for SF remaining 30 kcal/mol. These results indi- 6 SF6, for which we showed that the DFO basis sets that cate that the DFO and DFO+ basis sets are woefully have been commonly used in FLO-SIC calculations in the inaccurate for quantitative applications in chemistry, un- literature exhibit truncation errors of tens of kcal/mol. derlining the need to support high-angular momentum The most important features of PyFLOSIC are sum- basis sets in KS-DFT as well as FLO-SIC calculations. marized in figure 5. Our implementation offers the pos- An examination into the three flavors of FLO-SIC sibility to use FLO-SIC with any GTO basis set, var- studied in table I shows that optimization of the FODs ious types of quadrature grids, and hundreds of LDA, is clearly necessary, the guess-pc-0 atomization energies GGA, and meta-GGA exchange-correlation functionals being very far from the fully optimized FLO-SIC values, available in the Libxc and XCFun libraries. FLO-SIC guess-pc-0 overestimating the atomization energy of SF6 depends on FODs, which need to be initialized and op- by over 300 kcal/mol. In contrast, the atomization en- timized. Importantly, the FODs can be initialized au- ergies obtained with FODs frozen to the pc-0 optimal tomatically in PyFLOSIC with PyCOM—the internal values are surprisingly close to the fully variational re- routine based on localized orbitals’ centroids—or read sults, showing differences below 1 kcal/mol for all values in from other FOD generators, such as the PyEFF, except for the pc-3 value for SO2 that differs from the PyLEWIS, and fodMC routines presented in ref. 84. fully variational one by 3.2 kcal/mol. These results sug- FOD optimization is carried out in PyFLOSIC through gest using FODs frozen to preoptimized values for a small an interface to ASE. The electron density is optimized in basis set may be an acceptable alternative to fully varia- PyFLOSIC with an SCF procedure employing a unified tional FLO-SIC calculations, if the goal is to estimate the Hamiltonian. importance of self-interaction corrections to e.g. atom- The version of PyFLOSIC described in this work has ization energies. The preoptimized FODs also allow for already been used in some applications.84,123 However, straightforward basis set convergence studies, as omit- PySCF offers features that have not yet been exploited ting the optimization of the FODs implies a significant in FLO-SIC calculations. For instance, PySCF imple- speed-up of FLO-SIC calculations in large basis sets. ments solvation models, such as the domain-decomposed As a further topic, the contraction error in the pc- conductor-like screening model128–130 (COSMO) as n basis sets for DFT calculations and fully optimized well as the domain-decomposed polarizable continuum FLO-SIC calculations were also studied, since the self- model131,132 (PCM), which could be combined with FLO- interaction correction may affect the core orbitals signif- SIC in PyFLOSIC for studying systems in solution. icantly. The atomization energies in the uncontracted For the next iteration of PyFLOSIC, we also aim pc-n (unc-pc-n) basis sets are shown in table II. As to extend the program to include fully variational FLO- shown by the comparison of the results in tables I and SIC, as well as conventional RSIC and CSIC approaches II, the contraction errors are under 1 kcal/mol for all ba- based on orbital rotation techniques which are presently sis sets larger than pc-1, while the contraction error for available only in ERKALE. Since SIC is sensitive to the the largest pc-3 and pc-4 basis sets is in the order of 0.4 details of the evaluation of the density functional (e.g., kcal/mol for FLO-SIC calculations, which is (unsurpris- the numerical quadrature), consistent implementations of 10

TABLE I. Basis set convergence for the atomization energy of SO2 and SF6 in kcal/mol, calculated with an unpruned (200,590) quadrature grid and the PBEsol exchange-correlation functional. W4-17127 nuclear geometries were used. KS-DFT FLO-SIC (guess-pc-0) FLO-SIC (opt-pc-0) FLO-SIC (opt-basis) basis SO2 SF6 SO2 SF6 SO2 SF6 SO2 SF6 pc-0 155.876 396.551 298.089 574.227 56.087 241.725 56.087 241.725 pc-1 268.312 512.775 412.687 729.099 184.844 388.415 185.488 388.740 DFO 289.147 538.089 434.999 763.010 207.368 422.575 207.406 422.623 DFO+ 289.261 538.638 435.792 763.500 207.205 423.141 207.626 423.183 pc-2 294.159 556.651 440.800 786.006 215.682 442.813 214.234 443.019 pc-3 302.640 568.079 448.222 792.663 220.803 452.028 224.012 452.425 pc-4 303.179 568.534 448.727 792.518 224.647 452.562 224.629 452.898

TABLE II. Atomization energies of SO2 and SF6 in kcal/mol with uncontracted basis sets, calculated with an unpruned (200,590) quadrature grid and the PBEsol exchange-correlation functional. W4-17 nuclear geometries were used. KS-DFT FLO-SIC (opt-basis) basis SO2 SF6 SO2 SF6 unc-pc-0 158.414 400.937 59.557 245.947 unc-pc-1 269.065 515.849 186.428 395.124 unc-pc-2 294.361 557.464 215.115 443.889 unc-pc-3 302.614 568.003 224.295 452.023 unc-pc-4 303.126 568.411 224.235 452.456 the various schemes will allow for unbiased comparison DATA AVAILABILITY of various SIC methods. In addition, FLO-SIC stability analysis along the lines of ref. 68 will be investigated. Due The PyFLOSIC code, presented and used within this to the simple interfacing required for SIC calculations and study, is openly available on GitHub (https://github. the modularity of PyFLOSIC, additional interfaces to ), and can be referenced via 8 com/pyflosic/pyflosic other electronic structure codes such as PSI4 could be https://doi.org/10.5281/zenodo.3948143. provided in the future.

An important feature of PySCF are calculations on ACKNOWLEDGMENTS crystalline systems through the use of periodic boundary conditions.7,14 Although calculations on periodic solids J. Kortus and S. Schwalbe have been funded by the with exact exchange are tractable within a GTO basis Deutsche Forschungsgemeinschaft (DFG, German Re- set as in PySCF,133 allowing the use of various hybrid search Foundation) - project-id 421663657 - KO 1924/9- functionals that are less prone to self-interaction error, 1. J. Kortus and J. Kraus have been funded by the DFG exact exchange is undesirable in many cases in the study - project-id 169148856 - SFB 920, subproject A04. K. of the solid state due to problems with, e.g., metallic sys- Trepte has been supported by the U.S. Department of tems. SIC does not pose such problems, and the applica- Energy, Office of Science, Office of Basic Energy Sciences, tion of SIC to solid-state systems has been an active field as part of the Computational Chemical Sciences Pro- of research for several decades.45,64,67,134–157 PyFLOSIC gram under Award Number #de-sc0018331. S. Lehtola could be extended to solids in the future as well. The ne- has been supported by the Academy of Finland (Suomen cessity of a localized orbital picture within a periodic sys- Akatemia) through project number 311149. We also tem is a complication for pushing FLO-SIC to the solid thank the ZIH in Dresden for computational time and state, requiring changes to the algorithms as well as to support, Prof. Mark R. Pederson for creating the FLO- the initialization of FODs. Although orbital localization SIC method that is the core of everything discussed methods are less developed for the solid state than for within the manuscript, and Dr. Torsten Hahn for var- molecular studies, solid-state FODs could be initialized ious discussions and his contributions in an earlier stage in analogy to the present work either with maximally lo- of the project. calized Foster–Boys158 or generalized Pipek–Mezey Wan- nier functions.159 The other methods of ref. 84 might also 1Q. Sun, “Libcint: An efficient general integral library for Gaus- be extended for the solid state; for instance, developmen- sian basis functions,” J. Comput. Chem. 36, 1664 (2015), tal versions of PyEFF and fodMC are already able to arXiv:1412.0649. 2J. Kim, A. D. Baczewski, T. D. Beaudet, A. Benali, M. C. Ben- provide FODs that appear reasonable for simple solids nett, M. A. Berrill, N. S. Blunt, E. J. L. Borda, M. Casula, 160 (e.g., solid deuterium or lithium) and metal-organic D. M. Ceperley, et al., “QMCPACK: an open source ab ini- frameworks (MOFs).161–165 tio quantum Monte Carlo package for the electronic structure 11

all LDA, GGA, meta-GGA functionals from Libxc/XCFun all GTO basis sets supported i.e., all-electron or effective core potential, arbitrary angular momentum transparent usage of PySCF features e.g., density fitting procedures, custom numerical quadrature

automatic generation PyCOM, interfaces to external generators FODs analytical FOD forces automatic optimization via ASE e.g., using CG, FIRE, (L-)BFGS

two unified SIC Hamiltonians Density available for SCF and analysis (e.g., eigenvalues)

FLO-SIC total energy and eigenvalues Output optimal set of FLOs/FODs

of FLOs Post-Processing FOD bond order

FIG. 5. PyFLOSIC code features.

of atoms, molecules and solids,” J. Phys. Condens. Matter 30, density functional theory,” J. Chem. Phys. 140, 121103 (2014). 195901 (2018), arXiv:1802.06922. 10M. R. Pederson, “Fermi orbital derivatives in self-interaction 3S. Lehtola, C. Steigemann, M. J. T. Oliveira, and M. A. L. corrected density functional theory: Applications to closed shell Marques, “Recent developments in LIBXC – A comprehensive atoms,” J. Chem. Phys. 142, 064112 (2015), arXiv:1412.3101. library of functionals for density functional theory,” SoftwareX 11M. R. Pederson and T. Baruah, “Self-interaction corrections 7, 1 (2018). within the Fermi-orbital-based formalism,” Adv. At., Mol., Opt. 4M. F. Herbst, M. Scheurer, T. Fransson, D. R. Rehn, and Phys. 64, 153 (2015). A. Dreuw, “adcc: A versatile toolkit for rapid development of 12Z.-h. Yang, M. R. Pederson, and J. P. Perdew, “Full self- algebraic-diagrammatic construction methods,” Wiley Interdis- consistency in the Fermi-orbital self-interaction correction,” cip. Rev.: Comput. Mol. Sci. , e1462 (2019), arXiv:1910.07757. Phys. Rev. A 95, 052505 (2017). 5P. Koval, M. Barbry, and D. Sánchez-Portal, “PySCF-NAO: 13L. Fiedler, “Implementation and reassessment of the Fermi– An efficient and flexible implementation of linear response time- Löwdin orbital self-interaction correction for LDA, GGA dependent density functional theory with numerical atomic or- and mGGA functionals, Master’s thesis, TU Bergakademie bitals,” Comput. Phys. Commun. 236, 188 (2019). Freiberg,” (2018). 6S. Iskakov, A. A. Rusakov, D. Zgid, and E. Gull, “Effect of prop- 14Q. Sun, T. C. Berkelbach, N. S. Blunt, G. H. Booth, S. Guo, agator renormalization on the band gap of insulating solids,” Z. Li, J. Liu, J. D. McClain, E. R. Sayfutyarova, S. Sharma, Phys. Rev. B 100, 085112 (2019), arXiv:1812.07027. et al., “PySCF: The Python-based simulations of chemistry 7Q. Sun, X. Zhang, S. Banerjee, P. Bao, M. Barbry, N. S. Blunt, framework,” Wiley Interdiscip. Rev. Comput. Mol. Sci. 8, e1340 N. A. Bogdanov, G. H. Booth, J. Chen, Z.-H. Cui, J. J. Eriksen, (2018). Y. Gao, S. Guo, J. Hermann, M. R. Hermes, K. Koh, P. Ko- 15P. Hohenberg and W. Kohn, “Inhomogeneous electron gas,” val, S. Lehtola, Z. Li, J. Liu, N. Mardirossian, J. D. McClain, Phys. Rev. 136, B864 (1964). M. Motta, B. Mussard, H. Q. Pham, A. Pulkin, W. Purwanto, 16W. Kohn and L. J. Sham, “Self-consistent equations includ- P. J. Robinson, E. Ronca, E. R. Sayfutyarova, M. Scheurer, ing exchange and correlation effects,” Phys. Rev. 140, A1133 H. F. Schurkus, J. E. T. Smith, C. Sun, S.-N. Sun, S. Upad- (1965). hyay, L. K. Wagner, X. Wang, A. White, J. D. Whitfield, M. J. 17Q. Sun, “Co-iterative augmented Hessian method for orbital op- Williamson, S. Wouters, J. Yang, J. M. Yu, T. Zhu, T. C. Berkel- timization,” arXiv:1610.08423 (2016). bach, S. Sharma, A. Y. Sokolov, and G. K.-L. Chan, “Recent 18M. R. Pederson and K. A. Jackson, “Variational mesh for developments in the pyscf program package,” J. Chem. Phys. quantum-mechanical simulations,” Phys. Rev. B 41, 7453 153, 024109 (2020), arXiv:2002.12531. (1990). 8D. G. A. Smith, L. A. Burns, A. C. Simmonett, R. M. Parrish, 19M. R. Pederson, K. A. Jackson, and W. E. Pickett, “Local- M. C. Schieber, R. Galvelis, P. Kraus, H. Kruse, R. Di Remi- density-approximation-based simulations of hydrocarbon inter- gio, A. Alenaizan, A. M. James, S. Lehtola, J. P. Misiewicz, actions with applications to diamond chemical vapor deposi- M. Scheurer, R. A. Shaw, J. B. Schriber, Y. Xie, Z. L. Glick, tion,” Phys. Rev. B 44, 3891 (1991). D. A. Sirianni, J. S. O’Brien, J. M. Waldrop, A. Kumar, E. G. 20M. R. Pederson and K. A. Jackson, “Pseudoenergies for simula- Hohenstein, B. P. Pritchard, B. R. Brooks, H. F. Schaefer, A. Y. tions on metallic systems,” Phys. Rev. B 43, 7312 (1991). Sokolov, K. Patkowski, A. E. DePrince, U. Bozkaya, R. A. King, 21J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. F. A. Evangelista, J. M. Turney, T. D. Crawford, and C. D. Pederson, D. J. Singh, and C. Fiolhais, “Atoms, molecules, Sherrill, “PSI4 1.4: Open-source software for high-throughput solids, and surfaces: Applications of the generalized gradient quantum chemistry,” J. Chem. Phys. 152, 184108 (2020). approximation for exchange and correlation,” Phys. Rev. B 46, 9M. R. Pederson, A. Ruzsinszky, and J. P. Perdew, “Commu- 6671 (1992). nication: Self-interaction correction with unitary invariance in 12

22D. Porezag and M. R. Pederson, “Infrared intensities and reaction. Effect of self-interaction correction,” Chem. Phys. Lett. Raman-scattering activities within density-functional theory,” 221, 100 (1994). Phys. Rev. B 54, 7830 (1996). 44E. Kraisler and L. Kronik, “Piecewise linearity of approximate 23D. V. Porezag, “Development of ab-initio and approximate density functionals revisited: Implications for frontier orbital density functional methods and their application to complex energies,” Phys. Rev. Lett. 110, 126403 (2013), arXiv:1211.5950. fullerene systems, PhD thesis, TU Chemnitz,” (1997). 45J. P. Perdew and A. Zunger, “Self-interaction correction to 24D. Porezag and M. R. Pederson, “Optimization of Gaussian basis density-functional approximations for many-electron systems,” sets for density-functional calculations,” Phys. Rev. A 60, 2840 Phys. Rev. B 23, 5048 (1981). (1999). 46M. R. Pederson and C. C. Lin, “Localized and canonical atomic 25J. Kortus and M. R. Pederson, “Magnetic and vibrational prop- orbitals in self-interaction corrected local density functional ap- erties of the uniaxial Fe13O8 cluster,” Phys. Rev. B 62, 5755 proximation,” J. Chem. Phys. 88, 1807 (1988). (2000). 47O. A. Vydrov and G. E. Scuseria, “Ionization potentials and elec- 26M. R. Pederson, D. V. Porezag, J. Kortus, and D. C. Patton, tron affinities in the Perdew—Zunger self-interaction corrected “Strategies for massively parallel local-orbital-based electronic density-functional theory,” J. Chem. Phys. 122, 184107 (2005). structure methods,” Phys. Status Solidi B 217, 197 (2000). 48S. Klüpfel, P. Klüpfel, and H. Jónsson, “Importance of complex 27F. W. Aquino and B. M. Wong, “Additional insights between orbitals in calculating the self-interaction-corrected ground state Fermi–Löwdin orbital SIC and the localization equation con- of atoms,” Phys. Rev. A 84, 050501 (2011), arXiv:1308.6063. straints in SIC-DFT,” J. Phys. Chem. Lett. 9, 6456 (2018). 49H. Gudmundsdóttir, Y. Zhang, P. M. Weber, and H. Jóns- 28F. W. Aquino, R. Shinde, and B. M. Wong, “Fractional occu- son, “Self-interaction corrected density functional calculations of pation numbers and self-interaction correction-scaling methods molecular Rydberg states,” J. Chem. Phys. 139, 194102 (2013). with the Fermi–Löwdin orbital self-interaction correction ap- 50G. Borghi, A. Ferretti, N. L. Nguyen, I. Dabo, and proach,” J. Comput. Chem. 41, 1200 (2020). N. Marzari, “Koopmans-compliant functionals and their per- 29B. P. Pritchard, D. Altarawy, B. T. Didier, T. D. Gibson, and formance against reference molecular data,” Phys. Rev. B 90, T. L. Windus, “A new basis set exchange: An open, up-to-date 075135 (2014), arXiv:1405.4635. resource for the molecular sciences community,” J. Chem. Inf. 51H. Gudmundsdóttir, Y. Zhang, P. M. Weber, and Model. 59, 4814 (2019). H. Jónsson, “Self-interaction corrected density functional cal- 30S. R. Jensen, S. Saha, J. A. Flores-Livas, W. Huhn, V. Blum, culations of Rydberg states of molecular clusters: N,N- S. Goedecker, and L. Frediani, “The elephant in the room of dimethylisopropylamine,” J. Chem. Phys. 141, 234308 (2014). density functional theory calculations,” J. Phys. Chem. Lett. 8, 52J. P. Perdew, A. Ruzsinszky, J. Sun, and M. R. Pederson, 1449 (2017), arXiv:1702.00957. “Paradox of self-interaction correction: How can anything so 31F. Jensen, “How large is the elephant in the density func- right be so wrong?” Adv. At., Mol., Opt. Phys. 64, 1 (2015). tional theory room?” J. Phys. Chem. A 121, 6104 (2017), 53X. Cheng, Y. Zhang, E. Jónsson, H. Jónsson, and P. M. Weber, arXiv:1704.08832. “Charge localization in a diamine cation provides a test of energy 32D. Feller and D. A. Dixon, “Density functional theory and the functionals and self-interaction correction,” Nat. Commun. 7, basis set truncation problem with correlation consistent basis 11013 (2016). sets: Elephant in the room or mouse in the closet?” J. Phys. 54Y. Zhang, P. M. Weber, and H. Jónsson, “Self-interaction cor- Chem. A 122, 2598 (2018). rected functional calculations of a dipole-bound molecular an- 33S. Lehtola, “Polarized Gaussian basis sets from one-electron ion,” J. Phys. Chem. Lett. 7, 2068 (2016). ions,” J. Chem. Phys. 152, 134108 (2020), arXiv:2001.04224. 55S. Schwalbe, T. Hahn, S. Liebing, K. Trepte, and J. Kortus, 34S. Lehtola, “Curing basis set overcompleteness with pivoted “Fermi–Löwdin orbital self-interaction corrected density func- Cholesky decompositions,” J. Chem. Phys. 151, 241102 (2019), tional theory: Ionization potentials and enthalpies of forma- arXiv:1911.10372. tion,” J. Comput. Chem. 39, 2463 (2018). 35S. Lehtola, “Accurate reproduction of strongly repulsive in- 56G. Borghi, C.-H. Park, N. L. Nguyen, A. Ferretti, and teratomic potentials,” Phys. Rev. A 101, 032504 (2020), N. Marzari, “Variational minimization of orbital-density- arXiv:1912.12624. dependent functionals,” Phys. Rev. B 91, 155112 (2015). 36U. Ekström, L. Visscher, R. Bast, A. J. Thorvaldsen, and 57J. Garza, J. A. Nichols, and D. A. Dixon, “The optimized ef- K. Ruud, “Arbitrary-order density functional response theory fective potential and the self-interaction correction in density from automatic differentiation,” J. Chem. Theory Comput. 6, functional theory: Application to molecules,” J. Chem. Phys. 1971 (2010). 112, 7880 (2000). 37“See https://www.tddft.org/programs/libxc/functionals/ 58J. Garza, R. Vargas, J. A. Nichols, and D. A. Dixon, “Or- for list of functionals in Libxc (accessed 17 July 2020) and bital energy analysis with respect to LDA and self-interaction https://xcfun.readthedocs.io/en/latest/functionals.html corrected exchange-only potentials,” J. Chem. Phys. 114, 639 for those in XCFun (accessed 17 July 2020).”. (2001). 38S. Lehtola and H. Jónsson, “Variational, self-consistent imple- 59S. Patchkovskii, J. Autschbach, and T. Ziegler, “Curing diffi- mentation of the Perdew–Zunger self-interaction correction with cult cases in magnetic properties prediction with self-interaction complex optimal orbitals,” J. Chem. Theory Comput. 10, 5324 corrected density functional theory,” J. Chem. Phys. 115, 26 (2014). (2001). 39C. D. Sherrill, “Frontiers in electronic structure theory,” J. 60S. Patchkovskii and T. Ziegler, “Improving “difficult” reaction Chem. Phys. 132, 110902 (2010). barriers with self-interaction corrected density functional the- 40K. Burke, “Perspective on density functional theory,” J. Chem. ory,” J. Chem. Phys. 116, 7806 (2002). Phys. 136, 150901 (2012), arXiv:1201.3679. 61S. Patchkovskii and T. Ziegler, “Phosphorus NMR chemical 41A. J. Cohen, P. Mori-Sánchez, and W. Yang, “Insights into shifts with self-interaction free, gradient-corrected DFT,” J. current limitations of density functional theory,” Science 321, Phys. Chem. A 106, 1088 (2002). 792 (2008). 62O. A. Vydrov and G. E. Scuseria, “Effect of the Perdew-–Zunger 42J. P. Perdew, A. Ruzsinszky, L. A. Constantin, J. Sun, and self-interaction correction on the thermochemical performance G. I. Csonka, “Some fundamental issues in ground-state density of approximate density functionals,” J. Chem. Phys. 121, 8187 functional theory: A guide for the perplexed,” J. Chem. Theory (2004). Comput. 5, 902 (2009). 63O. A. Vydrov, G. E. Scuseria, J. P. Perdew, A. Ruzsinszky, and 43B. G. Johnson, C. A. Gonzales, P. M. W. Gill, and J. A. Pople, G. I. Csonka, “Scaling down the Perdew–Zunger self-interaction “A density functional study of the simplest hydrogen abstraction correction in many-electron regions,” J. Chem. Phys. 124, 13

094108 (2006). the direction of resolving the paradox of Perdew–Zunger self- 64C. D. Pemmaraju, T. Archer, D. Sánchez-Portal, and S. San- interaction correction,” J. Chem. Phys. 151, 214108 (2019), vito, “Atomic-orbital-based approximate self-interaction correc- arXiv:1911.08659. tion scheme for molecules and solids,” Phys. Rev. B 75, 045101 81K. P. K. Withanage, S. Akter, C. Shahi, R. P. Joshi, C. Diaz, (2007). Y. Yamamoto, R. Zope, T. Baruah, J. P. Perdew, J. E. Per- 65S. Klüpfel, P. Klüpfel, and H. Jónsson, “The effect of the alta, and K. A. Jackson, “Self-interaction-free electric dipole Perdew–Zunger self-interaction correction to density function- polarizabilities for atoms and their ions using the Fermi–Löwdin als on the energetics of small molecules,” J. Chem. Phys. 137, self-interaction correction,” Phys. Rev. A 100, 012505 (2019). 124102 (2012). 82B. Santra and J. P. Perdew, “Perdew–Zunger self-interaction 66A. Valdes, J. Brillet, M. Gratzel, H. Gudmundsdóttir, H. A. correction: How wrong for uniform densities and large-Z Hansen, H. Jónsson, P. Klüpfel, G.-J. Kroes, F. Le Formal, I. C. atoms?” J. Chem. Phys. 150, 174106 (2019), arXiv:1902.00117. Man, R. S. Martins, J. K. Norskov, J. Rossmeisl, K. Sivula, 83K. Trepte, S. Schwalbe, T. Hahn, J. Kortus, D.-y. Kao, Y. Ya- A. Vojvodic, and M. Zach, “Solar hydrogen production with mamoto, T. Baruah, R. R. Zope, K. P. K. Withanage, J. E. semiconductor metal oxides: New directions in experiment and Peralta, and K. A. Jackson, “Analytic atomic gradients in the theory,” Phys. Chem. Chem. Phys. 14, 49 (2012). Fermi–Löwdin orbital self-interaction correction,” J. Comput. 67H. Gudmundsdóttir, E. Ö. Jónsson, and H. Jónsson, “Calcula- Chem. 40, 820 (2019). tions of Al dopant in α-quartz using a variational implementa- 84S. Schwalbe, K. Trepte, L. Fiedler, A. I. Johnson, J. Kraus, tion of the Perdew–Zunger self-interaction correction,” New J. T. Hahn, J. E. Peralta, K. A. Jackson, and J. Kortus, “Interpre- Phys. 17, 083006 (2015). tation and automatic generation of Fermi-orbital descriptors,” 68S. Lehtola, M. Head-Gordon, and H. Jónsson, “Complex J. Comput. Chem. 40, 2843 (2019). orbitals, multiple local minima, and symmetry breaking in 85Y. Yamamoto, C. M. Diaz, L. Basurto, K. A. Jackson, Perdew—Zunger self-interaction corrected density functional T. Baruah, and R. R. Zope, “Fermi–Löwdin orbital self- theory calculations,” J. Chem. Theory Comput. 12, 3195 (2016). interaction correction using the strongly constrained and appro- 69J. Lehtola, M. Hakala, A. Sakko, and K. Hämäläinen, priately normed meta-GGA functional,” J. Chem. Phys. 151, “ERKALE - A flexible program package for X-ray properties 154105 (2019). of atoms and molecules,” J. Comput. Chem. 33, 1572 (2012). 86J. Vargas, P. Ufondu, T. Baruah, Y. Yamamoto, K. A. Jackson, 70S. Lehtola. ERKALE – HF/DFT from Hel, “http://github. and R. R. Zope, “Importance of self-interaction-error removal in com/susilehtola/erkale,” (2016). density functional calculations on water cluster anions,” Phys. 71T. Hahn, S. Liebing, J. Kortus, and M. R. Pederson, “Fermi or- Chem. Chem. Phys. 22, 3789 (2020). bital self-interaction corrected electronic structure of molecules 87S. Lehtola, E. Ö. Jónsson, and H. Jónsson, “Effect of beyond local density approximation,” J. Chem. Phys. 143, complex-valued optimal orbitals on atomization energies with 224104 (2015), arXiv:1508.00745. the Perdew—Zunger self-interaction correction to density func- 72M. R. Pederson, T. Baruah, D.-y. Kao, and L. Basurto, “Self- tional theory,” J. Chem. Theory Comput. 12, 4296 (2016). 88 interaction corrections applied to Mg-porphyrin, C60, and pen- U. Lundin and O. Eriksson, “Novel method of self-interaction tacene molecules,” J. Chem. Phys. 144, 164117 (2016). corrections in density functional calculations,” Int. J. Quantum 73T. Hahn, S. Schwalbe, J. Kortus, and M. R. Pederson, “Symme- Chem. 81, 247 (2001). try breaking within Fermi–Löwdin orbital self-interaction cor- 89S. Lehtola, F. Blockhuys, and C. Van Alsenoy, “An overview rected density functional theory,” J. Chem. Theory Comput. of self-consistent field calculations within finite basis sets,” 13, 5823 (2017). Molecules 25, 1218 (2020), arXiv:1912.12029. 74D.-y. Kao, K. Withanage, T. Hahn, J. Batool, J. Kortus, and 90J. G. Harrison, R. A. Heaton, and C. C. Lin, “Self-interaction K. Jackson, “Self-consistent self-interaction corrected density correction to the local density Hartree–Fock atomic calculations functional theory calculations for atoms using Fermi–Löwdin or- of excited and ground states,” J. Phys. B: At. Mol. Phys. 16, bitals: Optimized Fermi-orbital descriptors for Li-Kr,” J. Chem. 2079 (1983). Phys. 147, 164107 (2017). 91M. R. Pederson, R. A. Heaton, and C. C. Lin, “Local-density 75K. Sharkas, L. Li, K. Trepte, K. P. K. Withanage, R. P. Joshi, Hartree–Fock theory of electronic states of molecules with self- R. R. Zope, T. Baruah, J. K. Johnson, K. A. Jackson, and interaction correction,” J. Chem. Phys. 80, 1972 (1984). J. E. Peralta, “Shrinking self-interaction errors with the Fermi– 92S. Lehtola and H. Jónsson, “Unitary optimization of localized Löwdin orbital self-interaction-corrected density functional ap- molecular orbitals,” J. Chem. Theory Comput. 9, 5365 (2013). proximation,” J. Phys. Chem. A 122, 9307 (2018). 93S. F. Boys, “Construction of some molecular orbitals to be 76R. P. Joshi, K. Trepte, K. P. K. Withanage, K. Sharkas, Y. Ya- approximately invariant for changes from one molecule to an- mamoto, L. Basurto, R. R. Zope, T. Baruah, K. A. Jackson, other,” Rev. Mod. Phys. 32, 296 (1960). and J. E. Peralta, “Fermi–Löwdin orbital self-interaction cor- 94C. Edmiston and K. Ruedenberg, “Localized atomic and molec- rection to magnetic exchange couplings,” J. Chem. Phys. 149, ular orbitals,” Rev. Mod. Phys. 35, 457 (1963). 164101 (2018). 95J. Pipek and P. G. Mezey, “A fast intrinsic localization proce- 77K. P. K. Withanage, K. Trepte, J. E. Peralta, T. Baruah, dure applicable for ab initio and semiempirical linear combina- R. Zope, and K. A. Jackson, “On the question of the total tion of atomic orbital wave functions,” J. Chem. Phys. 90, 4916 energy in the Fermi–Löwdin Orbital self-interaction correction (1989). method,” J. Chem. Theory Comput. 14, 4122 (2018). 96S. Lehtola and H. Jónsson, “Pipek–Mezey orbital localization 78A. I. Johnson, K. P. K. Withanage, K. Sharkas, Y. Yamamoto, using various partial charge estimates,” J. Chem. Theory Com- T. Baruah, R. R. Zope, J. E. Peralta, and K. A. Jackson, “The put. 10, 642 (2014). effect of self-interaction error on electrostatic dipoles calculated 97W. L. Luken and J. C. Culberson, “Mobility of the Fermi hole using density functional theory,” J. Chem. Phys. 151, 174106 in a single-determinant wavefunction,” Int. J. Quantum Chem. (2019). 22, 265 (1982). 79K. A. Jackson, J. E. Peralta, R. P. Joshi, K. P. Withanage, 98W. L. Luken and D. N. Beratan, “Localized orbitals and the K. Trepte, K. Sharkas, and A. I. Johnson, “Towards efficient Fermi hole,” Theor. Chim. Acta 61, 265 (1982). density functional theory calculations without self-interaction: 99W. L. Luken, “Properties of the Fermi hole in molecules,” Croat. The Fermi–Löwdin orbital self-interaction correction,” J. Phys.: Chem. Acta 57, 1283 (1984). Conf. Ser. 1290, 012002 (2019). 100W. L. Luken and J. C. Culberson, “Localized orbitals based on 80R. R. Zope, Y. Yamamoto, C. M. Diaz, T. Baruah, J. E. Per- the Fermi hole,” Theor. Chim. Acta 66, 279 (1984). alta, K. A. Jackson, B. Santra, and J. P. Perdew, “A step in 14

101P.-O. Löwdin, “On the non-orthogonality problem connected 125E. Ö. Jónsson, S. Lehtola, and H. Jónsson, “Towards an opti- with the use of atomic wave functions in the theory of molecules mal gradient-dependent energy functional of the PZ-SIC form,” and crystals,” J. Chem. Phys. 18, 365 (1950). Procedia Comput. Sci. 51, 1858 (2015). 102G. N. Lewis, “The atom and the molecule,” J. Am. Chem. Soc. 126T. H. Dunning, “Gaussian basis sets for use in correlated molec- 38, 762 (1916). ular calculations. I. The atoms boron through neon and hydro- 103J. W. Linnett, “Valence-bond structures: A new proposal,” Na- gen,” J. Chem. Phys. 90, 1007 (1989). ture 187, 859 (1960). 127A. Karton, N. Sylvetsky, and J. M. L. Martin, “W4-17: A 104J. W. Linnett, “A modification of the Lewis–Langmuir octet diverse and high-confidence dataset of atomization energies for rule,” J. Am. Chem. Soc. 83, 2643 (1961). benchmarking high-level electronic structure methods,” J. Com- 105J. W. Linnett, Electronic structure of molecules (Methuen & put. Chem. 38, 2063 (2017). Co. Ltd., London, 1964). 128E. Cancès, Y. Maday, and B. Stamm, “Domain decomposi- 106W. F. Luder, “Electronic structure of molecules (Linnett, JW),” tion for implicit solvation models,” J. Chem. Phys. 139, 054111 J. Chem. Educ. 43, 55 (1964). (2013). 107J. Kraus, “FLOSIC-DFT analysis of chemical bonding: applica- 129F. Lipparini, B. Stamm, E. Cancès, Y. Maday, and B. Men- tion to diatomic molecules, Bachelor’s thesis, TU Bergakademie nucci, “Fast domain decomposition algorithm for continuum sol- Freiberg,” (2017). vation models: Energy and first derivatives,” J. Chem. Theory 108A. H. Larsen, J. J. Mortensen, J. Blomqvist, I. E. Castelli, Comput. 9, 3637 (2013). R. Christensen, M. Dułak, J. Friis, M. N. Groves, B. Ham- 130F. Lipparini, G. Scalmani, L. Lagardère, B. Stamm, E. Can- mer, C. Hargus, et al., “The atomic simulation environment - cès, Y. Maday, J.-P. Piquemal, M. J. Frisch, and B. Mennucci, a Python library for working with atoms,” J. Phys. Condens. “Quantum, classical, and hybrid QM/MM calculations in solu- Matter 29, 273002 (2017). tion: General implementation of the ddCOSMO linear scaling 109C. G. Broyden, “The convergence of a class of double-rank min- strategy,” J. Chem. Phys. 141, 184108 (2014). imization algorithms 1. General considerations,” IMA J. Appl. 131B. Stamm, E. Cancès, F. Lipparini, and Y. Maday, “A new Math. 6, 76 (1970). discretization for the polarizable continuum model within the 110R. Fletcher, “A new approach to variable metric algorithms,” domain decomposition paradigm,” J. Chem. Phys. 144, 054101 Comput. J. 13, 317 (1970). (2016). 111D. Goldfarb, “A family of variable-metric methods derived by 132F. Lipparini and B. Mennucci, “Perspective: Polarizable con- variational means,” Math. Comput. 24, 23 (1970). tinuum models for quantum-mechanical descriptions,” J. Chem. 112D. F. Shanno, “Conditioning of quasi-Newton methods for func- Phys. 144, 160901 (2016). tion minimization,” Math. Comput. 24, 647 (1970). 133C. Pisani, R. Dovesi, and C. Roetti, Hartree–Fock ab initio 113J. Nocedal, “Updating quasi-Newton matrices with limited stor- treatment of crystalline systems, Lecture Notes in Chemistry, age,” Math. Comput. 35, 773 (1980). Vol. 48 (Springer Berlin Heidelberg, Berlin, Heidelberg, 1988). 114D. C. Liu and J. Nocedal, “On the limited memory BFGS 134R. A. Heaton, J. G. Harrison, and C. C. Lin, “Self-interaction method for large scale optimization,” Math. Program. 45, 503 correction for density-functional theory of electronic energy (1989). bands of solids,” Phys. Rev. B 28, 5992 (1983). 115R. H. Byrd, P. Lu, J. Nocedal, and C. Zhu, “A limited memory 135R. A. Heaton and C. C. Lin, “Self-interaction-correction theory algorithm for bound constrained optimization,” SIAM J. Sci. for density functional calculations of electronic energy bands for Comput. 16, 1190 (1995). the lithium chloride ,” J. Phys. C: Solid State Phys. 17, 116C. Zhu, R. H. Byrd, P. Lu, and J. Nocedal, “Algorithm 778: L- 1853 (1984). BFGS-B: Fortran subroutines for large-scale bound-constrained 136A. Svane and O. Gunnarsson, “Anti-ferromagnetic moment for- optimization,” ACM Trans. Math. Softw. 23, 550 (1997). mation in the self-interaction-corrected density functional for- 117E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumb- malism,” EPL 7, 171 (1988). sch, “Structural relaxation made simple,” Phys. Rev. Lett. 97, 137A. Svane and O. Gunnarsson, “Localization in the self- 170201 (2006). interaction-corrected density-functional formalism,” Phys. Rev. 118R. A. Heaton, J. G. Harrison, and C. C. Lin, “Self-interaction B 37, 9919 (1988). correction for energy band calculations: Application to LiCl,” 138S. C. Erwin and C. C. Lin, “The self-interaction-corrected elec- Solid State Commun. 41, 827 (1982). tronic band structure of six alkali fluoride and chloride crystals,” 119S. Kim, J. Chen, T. Cheng, A. Gindulyte, J. He, S. He, Q. Li, J. Phys. C: Solid State Phys. 21, 4285 (1988). B. A. Shoemaker, P. A. Thiessen, B. Yu, et al., “PubChem 2019 139Z. Szotek, W. M. Temmerman, and H. Winter, “On the self- update: Improved access to chemical data,” Nucleic Acids Res. interaction correction of localized bands: Application to rare 47, D1102 (2019). gas solids,” Solid State Commun. 74, 1031 (1990). 120J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov, G. E. 140Z. Szotek, W. M. Temmerman, and H. Winter, “On the self- Scuseria, L. A. Constantin, X. Zhou, and K. Burke, “Restoring interaction correction of localized bands: Application to the 4p the density-gradient expansion for exchange in solids and sur- semi-core states in Y,” Phys. B (Amsterdam, Neth.) 165, 275 faces,” Phys. Rev. Lett. 100, 136406 (2008), arXiv:0711.0156. (1990). 121F. Jensen, “Polarization consistent basis sets: Principles,” J. 141A. Svane and O. Gunnarsson, “Hydrogen solid in self- Chem. Phys. 115, 9113 (2001). interaction-corrected local-spin-density approximation,” Solid 122F. Jensen, “Polarization consistent basis sets. II. Estimating the State Commun. 76, 851 (1990). Kohn–Sham basis set limit,” J. Chem. Phys. 116, 7372 (2002). 142A. Svane and O. Gunnarsson, “Transition-metal oxides in the 123C. Shahi, P. Bhattarai, K. Wagle, B. Santra, S. Schwalbe, self-interaction-corrected density-functional formalism,” Phys. T. Hahn, J. Kortus, K. A. Jackson, J. E. Peralta, K. Trepte, Rev. Lett. 65, 1148 (1990). 143 S. Lehtola, N. K. Nepal, H. Myneni, B. Neupane, S. Adhikari, A. Svane, “Electronic structure of La2CuO4 in the self- A. Ruzsinszky, Y. Yamamoto, T. Baruah, R. R. Zope, and interaction-corrected density-functional formalism,” Phys. Rev. J. P. Perdew, “Stretched or noded orbital densities and self- Lett. 68, 1900 (1992). interaction correction in density functional theory,” J. Chem. 144M. M. Rieger and P. Vogl, “Self-interaction corrections in semi- Phys. 150, 174102 (2019). conductors,” Phys. Rev. B 52, 16567 (1995). 124J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized gra- 145D. Vogel, P. Krüger, and J. Pollmann, “Ab initio electronic- dient approximation made simple,” Phys. Rev. Lett. 77, 3865 structure calculations for II-VI semiconductors using self- (1996). interaction-corrected pseudopotentials,” Phys. Rev. B 52, R14316 (1995). 15

146M. Arai and T. Fujiwara, “Electronic structures of transition- 156M. Däne, M. Lueders, A. Ernst, D. Ködderitzsch, W. M. Tem- metal mono-oxides in the self-interaction-corrected local-spin- merman, Z. Szotek, and W. Hergert, “Self-interaction correction density approximation,” Phys. Rev. B 51, 1477 (1995). in multiple scattering theory: Application to transition metal 147D. Vogel, P. Krüger, and J. Pollmann, “Self-interaction oxides,” J. Phys.: Condens. Matter 21, 045604 (2009). and relaxation-corrected pseudopotentials for II-VI semiconduc- 157N. L. Nguyen, N. Colonna, A. Ferretti, and N. Marzari, tors,” Phys. Rev. B 54, 5495 (1996). “Koopmans-compliant spectral functionals for extended sys- 148A. Svane, “Electronic structure of cerium in the self-interaction- tems,” Phys. Rev. X 8, 021051 (2018), arXiv:1708.08518. corrected local-spin-density approximation,” Phys. Rev. B 53, 158N. Marzari and D. Vanderbilt, “Maximally localized generalized 4275 (1996). Wannier functions for composite energy bands,” Phys. Rev. B 149A. Svane, W. M. Temmerman, Z. Szotek, J. Laegsgaard, and 56, 12847 (1997), arXiv:cond-mat/9707145. H. Winter, “Self-interaction-corrected local-spin-density calcu- 159E. Ö. Jónsson, S. Lehtola, M. Puska, and H. Jónsson, lations for rare earth materials,” Int. J. Quantum Chem. 77, “Theory and applications of generalized Pipek–Mezey Wan- 799 (2000). nier functions,” J. Chem. Theory Comput. 13, 460 (2017), 150A. Filippetti and N. A. Spaldin, “Self-interaction-corrected arXiv:1608.06396. pseudopotential scheme for magnetic and strongly-correlated 160J. T. Su and W. A. Goddard, “Excited electron dynamics model- systems,” Phys. Rev. B 67, 125109 (2003), arXiv:cond- ing of warm dense matter,” Phys. Rev. Lett. 99, 185003 (2007). mat/0303042. 161K. Trepte, S. Schwalbe, and G. Seifert, “Electronic and mag- 151E. Bylaska, K. Tsemekhman, and H. Jónsson, “Self-consistent netic properties of DUT-8 (Ni),” Phys. Chem. Chem. Phys. 17, self-interaction corrected DFT: The method and applications 17122 (2015). to extended and confined systems,” in APS Meeting Abstracts 162S. Schwalbe, K. Trepte, G. Seifert, and J. Kortus, “Screening for (2004). high-spin metal organic frameworks (MOFs): density functional 152M. Lüders, A. Ernst, M. Däne, Z. Szotek, A. Svane, D. Köd- theory study on DUT-8(M1,M2) (with Mi = V,...,Cu),” Phys. deritzsch, W. Hergert, B. L. Györffy, and W. M. Temmerman, Chem. Chem. Phys. 18, 8075 (2016). “Self-interaction correction in multiple scattering theory,” Phys. 163K. Trepte, J. Schaber, S. Schwalbe, F. Drache, I. Senkovska, Rev. B 71, 205109 (2005). S. Kaskel, J. Kortus, E. Brunner, and G. Seifert, “The origin 153E. J. Bylaska, K. Tsemekhman, and F. Gao, “New development of the measured chemical shift of 129Xe in UiO-66 and UiO-67 of self-interaction corrected DFT for extended systems applied revealed by DFT investigations,” Phys. Chem. Chem. Phys. 19, to the calculation of native defects in 3C–SiC,” Phys. Scr. T124, 10020 (2017). 86 (2006). 164K. Trepte, S. Schwalbe, J. Schaber, S. Krause, I. Senkovska, 154B. Hourahine, S. Sanna, B. Aradi, C. Köhler, T. Niehaus, S. Kaskel, E. Brunner, J. Kortus, and G. Seifert, “Theoretical and T. Frauenheim, “Self-interaction and strong correlation in and experimental investigations of 129Xe NMR chemical shift DFTB,” J. Phys. Chem. A 111, 5671 (2007). isotherms in metal-organic frameworks,” Phys. Chem. Chem. 155M. Stengel and N. A. Spaldin, “Self-interaction correction with Phys. 20, 25039 (2018). Wannier functions,” Phys. Rev. B 77, 155106 (2008). 165K. Trepte and S. Schwalbe, “Systematic analysis of porosities in metal-organic frameworks,” ChemRxiv (2019), 10.26434/chem- rxiv.10060331.v1.