Standard spin-density functionals for the exchange-correlation energy of a many-electron ground state make serious self-interaction errors which can be corrected by the Perdew-Zunger self-interaction correction (SIC). We propose a size-extensive construction of SIC orbitals which, unlike earlier constructions, makes SIC computationally efficient, and a true spin-density functional. The SIC orbitals are constructed from a unitary transformation that is explicitly dependent on the non-interacting one-particle density matrix. When this SIC is applied to the local spin-density approximation, improvements are found for the atomization energies of molecules.
I. INTRODUCTION AND MOTIVATION
Density Functional Theory (DFT)1,2 allows for a quantum-mechanical description of electrons without direct calculation of a many-electron wavefunction. Approximations to DFT, including the local-spin-density (LSD) approximation2 and gradient approximations,3,4 are widely and successfully used to predict, understand, and optimize many physical and chemical phenomena. However, such approximations have yet to be cast in a form that is both efficient and free of the self-interaction error (SIE).5 For example, in neutral systems the long-range behavior of the Kohn-Sham (KS) potential does not reduce to the −1/r form expected from general considerations. As discussed in Refs. 6,7, this and other SIEs lead to a range of related issues, sometimes referred to as delocalization errors,8 when using DFT approximations to understand chemistry, materials, and physics. Prior to introducing a size-extensive unitarily invariant self-interaction corrected (SIC) density functional theory, we review some physical and chemical systems where this class of functionals could offer improvements.
While standard DFT expressions allow for accurate predictions of electron affinities, based upon total energy differences, the SIE always pushes the occupied-orbital energies upward in a non-systematic way. This problem prevents the calculation of dipole-bound anions, such as ethylene carbonate,9,10 and causes errors when DFT approximations are used for predicting the location and number of defect-levels in solids,11,12 Rydberg excitations in atoms and molecules, or chemical processes mediated by differences in atomic- or fragment-electronegativity. With respect to the latter, the SIE is responsible for dissociation of molecules and crystals into fractionally charged fragments,13 and complicates the understanding of transport in systems containing gold leads and molecular islands.14,15 While constrained DFT methods12,16,17 have been used to address problems related to charge transfer, such constraints are also tacitly assumed when atomization energies are calculated. Although past applications of SIC may not provide empirical evidence that atomization energies can be improved in molecules,18,19 the fact that published DFT atomization energies are generally constrained to non-fractional dissociation blunts this criticism of SIC. Calculated DFT atomic energies as a function of fractional occupation show, for at least localized systems, that the energy as a function of charge state is not linear for fractional electron numbers as required by the Perdew-Levy theorem that requires piece-wise linear energies and derivative discontinuities at integer electron states.6 Time-dependent DFT (TDDFT)20,21 could be improved by inclusion of SIC. However, any new SIC method must address both DFT-based issues and issues related to the explicitly complex canonical orbitals that may result from TDDFT methods. Other manifestations of the self-interaction-induced delocalization error impact predictions of magnetic properties due to qualitatively incorrect predictions of valences in some 3d- and many 4f-electron systems,22–25 and/or the delocalization of 3d orbitals, which causes overestimates in spin-excitations.26 The broken-symmetry calculations that are often used for such magnetic calculations are also needed for describing bond-breaking of condensed phases into separated atoms. This raises questions about the so-called symmetry dilemma, due to the fact that a single determinant constructed from the KS orbitals is not always an eigenstate of symmetry or spin. There are good physical reasons27 to permit such symmetry breaking in density functional approaches including ours here.
In the original formulation of the problem, for a given approximation to the exchange-correlation functional,
In the above equation, the SIC (localized) orbitals {ϕiσ} are used to define orbital densities according to:
with
The one-particle density matrices,
II. FORMULATION OF SELF-INTERACTION CORRECTED DFT WITH UNITARY INVARIANCE
Here, a modification of the original formulation of the Perdew-Zunger self-interaction correction is introduced. For notational simplicity, we consider the large number of molecular and crystalline systems that can, in principle, be described by real KS orbitals and that have an integer number of spin up and spin down electrons. This formulation leads to a size-consistent SIC spin-density functional that is invariant to unitary transformations in the occupied orbital space and leads to a long-range effective potential that scales as −1/r. Rather than allowing the SIC localized orbitals to be any unitary transformation within the occupied orbital space, we introduce a constraint that the orbitals used for constructing Eq. (1) must be explicitly dependent on a quantity that is itself unitarily invariant. The Fermi orbital (FO)35,36 is a specific example of such a quantity. Given a trial set of KS orbitals, the FO (Fiσ) is defined at any point in space, aiσ, according to
In other words, the FO is simply the ratio of the one-particle spin-density matrix to the square root of the spin density and is ultimately a simple transformation of the KS orbitals. It is easy to verify that Wannier functions are a sub-class of Fermi Orbitals, and a few more comments illustrate their physical and chemical nature. At r = aiσ, the value of the absolute square of the FO is identically equal to the total spin density at r = aiσ. Further, the FO associated with any position, aiσ, in space is normalized to unity. For special sets of points, the FOs can be immediately orthogonal (e.g., Wannier Functions) but for general sets of points, they are not orthogonal. Second, the absolute square of the FO is minus the exchange-hole density at r around an electron at aiσ. The exact exchange energy of a single Slater determinant is simply
Because the FO accounts for all the spin density at a given point in space it is expected to be a rather localized function. So, to incorporate self-interaction corrections for the states with spin σ, in a system with N↑ + N↓ electrons, the PZ-SIC formulation can be slightly constrained by invoking the following strategy for each spin:
For a trial set of KS orbitals {ψασ} find Nσ centroids
$\lbrace {\bf a}_{1\sigma }, {\bf a}_{2\sigma },...,{\bf a}_{N_\sigma \sigma }\rbrace$which provide a set of Nσ normalized linearly independent, but not orthogonal FO$\lbrace F_{1\sigma },F_{2\sigma }...F_{N_\sigma \sigma }\rbrace$which, from Eq. (4), will always lie in the space spanned by the KS orbitals.Use Löwdin's method of symmetric orthonormali-zation31,38 to construct a set of localized orthonormal orbitals
$\lbrace \phi _{1\sigma },\phi _{2\sigma },...,\phi _{N_\sigma \sigma }\rbrace$and construct the SIC-DFT energy of Eq. (1) from the set of FOs. This set is a unitary transformation on the KS orbitals.Minimize the SIC-DFT energy as a function of the KS orbitals and the classical centroids of the FOs.
This approach bypasses solution of the localization equations in entirety at the lesser expense of the search for quasi-transferable and chemically appealing semi-classical FOCs (Fig. 1) on which the FO and unitary transformation depend.
Valence Fermi Orbital Centroids (FOC) are superimposed over molecular equilibrium geometries for CO2, C2 (which retains spin-ordering) and N2. In all cases the FOCs form vertices on distorted tetrahedra.
Valence Fermi Orbital Centroids (FOC) are superimposed over molecular equilibrium geometries for CO2, C2 (which retains spin-ordering) and N2. In all cases the FOCs form vertices on distorted tetrahedra.
Compared to the localization equations,31 the FO-SIC (FSIC) formalism adds a constraint that prevents consideration of all unitary transformations. An advantageous example is that the FO-formalism especially discourages the use of the KS orbitals such as Bloch functions, molecular orbitals, or orbital-angular momentum states in atoms for construction of the SIC energy (Eq. (1)). This point is key to regaining size extensivity. A simple molecular case that illustrates this is the He2 molecule. To obtain σg and σu FO, it would be necessary to find a point where the value of the nodeless σg state is zero and the value of the σu state is non-zero, which is impossible.
The PW92 LSD functional was used to calculate all FSIC energies presented here. While future FSIC-DFT approaches will undoubtedly exploit sparsity that is provided by a direct self-consistent solution of the FSIC localized orbitals, in this work we have used the set of self-consistent KS orbitals to construct the spin-density matrices and then minimized the SIC-DFT energy of Eq. (1) by a rather tedious but straightforward direct brute-force variation of the FO-centroids {aiσ}. To account for SIC-induced relaxation of the KS orbitals we used a strategy similar to the optimized effective potential method. Once a set of FO-centroids was obtained from a specific set of KS orbitals, we recalculated the PW92 FSIC-LSD energy using SCF orbitals obtained from different mixtures of all the functionals available in NRLMOL (the generalized-gradient approximation (GGA) of Perdew, Burke, and Ernzerhof (PBE), LSD(PW92), KS-Exchange Only, etc.). In some cases we found that the SCF GGA(PBE) KS orbitals provided slightly lower, albeit chemically insignificant, PW92 FSIC-LSD energies than the SCF LSD(PW92) KS orbitals. This two-step brute force process was iterated until the FSIC-LSD energy was minimized. The eigenvalues of the KS orbitals (Table II) were found by diagonalizing a Hamiltonian similar to that of Ref. 31 given by
with
Eigenvalues (Hartree) and dissociation or atomization energy (De in eV) for the N2 molecule for Hartree-Fock (HF) and SIC exchange-only (SIC-LSDX),30,32 and for FSIC-LSD (this work), LSD(PW92) (this work), and experiment. Columns 2–3 and 4–6 provide results for exchange-only and correlated methods, respectively. The CASSCF41 gives De = 9.85 eV.
. | . | SIC- . | FSIC- . | LSD . | . |
---|---|---|---|---|---|
State . | HF . | LSDX . | LSD . | (PW92) . | Expt. . |
1σg (au) | −15.709 | −15.639 | −15.68 | −13.966 | |
1σu (au) | −15.706 | −15.637 | −15.67 | −13.965 | |
2σg (au) | −1.525 | −1.428 | −1.371 | −1.0422 | |
2σu (au) | −0.775 | −0.745 | −0.788 | −0.492 | |
1πu (au) | −0.62 | −0.639 | −0.687 | −0.438 | |
3σg (au) | −0.631 | −0.600 | −0.658 | −0.388 | −0.573 |
De (eV) | 7.26 | 9.80 | 11.58 | 9.84 |
. | . | SIC- . | FSIC- . | LSD . | . |
---|---|---|---|---|---|
State . | HF . | LSDX . | LSD . | (PW92) . | Expt. . |
1σg (au) | −15.709 | −15.639 | −15.68 | −13.966 | |
1σu (au) | −15.706 | −15.637 | −15.67 | −13.965 | |
2σg (au) | −1.525 | −1.428 | −1.371 | −1.0422 | |
2σu (au) | −0.775 | −0.745 | −0.788 | −0.492 | |
1πu (au) | −0.62 | −0.639 | −0.687 | −0.438 | |
3σg (au) | −0.631 | −0.600 | −0.658 | −0.388 | −0.573 |
De (eV) | 7.26 | 9.80 | 11.58 | 9.84 |
The FSIC formalism reduces the number of unknown parameters needed for describing the unitary transformation from O(N2) to 3N and reduces the number of equations from O(N2) to 3N. Thus, our analysis is that an optimal algorithm for solving the FSIC equations can easily scale as favorably as DFT and that, accounting for sparsity, an O(N) algorithm should be possible in the many-electron limit. In contrast, solution of the localization equations scales as poorly as O(N6) for a small number of electrons. After accounting for sparsity, solution of the localization equations should be slower by O(M3) with M a characteristic number of close neighbors that each atom has. This scaling argument shows that FSIC is potentially much faster than SIC and will actually be once an optimal algorithm is developed.
III. DISCUSSION AND APPLICATIONS TO ATOMS AND MOLECULES
As a first test we have performed FSIC calculations on the six lightest closed shell-atoms (He, Be, Ne, Mg, Ar, Ca). In all cases, the FO procedure provides localized orbitals that resemble the sp3 hybrids and 1s-core orbitals identified in Ref. 32.
In Table I, we compare calculated atomization energies from LSD(PW92),3 GGA(PBE),4 SIC-LSD,33 FSIC-LSD, and experiment. The FSIC-LSD provides significant improvements over LSD as compared to experiment and does as well or better than GGA(PBE) in some cases. The mean absolute error (MAE) for FSIC-LSD in Table I is significantly improved over SIC-LSD. This average improvement is almost entirely due to systematically better performance of FSIC in molecules with π bonds.
Atomization energies (eV) of molecules, with π bonds, ionic bonds, and sigma bonds respectively, as determined from FSIC-LSD (this work). LSD(PW92),3 GGA(PBE),4 and experimental results are from Refs. 18,19,40. SIC-LSD results are from Ref. 33. Energies are relative to spin-polarized atoms. The mean absolute error (MAE) per atom is also provided.
. | LSD . | GGA . | SIC- . | FSIC- . | . |
---|---|---|---|---|---|
Mol. . | (PW92) . | (PBE) . | LSD . | LSD . | Expt. . |
N2 | 11.58 | 10.49 | 10.89 | 9.80 | 9.84 |
O2 | 7.62 | 6.30 | 5.77 | 4.80 | 5.12 |
CO | 12.94 | 11.65 | 12.02 | 11.00 | 11.32 |
CO2 | 20.57 | 18.16 | 18.29 | 16.88 | 17.00 |
C2H2 | 19.93 | 18.01 | 19.81 | 18.93 | 17.52 |
LiF | 6.75 | 6.01 | 6.17 | 5.61 | 6.03 |
H2 | 4.91 | 4.51 | 4.97 | 4.97 | 4.77 |
Li2 | 1.03 | 1.06 | 1.04 | 1.02 | 1.06 |
CH4 | 20.06 | 18.24 | 20.25 | 20.23 | 18.21 |
NH3 | 14.56 | 13.05 | 14.21 | 14.24 | 12.88 |
H2O | 11.64 | 10.27 | 10.68 | 10.71 | 10.10 |
MAE | 1.62 | 0.41 | 0.94 | 0.62 |
. | LSD . | GGA . | SIC- . | FSIC- . | . |
---|---|---|---|---|---|
Mol. . | (PW92) . | (PBE) . | LSD . | LSD . | Expt. . |
N2 | 11.58 | 10.49 | 10.89 | 9.80 | 9.84 |
O2 | 7.62 | 6.30 | 5.77 | 4.80 | 5.12 |
CO | 12.94 | 11.65 | 12.02 | 11.00 | 11.32 |
CO2 | 20.57 | 18.16 | 18.29 | 16.88 | 17.00 |
C2H2 | 19.93 | 18.01 | 19.81 | 18.93 | 17.52 |
LiF | 6.75 | 6.01 | 6.17 | 5.61 | 6.03 |
H2 | 4.91 | 4.51 | 4.97 | 4.97 | 4.77 |
Li2 | 1.03 | 1.06 | 1.04 | 1.02 | 1.06 |
CH4 | 20.06 | 18.24 | 20.25 | 20.23 | 18.21 |
NH3 | 14.56 | 13.05 | 14.21 | 14.24 | 12.88 |
H2O | 11.64 | 10.27 | 10.68 | 10.71 | 10.10 |
MAE | 1.62 | 0.41 | 0.94 | 0.62 |
For example, the strongly covalent singlet N2 molecule,
To summarize, we present and test a simplified and systematic theoretical framework for incorporating self-interaction corrections into a density-functional approximation. Our approach achieves computational efficiency and guarantees size-extensivity, as previous approaches may not. Compared to the accuracy of GGA(PBE), the results for FSIC-LSD are encouraging and greater accuracy is expected when FSIC is applied to meta-generalized gradient approximations.42 The localized orbitals obtained from the FO formulation are topologically similar to those obtained from the localization equations used earlier.32 However, the FSIC does not allow consideration of orbitals that correspond to saddle points and energy maxima whereas the original version of SIC did. Relative to the original formulation, the FO-induced constraint on the unitary transformations leading to localized orbitals allows FSIC to retain the symmetries and size-extensivity that are present in density-functional theory.
ACKNOWLEDGMENTS
M.R.P. thanks Dr. R. A. Heaton who first suggested the possibility of using a FO in the SIC-LSD functional. J.P.P. acknowledges support from NSF Grant No. DMR-1305135.