Next Article in Journal
Note on F-Graph Construction
Next Article in Special Issue
Is There a Quadruple Fe-C Bond in FeC(CO)3?
Previous Article in Journal
District-Heating-Grid Simulation in Python: DiGriPy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Density Functional Theory of Highly Excited States of Coulomb Systems

Department of Theoretical Physics, University of Debrecen, H-4002 Debrecen, Hungary
Submission received: 20 May 2021 / Revised: 13 June 2021 / Accepted: 18 June 2021 / Published: 21 June 2021
(This article belongs to the Special Issue Electronic Correlation)

Abstract

:
The density functional theory proposed earlier for excited states of Coulomb systems is discussed. The localized Hartree–Fock (LHF) and the Krieger, Li, and Iafrate (KLI) methods combined with correlation are generalized for excited states. Illustrative examples include some highly excited states of Li and Na atoms.

1. Introduction

The density functional theory (DFT) was initially a ground-state scheme [1,2], although in special cases the ground-state constrained-search functional [3] yields exact excited-state energies. Rigorously, DFT was extended to excited states as a subspace theory by Theophilou [4] and later to a more general ensemble theory by Gross, Oliveira and Kohn [5,6,7]. These approaches have found several extensions and applications (e.g., [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26]). Nevertheless, they have the unfavorable feature that all levels lying below the excited state intended to study should be taken into account in the calculation. It can be rather inconvenient in the case of higher excited states. Hence, approaches for an individual excited state were worked out.
It was proved that in Coulomb systems the density determines the external potential even in excited states [27,28,29]. Then, a variational bifunctional theory was put forward [30,31] and analyzed [32,33]. Several other significant schemes were developed [34,35,36]; however, time-dependent DFT [37,38,39] is still usually employed for excited states.
In a series of papers [40,41,42], a comprehensive theory for excited states of Coulomb systems is put forward. It is based on the fact that the Coulomb density determines not only its Hamiltonian but the degree of excitation as well. It makes it possible to develop a universal functional valid for any excited state. Moreover, the excited-state Kohn–Sham (KS) equations are similar to the ground-state KS equations.
Unfortunately, the exact form of the exchange-correlation functional is not known. It is not unexpected, as the exchange-correlation functional is unknown even for the ground state. Therefore, this functional should be approximated in calculations. We mention in passing that several exact constraints that the excited-state functionals should satisfy have recently been derived [43,44].
If we consider the exchange as a functional of the orbitals instead of the density, the energy functional is known: it is the well-known Hartree–Fock expression with the KS orbitals, of course. The exchange potential—which is a local potential in DFT—can be obtained by the optimized potential method (OPM). In the ground-state theory, several methods have been proposed to find the local potential whose eigenfunctions would minimize a given energy functional [45,46]. The localized Hartree–Fock (LHF) [47,48,49] and the KLI (Krieger, Li, and Iafrate) [50,51,52] methods proved to be excellent approximations to the OPM. An alternative derivation to the KLI method was also proposed by the present author [53].
Here, extensions of the LHF and KLI methods combined with correlation are proposed to DFT of Coulombic excited states. These approaches provide an almost exact treatment of exchange and can be coupled with any approximate correlation functional. It is worth starting with a very simple approximate correlation functional, so the local Wigner expression is taken here [54]. As an illustration total and excitation energies are presented for some (including highly) excited states of Li and Na atoms.
The paper is organized as follows. In Section 2, the DFT for Coulombic excited states [40,41,42] is summarized. Section 3 presents generalization of the LHF and KLI methods combined with correlation. Section 4 is dedicated to the discussion.

2. DFT for Coulombic Excited States

First, the theory of Coulombic excited states is summarized. Consider a system in an external Coulomb potential of the form
v C o u l ( r ) = β = 1 M Z β r β ,
where r β = | r R β | and M is the number of nuclei. R β and Z β stand for the position and the charge of the nucleus β . The Hamiltonian is
H ^ = T ^ + V ^ e e + i = 1 N v C o u l ( r i ) ,
where T ^ and V ^ e e are the kinetic energy and the electron-electron energy operators. Kato’s theorem [55,56,57,58,59,60,61]
n ¯ β ( r β ) r β r β = 0 = 2 Z β n ( r = R β )
is valid for an excited state, too. Therefore, the cusps of the density n uncover the atomic numbers and the positions of the nuclei. On the other hand, the integral of the density generates the number of electrons. That is, the density provides all parameters of the Coulomb potential (1), consequently determining the external potential, the Hamiltonian (2), and all properties of the Coulomb system. Moreover, a Coulombic electron density cannot be a stationary state density for any other Coulomb external potential, and two different excited states cannot have the same electron density, as proved in [40]. Therefore, the functional
F C o u l [ n ] = E [ n ] n ( r ) v C o u l [ n ; r ] d r
can be defined for Coulombic densities. Unfortunately, no easy method is available to decide whether a given density is Coulombic or not; consequently, it is worth defining the functional F for all electron densities.
Consider first a bifunctional
F [ n , n C o u l ] = min Ψ n { Ψ | Ψ l C o u l [ n C o u l ] = 0 } l = 1 k 1 Ψ | T ^ + V ^ e e | Ψ .
In Equation (5), the minimization is performed with the constraint that each wave function gives the excited-state density n and is orthogonal to the first k 1 eigenfunctions of the Coulomb system of n C o u l . The existence of a Coulomb density close to n is assumed. If more than one Coulomb density can be found at the same distance from n, the one with the smallest F is taken.
F C o u l [ n ] = F ϵ m i n C o u l [ n ] ,
where
F ϵ C o u l [ n ] = min n C o u l F [ n , n C o u l ] ; | | n C o u l n | | ϵ .
The best measure for the distance is not known yet. We prefer the Sobolev-type norm:
d ( n C o u l , n ) n C o u l ( r ) n ( r ) 2 d r + n C o u l ( r ) n ( r ) 2 d r .
The mathematical properties of the functional might depend on the definition. Therefore, this problem should be the subject of future investigation (see further details in [40]).
The minimization leads to the Euler equation
v C o u l ( [ n ] , r ) = δ F C o u l [ n ] δ n ( r )
up to a constant.
In calculations, the Kohn–Sham (KS) system is preferred; therefore, it is valuable to define the non-interacting kinetic energy
T s [ n C o u l ] = min Φ n C o u l { Φ | Φ j [ n C o u l ] = 0 } j = 1 l 1 | | n 1 C o u l n 1 0 | | δ Φ | T ^ | Φ .
The minimization is over the wave functions Φ having the excited-state density n C o u l and orthogonal to the first l 1 eigenfunctions of the non-interacting system. That is, n C o u l is the same in the real and the KS systems. However, the ground states can be different. It may happen that there are more than one KS system with the density n C o u l . Then, take the one in which the KS ground-state density n 1 0 is closest to the true ground-state density n 1 C o u l .
We need the KS kinetic energy as a functional of a not necessarily Coulomb density n. First, a bifunctional is defined:
T s C o u l [ n , n C o u l ] = min Φ n { Φ | Φ l [ n C o u l ] = 0 } l = 1 k 1 | | n 1 C o u l n 1 0 | | δ Φ | T ^ | Φ .
The minimum is taken with the constraint that the density of the lth KS state Φ equals n, Φ is orthogonal to all states Φ j , 1 j < l and the ground-state KS density n 1 0 is as close as possible to the real ground-state density n 1 C o u l . We assume the existence of a unique Coulomb density close to the non-Coulomb density n and construct the functional
T s , ϵ C o u l [ n ] = min n C o u l T s C o u l [ n , n C o u l ] ,
where
| | n C o u l n | | ϵ .
It is expected that there is at least one Coulomb density closer to n than ϵ , when taking a large enough value for ϵ . Finally, the smallest value of ϵ provides the kinetic energy functional
T s C o u l [ n ] = T s , ϵ m i n C o u l [ n ] .
The variational principle provides an Euler equation, within an additive constant,
w C o u l ( [ n ] , r ) = δ T s C o u l [ n ] δ n ( r ) .
To derive the KS equations it is helpful to partition F C o u l [ n ] as
F C o u l [ n ] = T s C o u l [ n ] + J C o u l [ n ] + E x c C o u l [ n ] ,
where J C o u l [ n ] and E x c C o u l [ n ] are the classical Coulomb and exchange-correlation energies. Comparing Equations (9), (15), and (16), the KS potential
w C o u l ( [ n ] , r ) = v C o u l ( [ n ] , r ) + v J C o u l ( [ n ] , r ) + v x c C o u l ( [ n ] , r )
is obtained as the sum of the external, the classical Coulomb and the exchange-correlation potentials. The KS equations have the form
1 2 2 + w C o u l ( [ n ] , r ) ϕ i = ε i ϕ i ,
where the KS orbitals ϕ i provide the density as
n = i = 1 K λ i | ϕ i | 2 ,
where the occupation numbers λ i are 0, 1, or 2 for a non-degenerate system. K stands for the orbital having the highest orbital energy with non-zero occupation number.

3. Orbital-Dependent Exchange-Correlation Functional

The exact form of the exchange-correlation functional is unknown and has to be approximated in calculations even in the original ground-state DFT. As the theory for Coulombic excited states studied in this paper is also valid for the ground state, it is worth testing how the ground-state functionals work. The exchange is known exactly as a functional of orbitals. This functional is the Hartree–Fock (HF) expression. The Kohn–Sham potential is local, therefore a local exchange potential should be generated. In the exchange-only approximation, the so-called optimized potential method (OPM) [45,46] or the localized Hartree–Fock method (LHF) [47,48,49] can be applied. In the latter, it is supposed that the HF and the exchange-only KS determinants are equal. The method have advantageous properties: invariant with respect to unitary transformations of orbitals, the local KS exchange potential is free of self-interaction and, consequently, has correct long-range behavior. Often, instead of OPM or LHF, the KLI (Krieger, Li, and Iafrate) [50,51,52] approximation is applied. KLI can also be obtained by neglecting certain terms from the LHF exchange potential. In KLI, the exchange potential exhibits the correct long-range behavior, but it is not invariant with respect to unitary transformations of orbitals. KLI is much simpler than OPM and more stable if finite-basis-set is applied. Before LHF approach appeared, the present author also provided an alternative derivation of the KLI method. Now, this method is utilized to extend the LHF and the KLI methods to include correlation in the frame of our excited-state theory.
First, the generalized LHF is derived. Denote it LHFC, where the last letter stands for correlation. The total energy can be written as a functional of orbitals:
E C o u l [ ψ ˜ 1 , , ψ ˜ N ] = T s C o u l [ ψ ˜ 1 , , ψ ˜ N ] + n ( r ) v C o u l ( r ) d r + J C o u l [ ψ ˜ 1 , , ψ ˜ N ] + E x C o u l [ ψ ˜ 1 , , ψ ˜ N ] + E c C o u l [ ψ ˜ 1 , , ψ ˜ N ] ,
where
T s C o u l = 1 2 i = 1 K λ i ψ ˜ i * ( x ) 2 ψ ˜ i x ) d x ,
E x C o u l = 1 2 i = 1 K j = 1 K λ i λ j ψ ˜ i * ( x 1 ) ψ ˜ j * ( x 2 ) ψ ˜ i ( x 2 ) ψ ˜ j ( x 1 ) 1 r 12 d x 1 d x 2 .
r 12 = | r 1 r 2 | and x : r , s stand for the spatial and spin coordinates, respectively. The variation of E C o u l with respect to the orbitals leads to the equations
1 2 2 + u ^ C o u l ψ ˜ i = j = 1 K ϵ ˜ i j ψ ˜ j ,
where
u ^ C o u l = v C o u l + v J C o u l + v ^ x C o u l + v ^ c C o u l .
v J C o u l is the Coulomb external potential and
v J C o u l ( r 1 ) = n ( r 2 ) r 12 d r 2
is the classical Coulomb potential. v ^ x C o u l is a Hartree–Fock-like exchange operator
v ^ x C o u l ψ ˜ i ( x 1 ) = j = 1 K λ j ψ ˜ j * ( x 2 ) ψ ˜ j ( x 1 ) 1 r 12 ψ ˜ i ( x 2 ) d x 2 .
The form of the correlation potential v ^ c C o u l can be obtained from the functional derivative of E c C o u l . The correlation functional is unknown; E c C o u l should be approximated. In this paper, a simple local approximation is applied, although the procedure described here can be used for any kind of approximation of the correlation. The right-hand side of Equation (23) arise from the orthonormalization conditions
ψ ˜ i * ( x ) ψ ˜ j ( x ) d x = δ i j ,
where ϵ ˜ i j are the Lagrange multipliers for the constraint (27).
Consider now the corresponding KS equations arising from the variation of E C o u l . In doing this, one must actually constrain the orbitals to orthonormal. Then, we arrive at
1 2 2 + w C o u l ( [ n ] , r ) ϕ ˜ i = j = 1 K ε ˜ i j ϕ ˜ j .
Now, we can compare Equation (28) with the correlated Hartree–Fock-like Equation (23). Multiplying Equation (28) by λ i ϕ ˜ i * and Equation (23) by λ i ψ ˜ i * , then summing for all occupied i, taking the difference of these equations, and finally using the approximation ψ ˜ i ϕ ˜ i , we arrive at the exchange-correlation potential of the LHFC approach
v x c L H F C C o u l = v L H F C S + v L H F C S c + v L H F C e ,
where
v L H F C S ( r ) = 1 n ( r ) i = 1 K λ i ϕ ˜ i * ( x ) v ^ x C o u l ϕ ˜ i ( x )
is the Slater potential,
v L H F C S c = 1 n i = 1 K λ i ϕ ˜ i * ( x ) v ^ c C o u l ϕ ˜ i ( x )
is a Slater-like potential originating from the correlation potential, and
v L H F C e = 1 n i = 1 K λ i j = 1 K ( ε ˜ i j ϵ ˜ i j ) ϕ ˜ i * ϕ ˜ j .
The last term v L H F C e appears because the KS and correlated Hartree–Fock-like equations have different Lagrange multipliers. Using Equations (28) and (23), the difference ε ˜ i j ϵ ˜ i j can be expressed as
ε ˜ i j ϵ ˜ i j = ϕ ˜ i | v x c C o u l v ^ x C o u l v ^ c C o u l | ϕ ˜ j .
Equation (29) provides the exchange-correlation potential. Observe that neglecting the correlation gives the LHF exchange potential of Della-Sala and Görling [47]. If all terms i j are neglected in Equation (32), we arrive at the KLI method with correlation. Omitting correlation, the original KLI approach is given.
The KLI and KLIC potentials can be derived explicitly from the canonical forms of Equations (23) and (28). Equations (23) and (28) can be reformalized by unitary transformations of orbitals:
1 2 2 + u ^ C o u l ψ i = ϵ i ψ i ,
and
1 2 2 + w C o u l ϕ i = ε i ϕ i ,
respectively. u ^ C o u l and w C o u l have the same form as before, but are expressed with the canonical orbitals ψ i and ϕ i instead of ψ ˜ i and ϕ ˜ i . Following the same steps we did in the derivation of Equation (29), we arrive at the KLI-like exchange-correlation potential
v x c K L I C C o u l = v K L I C S + v K L I C S c + v K L I C e ,
where
v K L I C S ( r ) = 1 n ( r ) i = 1 K λ i ϕ i * ( x ) v ^ x C o u l ϕ i ( x )
is the Slater potential,
v K L I C S c = 1 n i = 1 K λ i ϕ i * ( x ) v ^ c C o u l ϕ i ( x )
is a Slater-like potential originating from the correlation potential, and
v K L I C e = 1 n i = 1 K λ i ( ε i ϵ i ) | ϕ i | 2 .
The last term v K L I C e appears because the KS and correlated Hartree–Fock-like equations have different one-electron energies. Using Equations (35) and (34), the difference ε i ϵ i can be expressed as
ε i ϵ i = ϕ i | v x c C o u l v ^ x C o u l v ^ c C o u l | ϕ i .
The potential (36) reduces to the KLI exchange potential if the correlation term is omitted. Observe that, even though the same procedure is used to derive both the LHFC and the KLIC methods, different exchange-correlation potentials are obtained. The reason is that the Lagrange multipliers ε ˜ i j and ϵ ˜ i j are different. Therefore, the unitary transformations are also different. Consequently, even the same procedure leads to different approximation.
Now, the methods derived above are illustrated for some excited states of Li and Na atoms. It is natural to start with a very simple approximation for correlation: KLI is combined with the local Wigner approximation [54].
The correlation energy is given by
E c L W [ n ] = a n b + r s d r ,
where r s is the Wigner–Seitz radius:
r s = ( 3 / 4 π n ) 1 / 3 .
The parameters obtained by Süle and Nagy [62] are used: a = 0.02728 and b = 0.21882 .
The correlation potential obtained by functional derivation of (41)
v c L W = a ( b + 4 3 r s ) ( b + r s ) 2
should be substituted into Equations (38) and (40). Observe that the correlation is taken into account self-consistently.
Table 1 presents the total energies for the ground state and the excited states with configuration 1 s 2 m s , m = 2 , , 7 for the Li atom. The KLI values can be directly compared with the exchange-only results of the spin-dependent localized Hartree–Fock (SLHF) [63], xCOEP [64], and Hartree–Fock (HF) [65] methods. x-COEP refers to exchange-only constrained optimized effective potential (xCOEP) methodology [64]. SLHFc is localized Hartree–Fock combined with Lee–Yang–Parr (LYP) correlation [66]. WFLYP [67] refers to work-function-based exchange [68,69] and LYP correlation potentials. Exact energies obtained with accurate configuration interaction wave function in Hylleraas basis set [70] are presented in the last column. The KLI, SLHF, and xCOEP values are very close to the HF results, the SLHF values are closest. We can see that KLI approximates SLHF and HF excellently. The KLI with local Wigner correlation (KLILW) leads to lower total energies than the exact ones. SLHFc and WFLYP give more accurate total energies.
Calculated and experimental [71] excitation energies of the Li atom are shown in Table 2. The KLI, SLHF, and HF values are very close to each other, while the KLI is a bit worse. While the HF total energies are closer to the exact ones than the xCOEP values, for the excitation energies, xCOEP results are much closer to the exact ones than for the HF data. The KLI values are slightly worse than the HF results, while KLILW and SLHFc overestimate the excitation energies; KLILW values are somewhat better. The first excitation energy computed with time-dependent TDF with ALDA exchange-correlation [72] is less accurate.
Table 3 presents total energies for the ground and the excited states of the Na atom with configuration [ N e ] m s , m = 3 , , 7 . The KLI and ELP values are very close to the HF results, the ELP ones being a bit closer. ELP (effective local potential) [73] denotes an alternative way of solving the exact-exchange OPM. NHF stands for highly accurate numerical Hartree–Fock method [74]. The KLILW leads to lower total energies than the exact ones.
Table 4 displays calculated and experimental [71] excitation energies of the Na atom. The KLI, ELP, and HF values are very close to each other. The KLI with local Wigner correlation approximates experimental results quite well.
It can be concluded that KLI method provides results very close to the HF ones, while the KLILW leads to too low total energies. KLILW overestimates the excitation energies for the Li atom, but it gives much better excitation energies for the Na atom than the KLI or HF method. Of course, these illustrative examples cannot give us full knowledge of the efficiency of the KLILW method, and further studies are necessary. Investigation of other methods and correlation functionals (e.g., [75,76]) will be the subject of future research.

4. Discussion

The great advantage of our method is that a single functional is relevant for any bound (ground or excited) state of a Coulomb system. However, of course, we do not know this functional and its properties. It might happen that it is a jagged, discontinuous functional. It can appear if very similar Coulomb densities have very different values of F C o u l with the consequence that F C o u l would be discontinuous. In our approach, the excitation level of two densities may be different. If two Coulomb densities are close together, they still can have vastly different excitation levels. Therefore, F C o u l might be discontinuous. This problem can be avoided by defining functionals F k C o u l , that is, using different functionals for different excitation levels k. Discontinuities in F k C o u l are much less likely because of the additional dependence on the level of excitation (see further details in [40]). In this paper, it is supposed that the single functional F C o u l is well-behaved and approximations for it are proposed.
We mention in passing that, even though calculations are generally performed in the Kohn–Sham scheme, the Euler equation (9) can also be applied provided that appropriate approximation for the kinetic energy functional is available. Unfortunately, such an approximation is not accessible. However, the Euler equation is very useful. For example, the existence of excited-state Euler equations for specific Shannon information and Fisher information has recently been proved [77]. Even the Euler equation for the relatively specific Shannon information has been derived. It is interesting to note that the Ghosh–Berkowitz–Parr thermodynamic transcription [78] has been recently extended to excited states of Coulomb systems [79]. For Coulomb systems, there is a simple relation between the total energy and phase-space Fisher information both in the ground and excited states. Furthermore, relations for the phase-space fidelity, relative entropy, fidelity susceptibility, and Fisher information have been presented. These kinds of analysis of excited states seem to be important, as excited-state reactivity is a new frontier [80,81,82].
In summary, Coulombic excited states are studied within the density functional theory proposed earlier. Generalizations of the LHF and KLI methods combined with correlation are derived within DFT for Coulombic excited states. In these approaches, exchange is treated (almost) exactly and any approximate correlation functional can be incorporated. As an illustration, total and excitation energies are presented for some (including highly) excited states of Li and Na atoms.

Funding

This research was supported by the National Research, Development and Innovation Fund of Hungary, financed under 123988 funding scheme.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharring is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. J. Arch. 1964, 136, B864. [Google Scholar] [CrossRef] [Green Version]
  2. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. J. Arch. 1965, 140, A1133. [Google Scholar] [CrossRef] [Green Version]
  3. Perdew, J.P.; Levy, M. Extrema of the density functional for the energy: Excited states from the ground-state theory. Phys. Rev. B 1985, 31, 6264–6272. [Google Scholar] [CrossRef] [PubMed]
  4. Theophilou, A.K. The energy density functional formalism for excited states. J. Phys. C Solid State Phys. 1979, 12, 5419–5430. [Google Scholar] [CrossRef]
  5. Gross, E.K.U.; Oliveira, L.N.; Kohn, W. Rayleigh-Ritz variational principle for ensembles of fractionally occupied states. Phys. Rev. A 1988, 37, 2805–2808. [Google Scholar] [CrossRef]
  6. Gross, E.K.U.; Oliveira, L.N.; Kohn, W. Density-functional theory for ensembles of fractionally occupied states. I. Basic formalism. Phys. Rev. A 1988, 37, 2809–2820. [Google Scholar] [CrossRef]
  7. Oliveira, L.N.; Gross, E.K.U.; Kohn, W. Density-functional theory for ensembles of fractionally occupied states. II. Application to the He atom. Phys. Rev. A 1988, 37, 2821–2833. [Google Scholar] [CrossRef] [PubMed]
  8. Nagy, Á. Parameter-free exchange potential for excitation in the density-functional theory: Application to excitation energies within the fractional-occupation approach. Phys. Rev. A 1990, 42, 4388–4390. [Google Scholar] [CrossRef]
  9. Nagy, Á. Excitation energies calculated with parameter-free exchange potential in the density functional theory. J. Phys. B 1991, 24, 4691–4694. [Google Scholar] [CrossRef]
  10. Nagy, Á. Relativistic density-functional theory for ensembles of excited states. Phys. Rev. A 1994, 49, 3074–3076. [Google Scholar] [CrossRef] [PubMed]
  11. Nagy, Á. Density functional theory for excited states. Adv. Quant. Chem. 1998, 29, 159–178. [Google Scholar] [CrossRef]
  12. Nagy, Á.; Andrejkovics, I. Excitation energies in the local density functional theory. J. Phys. B 1994, 27, 233–240. [Google Scholar] [CrossRef]
  13. Andrejkovics, I.; Nagy, Á. Excitation energies in density functional theory: Comparison of several methods for the H2O, N2, CO and C2H4 molecules. Chem. Phys. Lett. 1998, 296, 489–493. [Google Scholar] [CrossRef]
  14. Nagy, Á. Coordinate scaling and adiabatic connection formula for ensembles of fractionally occupied excited states. Int. J. Quantum Chem. 1995, 56, 225–228. [Google Scholar] [CrossRef]
  15. Nagy, Á. Local ensemble exchange potential. J. Phys. B 1996, 29, 389–394. [Google Scholar] [CrossRef]
  16. Nagy, Á.; Howard, I.A.; March, N.H.; Jánosfalvi, Z.S. Subspace density of the first excited state for two harmonically interacting electrons with isotropic harmonic. confinement. Phys. Lett. A 2005, 335, 347–350. [Google Scholar] [CrossRef]
  17. Nagy, Á. Hardness and excitation energy. J. Chem. Sci. 2005, 117, 437–440. [Google Scholar] [CrossRef]
  18. Nagy, Á. Optimized potential method for ensembles of excited states. Int. J. Quantum Chem. 1998, 69, 247–254. [Google Scholar] [CrossRef]
  19. Gidopoulos, N.I.; Papaconstantinou, P.G.; Gross, E.K. Spurious Interactions, and Their Correction, in the Ensemble-Kohn-Sham Scheme for Excited States. Phys. Rev. Lett. 2001, 88, 033003. [Google Scholar] [CrossRef] [Green Version]
  20. Nagy, Á. An alternative optimized potential method for ensembles of excited states. J. Phys. B 2001, 34, 2363–2370. [Google Scholar] [CrossRef]
  21. Tasnádi, F.; Nagy, Á. Study of Subspace Density-Functional Theory. Application of LSDA to Excited States of Atoms. Int. J. Quantum Chem. 2003, 92, 234–238. [Google Scholar] [CrossRef]
  22. Tasnádi, F.; Nagy, Á. Ghost and self-interaction free ensemble calculations for atoms with local exchange-correlation potential. J. Phys. B 2003, 36, 4073–4075. [Google Scholar] [CrossRef]
  23. Tasnádi, F.; Nagy, Á. An approximation to the ensemble Kohn-Sham exchange potential for excited states of atoms. J. Chem. Phys. 2003, 119, 4141–4147. [Google Scholar] [CrossRef]
  24. Yang, Z.H.; Trail, J.R.; Pribram-Jones, A.; Burke, K.; Needs, R.J.; Ullrich, C.A. Exact and approximate Kohn-Sham potentials in ensemble density-functional theory. Phys. Rev. A 2014, 90, 042501. [Google Scholar] [CrossRef] [Green Version]
  25. Pribram-Jones, A.; Yang, Z.H.; Trail, J.R.; Burke, K.; Needs, R.J.; Ullrich, C.A. Excitations and benchmark ensemble density functional theory for two electrons. J. Chem. Phys. 2014, 140, 18A541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Pastorczak, E.; Pernal, K. Ensemble density variational methods with self- and ghost-interaction-corrected functionals. J. Chem. Phys. 2014, 140, 18A514. [Google Scholar] [CrossRef] [Green Version]
  27. Nagy, Á. Excited states in density functional theory. Int. J. Quantum Chem. 1998, 70, 681–691. [Google Scholar] [CrossRef]
  28. Nagy, Á. Electron Correlations and Materials Properties; Gonis, A., Kioussis, N., Ciftan, M., Eds.; Kluwer: New York, NY, USA, 1999; pp. 451–462. [Google Scholar]
  29. Nagy, Á. Theories for excited states. Adv. Quant. Chem. 2003, 42, 363–381. [Google Scholar] [CrossRef]
  30. Levy, M.; Nagy, Á. Variational Density-Functional Theory for an Individual Excited State. Phys. Rev. Lett. 1999, 83, 4361–4364. [Google Scholar] [CrossRef] [Green Version]
  31. Nagy, Á.; Levy, M. Variational density-functional theory for degenerate excited states. Phys. Rev. A 2001, 63, 052502. [Google Scholar] [CrossRef]
  32. Samal, P.; Harbola, M.K.; Holas, A. Density-to-potential map in time-independent excited-state density-functional theory. Chem. Phys. Lett. 2006, 419, 217–222. [Google Scholar] [CrossRef] [Green Version]
  33. Harbola, M.K.; Samal, P. Time-independent excited-state density functional theory: Study of 1s22p3(4S) and 1s22p3(2D) states of the boron isoelectronic series up to Ne5+. J. Phys. B 2009, 42, 015003. [Google Scholar] [CrossRef]
  34. Görling, A. Density-functional theory beyond the Hohenberg-Kohn theorem. Phys. Rev. A 1999, 59, 3359–3374. [Google Scholar] [CrossRef]
  35. Görling, A. Proper Treatment of Symmetries and Excited States in a Computationally Tractable Kohn-Sham Method. Phys. Rev. Lett. 2000, 85, 4229–4232. [Google Scholar] [CrossRef] [PubMed]
  36. Sahni, V.; Massa, L.; Singh, R.; Slamet, M. Quantal Density Functional Theory of Excited States. Phys. Rev. Lett. 2001, 87, 113002. [Google Scholar] [CrossRef] [Green Version]
  37. Petersilka, M.; Gossmann, U.J.; Gross, E.K.U. Excitation Energies from Time-Dependent Density-Functional Theory. Phys. Rev. Lett. 1996, 76, 1212–1215. [Google Scholar] [CrossRef] [Green Version]
  38. Casida, M.E. Time-dependent density-functional theory for molecules and molecular solids. J. Mol. Struct. Theochem 2009, 914, 3–18. [Google Scholar] [CrossRef]
  39. Appel, H.; Gross, E.K.U.; Burke, K. Excitation Energies from Time-Dependent Density-Functional Theory. Phys. Rev. Lett. 2010, 90, 043005. [Google Scholar] [CrossRef]
  40. Ayers, P.W.; Levy, M.; Nagy, Á. Time-independent density-functional theory for excited states of Coulomb systems. Phys. Rev. A 2012, 85, 042518. [Google Scholar] [CrossRef] [Green Version]
  41. Ayers, P.W.; Levy, M.; Nagy, Á. Kohn-Sham theory for excited states of Coulomb systems. J. Chem. Phys. 2015, 143, 191101. [Google Scholar] [CrossRef]
  42. Ayers, P.W.; Levy, M.; Nagy, Á. Time-independent density functional theory for degenerate excited states of Coulomb systems. Theor. Chim. Account. 2018, 137, 152. [Google Scholar] [CrossRef]
  43. Nagy, Á. Coordinate Scaling in Time-independent Excited-state Density Functional Theory for Coulomb Systems. Computation 2019, 7, 59. [Google Scholar] [CrossRef] [Green Version]
  44. Nagy, Á. Theories, Principles and Approaches. In Chemical Reactivity; von Szentpaly, L., Kaya, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; in press. [Google Scholar]
  45. Sharp, R.T.; Horton, G.K. A Variational Approach to the Unipotential Many-Electron Problem. Phys. Rev. 1953, 90, 317. [Google Scholar] [CrossRef]
  46. Talman, J.D.; Shadwick, W.F. Optimized effective atomic central potential. Phys. Rev. A 1976, 14, 36–40. [Google Scholar] [CrossRef]
  47. Della-Sala, F.; Görling, A. Efficient localized Hartree—Fock methods as effective exact-exchange Kohn—Sham methods for molecules. J. Chem. Phys. 2001, 115, 5718–5732. [Google Scholar] [CrossRef]
  48. Gritsenko, O.; Baerends, E.J. Exchange kernel of density functional response theory from the common energy denominator approximation (CEDA) for the Kohn—Sham Green’s function. Res. Chem. Intermed. 2004, 30, 87–98. [Google Scholar] [CrossRef]
  49. Staroverov, V.N.; Scuseria, G.E.; Davidson, E.R. Effective local potentials for orbital-dependent density functionals. J. Chem. Phys. 2006, 125, 081104. [Google Scholar] [CrossRef]
  50. Krieger, J.B.; Li, Y.; Iafrate, G.J. Derivation and application of an accurate Kohn-Sham potential with integer discontinuity. Phys. Lett. A 1990, 146, 256–260. [Google Scholar] [CrossRef]
  51. Krieger, J.B.; Li, Y.; Iafrate, G.J. Construction and application of an accurate local spin-polarized Kohn-Sham potential with integer discontinuity: Exchange-only theory. Phys. Rev. A 1992, 45, 101–126. [Google Scholar] [CrossRef] [PubMed]
  52. Krieger, J.B.; Li, Y.; Iafrate, G.J. Systematic approximations to the optimized effective potential: Application to orbital-density-functional theory. Phys. Rev. A 1992, 46, 5453–5458. [Google Scholar] [CrossRef] [PubMed]
  53. Nagy, Á. Alternative derivation of the Krieger-Li-Iafrate approximation to the optimized-effective-potential method. Phys. Rev. A 1997, 55, 3465–3468. [Google Scholar] [CrossRef]
  54. Wigner, E. Effects of the electron interaction on the energy levels of electrons in metals. Trans. Faraday Soc. 1938, 34, 678–685. [Google Scholar] [CrossRef]
  55. Kato, T. On the eigenfunctions of many particle systems in quantum mechanics. Commun. Pure Appl. Math. 1957, 10, 151–177. [Google Scholar] [CrossRef]
  56. Steiner, E. Charge Densities in Atoms. J. Chem. Phys. 1963, 39, 2365–2366. [Google Scholar] [CrossRef]
  57. March, N.H. Self-Consistent Fields in Atoms; Pergamon Press: Oxford, UK, 1975. [Google Scholar]
  58. Nagy, Á.; Sen, K.D. Exact results on the curvature of the electron density at the cusp in certain highly excited states of atoms. Chem. Phys. Lett. 2000, 332, 154–158. [Google Scholar] [CrossRef]
  59. Ayers, P.W. Density per particle as a descriptor of Coulombic systems. Proc. Natl. Acad. Sci. USA 2000, 97, 1959–1964. [Google Scholar] [CrossRef] [Green Version]
  60. Nagy, Á.; Sen, K.D. Higher-order cusp of the density in certain highly excited states of atoms and molecules. J. Phys. B 2000, 33, 1745–1752. [Google Scholar] [CrossRef]
  61. Nagy, Á.; Sen, K.D. Ground- and excited-state cusp conditions for the electron density. J. Chem. Phys. 2001, 115, 6300–6308. [Google Scholar] [CrossRef]
  62. Süle, P.; Nagy, Á. Comparative test of local and nonlocal Wigner-like correlation energy functionals. Acta Phys. Chem. Debr. 1994, 29, 31. [Google Scholar]
  63. Zhou, Z.; Chu, S. Spin-dependent localized Hartree-Fock density-functional calculation of singly, doubly, and triply excited and Rydberg states of He- and Li-like ions. Phys. Rev. A 2005, 71, 022513. [Google Scholar] [CrossRef] [Green Version]
  64. Glushkov, V.N.; Levy, M. Highly Excited States from a Time Independent Density Functional Method. Computation 2016, 4, 28. [Google Scholar] [CrossRef] [Green Version]
  65. Glushkov, V.N. Orthogonality of determinant functions in the Hartree-Fock method for highly excited electronic states. Opt. Spectrosc. 2015, 119, 1–5. [Google Scholar] [CrossRef]
  66. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Roy, A.K.; Chu, S.I. Density-functional calculations on singly and doubly excited Rydberg states of many-electron atoms. Phys. Rev. A 2002, 65, 052508. [Google Scholar] [CrossRef] [Green Version]
  68. Harbola, M.K.; Sahni, V. Quantum-Mechanical Interpretation of the Exchange-Correlation Potential of Kohn-Sham Density-Functional Theory. Phys. Rev. Lett. 1989, 62, 489–492. [Google Scholar] [CrossRef] [PubMed]
  69. Sahni, V.; Li, Y.; Harbola, M.K. Atomic structure in the Pauli-correlated approximation. Phys. Rev. A 1992, 45, 1434–1448. [Google Scholar] [CrossRef]
  70. Puchalski, M.; Kedziera, D.; Pachucki, K. Ionization potential for excited S states of the lithium atom. Phys. Rev. A 2010, 82, 062509. [Google Scholar] [CrossRef]
  71. Kramida, A.; Ralchenko, Y.; Reader, J.; NIST ASD Team. NIST Atomic Spectra Database (ver. 5.8); National Institute of Standards and Technology: Gaithersburg, MD, USA, 2020. Available online: https://physics.nist.gov/asd (accessed on 8 May 2021). [CrossRef]
  72. Orestes, E.; da Silva, A.B.F.; Capelle, K. Excitation energies from ground-state density-functionals by means of generator coordinates. Phys. Chem. Chem. Phys. 2009, 11, 4564–4569. [Google Scholar] [CrossRef]
  73. Staroverov, V.N.; Glushkov, V.N. Effective local potentials for excited states. J. Chem. Phys. 2010, 133, 244104. [Google Scholar] [CrossRef]
  74. Tatewaki, H.; Koga, T.; Sakai, Y.; Thakkar, A.J. Numerical Hartree-Fock energies of low-lying excited states of neutral atoms with Z≤18. J. Chem. Phys. 1994, 101, 4945–4948. [Google Scholar] [CrossRef]
  75. Casida, M.E. Generalization of the Optimized Effective Potential Model to Include Electron Correlation: A Variational Derivation of the Sham–Schluter Equation for the Exact Exchange-Correlation Potential. Phys. Rev. A 1995, 51, 2005–2013. [Google Scholar] [CrossRef]
  76. Grabowski, I.; Hirata, S.; Ivanov, S.; Bartlett, R.J. Ab initio density functional theory: OEP-MBPT(2). A new orbital-dependent correlation functional. J. Chem. Phys. 2002, 116, 4415–4425. [Google Scholar] [CrossRef]
  77. Nagy, Á. Relative information in excited state orbital free density functional theory. Int. J. Quantum Chem. 2020, 120, e26405. [Google Scholar] [CrossRef]
  78. Ghosh, S.K.; Berkowitz, M.; Parr, R.G. Transcription of ground-state density-functional theory into a local thermodynamics. Proc. Natl. Acad. Sci. USA 1984, 81, 8028–8031. [Google Scholar] [CrossRef] [Green Version]
  79. Nagy, Á. Excited-state Density Functional Theory of Coulomb systems. J. Chem. Phys. 2020, 153, 154103. [Google Scholar] [CrossRef] [PubMed]
  80. Liu, S.B. Information-Theoretic Approach in Density Functional Reactivity Theory. Acta Phys. Chim. Sin. 2016, 32, 98–118. [Google Scholar] [CrossRef]
  81. Rong, C.; Wang, B.; Zhao, D.; Liu, S.B. Information theoretic approach in density functional theory and its recent applications to chemical problems. WIREs Comput. Mol. Sci. 2019, 10, e1461. [Google Scholar] [CrossRef]
  82. Geerlings, P.; Chamorro, E.; Chattaraj, P.K.; De Proft, F.; Gázquez, J.L.; Liu, S.B.; Morell, C.; Toro-Labbé, A.; Vela, A.; Ayers, P.W. Conceptual density functional theory: Status, prospects, issues. Theor. Chem. Acc. 2020, 139, 36. [Google Scholar] [CrossRef]
Table 1. Total energies in Rydberg units for the ground and some excited states of the Li atom.
Table 1. Total energies in Rydberg units for the ground and some excited states of the Li atom.
StateKLIKLILWSLHFSLHFcWFLYPxCOEPHFExact
1 s 2 2 s −14.8643−14.9960−14.8650−14.9744 −14.8634−14.8655−14.9561
1 s 2 3 s −14.6198−14.7450−14.6201−14.7191−14.7155−14.6146−14.6204−14.7082
1 s 2 4 s −14.5494−14.6726−14.5496−14.6463−14.6396−14.5463−14.5498−14.6371
1 s 2 5 s −14.5197−14.6421−14.5198−14.6157−14.6093−14.5145−14.5200−14.6071
1 s 2 6 s −14.5045−14.6264−14.5045−14.6000 −14.4984−14.5046−14.5917
1 s 2 7 s −14.4956−14.6173−14.4957−14.5909 −14.4857−14.4957−14.5828
Table 2. Calculated and experimental excitation energies of the Li atom in Rydberg units.
Table 2. Calculated and experimental excitation energies of the Li atom in Rydberg units.
StateKLIKLILWSLHFSLHFcxCOEPHFTDDFTExactExp
1 s 2 3 s 0.24450.25100.24490.25530.24880.24500.22800.24790.2479
1 s 2 4 s 0.31490.32340.31540.32810.31720.3157 0.31910.3191
1 s 2 5 s 0.34460.35390.34520.35870.34890.3455 0.34900.3490
1 s 2 6 s 0.35980.36960.36050.37440.36510.3608 0.36440.3644
1 s 2 7 s 0.36860.37870.36930.38350.37780.3697 0.37330.3733
Table 3. Total energies in Rydberg units for the ground and some excited states of the Na atom.
Table 3. Total energies in Rydberg units for the ground and some excited states of the Na atom.
StateKLIKLILWELPHFNHFExact
[ N e ] 3 s −323.7106−324.5443−323.7116−323.7174−323.7178−324.5092
[ N e ] 4 s −323.4879−324.3153−323.4894−323.4938
[ N e ] 5 s −323.4224−324.2479−323.4232−323.4376
[ N e ] 6 s −323.3944−324.2190
[ N e ] 7 s −323.3798−324.2040
Table 4. Calculated and experimental excitation energies of the Na atom in Rydberg units.
Table 4. Calculated and experimental excitation energies of the Na atom in Rydberg units.
StateKLIKLILWELPHFExp
[ N e ] 4 s 0.22280.22900.22220.22360.2346
[ N e ] 5 s 0.28830.29640.28840.28980.3025
[ N e ] 6 s 0.31630.3254 0.3315
[ N e ] 7 s 0.33080.3403 0.3464
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nagy, Á. Density Functional Theory of Highly Excited States of Coulomb Systems. Computation 2021, 9, 73. https://0-doi-org.brum.beds.ac.uk/10.3390/computation9060073

AMA Style

Nagy Á. Density Functional Theory of Highly Excited States of Coulomb Systems. Computation. 2021; 9(6):73. https://0-doi-org.brum.beds.ac.uk/10.3390/computation9060073

Chicago/Turabian Style

Nagy, Ágnes. 2021. "Density Functional Theory of Highly Excited States of Coulomb Systems" Computation 9, no. 6: 73. https://0-doi-org.brum.beds.ac.uk/10.3390/computation9060073

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop