This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exposing minimal composition of Kohn-Sham theory and its extendability

H. Nakada [email protected] Department of Physics, Graduate School of Science, Chiba University,
Yayoi-cho 1-33, Inage, Chiba 263-8522, Japan
Abstract

Reducing the many-fermion problem to a set of single-particle (s.p.) equations, the Kohn-Sham (KS) theory has provided a practical tool to implement ab initio calculations of ground-state energies and densities in many-electron systems. There have been attempts to extend the KS theory so that it could describe other physical quantities, or it could be applied to other many-fermion systems. By generalizing and reformulating the KS theory in terms of the 1-body density matrix, we expose the minimal composition of the theory that enables the reduction of the many-fermion problem to the s.p. equations. Based on the reformulation, several basic issues are reconsidered. The vv- and NN-representabilities for the KS theory are distinguished from those for the Hohenberg-Kohn theorem. Criteria for the extendability of the KS theory are addressed.

Many-fermion systems, Kohn-Sham theory, Density matrix

I Introduction

The density-functional theory (DFT) successfully described many-electron systems ranging from atoms and molecules to solids Parr and Yang (1989); Giuliani and Vignale (2005); Engel and Dreizler (2011); Sahni (2016), in which the influence of the ions is handled as external fields via the Born-Oppenheimer approximation. Located at its center, the Hohenberg-Kohn (HK) theorem Hohenberg and Kohn (1964) has provided the firm basis of the DFT, and the Kohn-Sham (KS) theory Kohn and Sham (1965) has made the DFT practical for electronic systems. They have supplied an ab initio theory for ground-state (g.s.) energy E0E_{0}. The energy of system EE is represented as a functional of the local density n(𝐫)n(\mathbf{r}), and E0E_{0} is obtained via the variation of the energy density functional (EDF).

The KS theory has been attempted to be extended in various directions. In electronic systems, the EDFs including variables additional to n(𝐫)n(\mathbf{r}) were developed: for instance, the spin DFT von Barth and Hedin (1972); Gunnarsson and Lundqvist (1976); Jacob and Reiher (2012), the current DFT Vignale and Rasolt (1987, 1988), the DFT for superconductors Oliveira et al. (1988); Lüders et al. (2005), and the density-matrix functional theory Donnely and Parr (1978); Zumbach and Maschke (1985); Schindlmayr and Godby (1995); López-Sandoval and Pastor (2000); Requistl and Pankratov (2008); Pernal and Giesbertz (2016). A systematic method to involve additional quantities has been proposed Higuchi and Higuchi (2004). The DFT has also been extended to finite-temperature problems Mermin (1965) and to excitations via a time-dependent formalism Runge and Gross (1984); Gross and Kohn (1985), although we focus on the g.s. properties in this paper. Another direction is to extend the KS theory to other many-fermion systems. For atomic nuclei composed of nucleons, the self-consistent mean-field (SCMF) calculations have been implemented Vautherin and Brink (1972) by using effective interactions in which correlation effects are incorporated. The most popular form of effective interactions is the Skyrme interaction Skyrme (1959), representing energy via quasi-local currents, i.e., currents that depend only on a position 𝐫\mathbf{r}, by attributing influences of non-locality to momentum-dependence. These SCMF approaches have been reinterpreted in the context of the DFT Petkov and Stoitsov (1991); Colò (2020). Calculations using finite-range effective interactions, which lead to non-local EDFs, have been developed as well Dechargé and Gogny (1980); Nakada (2003). The SCMF approaches incorporating relativity Serot and Walecka (1986) are also renamed the covariant DFT Ring (2016).

The KS theory’s remarkable and essential property is that the many-fermion problem is reduced to a set of single-particle (s.p.) equations. In this paper, we shall generalize and reformulate the KS theory, aiming to expose the minimal composition of the theory, i.e., with what ingredients at minimum the many-fermion problem can be solved by s.p. equations. Through the reformulation, criteria for the extendability of the theory will be elucidated. They help to judge what physical quantities can be covered by extending it and to what system it can be extended.

The so-called orbital-free (OF) DFT, which has a root in the historical work by Thomas and Fermi Thomas (1927); Fermi (1928), has been developed Deb and Chattaraj (1988, 1992); Pearson et al. (1993) to reduce the computational cost further. Whereas the OF DFT is often considered an approximation of the KS DFT, it does not need s.p. equations and should be distinguished from the KS theory in this respect. As one of our central issues is the emergence of the s.p. equations, we will not handle the OF DFT in this paper, except for a brief referral.

We restrict our discussion to systems where the particle number NN is conserved. An arbitrary normalized state in the system is represented by |Ψ|\Psi\rangle. Denoting the Hamiltonian of this system by HH, the energy expectation value is regarded as a functional of |Ψ|\Psi\rangle,

E[Ψ]:=Ψ|H|Ψ.E[\Psi]:=\langle\Psi|H|\Psi\rangle\,. (1)

The variational principle tells us that E0E_{0} is equal to the minimum111 The expression minf(x)\min f(x) means the global minimum of f(x)f(x) as usual, throughout this paper. of E[Ψ]E[\Psi] with respect to |Ψ|\Psi\rangle,

E0=minΨNE[Ψ],E_{0}=\min_{\Psi\to N}E[\Psi]\,, (2)

where the subscript ΨN\Psi\to N stands for the condition that the state |Ψ|\Psi\rangle should belong to the NN-particle system. However, the underlying Hamiltonian is not well established in certain systems; the atomic nuclei are among them. For this reason, E[Ψ]E[\Psi] and the variational principle (2) are placed at the starting point in the arguments below, not necessarily presuming HH. Moreover, we shall present the formulation in terms of the 1-body density matrix, which carries properties of |Ψ|\Psi\rangle. While n(𝐫)n(\mathbf{r}) is adopted as the principal variable for the variation in the original HK theorem, we merely assume that the principal variables are all s.p. quantities, viz., functions of the 1-body density matrix, not restricting to n(𝐫)n(\mathbf{r}).

II Hartree-Fock approximation

The KS theory significantly overlaps with the SCMF approximations. Both approaches reduce many-fermion problems to a set of s.p. equations. In this section we review the Hartree-Fock (HF) approximation, which is a typical and rigorous example of the SCMF theories, in terms of the 1-body density matrix Ring and Schuck (1980). The arguments here will clarify the basic structure of the SCMF theories, which is possessed by the KS theory as well. Another example, the Hartree-Fock-Bogolyubov (HFB) approximation, is summarized in Appendix A.

Let {ϕk}\{\phi_{k}\} represent a complete orthonormal set of s.p. bases. The creation and annihilation operators of the basis state kk are denoted by aka^{\dagger}_{k} and aka_{k}. The 1-body density matrix222 This density matrix is called first-order reduced density matrix in some literature. ϱk\varrho_{k\ell} for an arbitrary normalized state |Ψ|\Psi\rangle is defined by

ϱk=ϱk[Ψ]:=Ψ|aak|Ψ=ϱk.\varrho_{k\ell}=\varrho_{k\ell}[\Psi]:=\langle\Psi|a^{\dagger}_{\ell}a_{k}|\Psi\rangle=\varrho_{\ell k}^{\ast}\,. (3)

The ν\nu-body density matrix Ψ|a1a2aνakνak2ak1|Ψ\langle\Psi|a^{\dagger}_{\ell_{1}}a^{\dagger}_{\ell_{2}}\cdots a^{\dagger}_{\ell_{\nu}}a_{k_{\nu}}\cdots a_{k_{2}}a_{k_{1}}|\Psi\rangle is decomposed into products of ϱk\varrho_{k\ell} and the correlation functions 𝒞(ν)\mathcal{C}^{(\nu^{\prime})} (ν=2,3,,ν\nu^{\prime}=2,3,\cdots,\nu), which obey the Bogolyubov-Born-Green-Kirkwood-Yvon hierarchy Reichl (1980). For instance, the 2-body correlation function is defined by

𝒞kk(2):=Ψ|aaakak|Ψϱkϱk+ϱkϱk.\mathcal{C}^{(2)}_{kk^{\prime}\ell\ell^{\prime}}:=\langle\Psi|a^{\dagger}_{\ell}a^{\dagger}_{\ell^{\prime}}a_{k^{\prime}}a_{k}|\Psi\rangle-\varrho_{k\ell}\varrho_{k^{\prime}\ell^{\prime}}+\varrho_{k\ell^{\prime}}\varrho_{\ell k^{\prime}}\,. (4)

E[Ψ]E[\Psi] can be represented by ϱ\varrho and 𝒞(ν)\mathcal{C}^{(\nu)},

E[Ψ]=E[ϱ,𝒞(2),𝒞(3),,𝒞(N)].E[\Psi]=E[\varrho,\mathcal{C}^{(2)},\mathcal{C}^{(3)},\cdots,\mathcal{C}^{(N)}]\,. (5)

By neglecting 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2), the energy functional in the HF approximation is obtained, E[Ψ]EHF[ϱ]E[\Psi]\approx E^{\mathrm{HF}}[\varrho]. Variation of EHF[ϱ]E^{\mathrm{HF}}[\varrho] for ϱk\varrho_{k\ell} yields a s.p. Hamiltonian (HF Hamiltonian),

hkHF:=EHFϱk=(hkHF).h^{\mathrm{HF}}_{k\ell}:=\frac{\partial E^{\mathrm{HF}}}{\partial\varrho_{\ell k}}=\big{(}h^{\mathrm{HF}}_{\ell k}\big{)}^{\ast}\,. (6)

The diagonalization of hHFh^{\mathrm{HF}} with an appropriate unitary transformation 𝒰HF\mathcal{U}^{\mathrm{HF}} yields

hkHF𝒰iHF=ϵiHF𝒰kiHF,\sum_{\ell}h^{\mathrm{HF}}_{k\ell}\,\mathcal{U}^{\mathrm{HF}}_{\ell i}=\epsilon^{\mathrm{HF}}_{i}\,\mathcal{U}^{\mathrm{HF}}_{ki}\,, (7)

which is identical to the HF equation, determining s.p. energy ϵiHF\epsilon^{\mathrm{HF}}_{i} and wave-function φiHF=k𝒰kiHFϕk\varphi^{\mathrm{HF}}_{i}=\sum_{k}\mathcal{U}^{\mathrm{HF}}_{ki}\,\phi_{k}.

The HF state |ΦHF|\Phi^{\mathrm{HF}}\rangle is obtained by minimizing EHFE^{\mathrm{HF}} with respect to ϱk\varrho_{k\ell} under the particle-number condition,

tr(ϱ)=N.\mathrm{tr}(\varrho)=N\,. (8)

The trace in Eq. (8) is taken in the s.p. space. To implement the minimization, Eq. (8) is handled with the Lagrange multiplier μ\mu, modifying EHFE^{\mathrm{HF}} as

E~HF:=EHFμ[tr(ϱ)N].\tilde{E}^{\mathrm{HF}}:=E^{\mathrm{HF}}-\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,. (9)

Variation of E~HF\tilde{E}^{\mathrm{HF}} yields

δE~HF=kEHFϱkδϱkμkδϱkkδμ[tr(ϱ)N]=tr[(hHFμ)δϱ]δμ[tr(ϱ)N].\begin{split}\delta\tilde{E}^{\mathrm{HF}}&=\sum_{k\ell}\frac{\partial E^{\mathrm{HF}}}{\partial\varrho_{\ell k}}\,\delta\varrho_{\ell k}-\mu\sum_{k}\delta\varrho_{kk}-\delta\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\\ &=\mathrm{tr}\big{[}(h^{\mathrm{HF}}-\mu)\,\delta\varrho\big{]}-\delta\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,.\end{split} (10)

We should have333 Note that δE~HF=0\delta\tilde{E}^{\mathrm{HF}}=0 is not necessarily satisfied because ϱk\varrho_{k\ell} does not vary freely, restricted by the bounds of Eq. (11). See also Footnote 4. δE~HF0\delta\tilde{E}^{\mathrm{HF}}\geq 0 at the minimum of EHFE^{\mathrm{HF}}.

The Pauli principle requires

0ϱii1for any representation.0\leq\varrho_{ii}\leq 1\quad\mbox{for any representation}\,. (11)

By combining δE~HF0\delta\tilde{E}^{\mathrm{HF}}\geq 0 with Eq. (11), it is concluded that |ΦHF|\Phi^{\mathrm{HF}}\rangle should be a state corresponding to a single Slater determinant, as proven below. In δE~HF\delta\tilde{E}^{\mathrm{HF}} of Eq. (10), the δμ\delta\mu term vanishes as long as Eq. (8) is fulfilled. Suppose that {ϱkHF(:=ϱk[ΦHF])}\{\varrho^{\mathrm{HF}}_{k\ell}\,(:=\varrho_{k\ell}[\Phi^{\mathrm{HF}}])\} gives the minimum of EHFE^{\mathrm{HF}} under the constraint (8), which is equal to the minimum of E~HF\tilde{E}^{\mathrm{HF}}. The s.p. energy ϵiHF\epsilon^{\mathrm{HF}}_{i} and wave-function φiHF\varphi^{\mathrm{HF}}_{i} are determined accordingly. It is convenient to work with the s.p. state ii instead of the basis kk in the vicinity of the HF solution. In this vicinity, Eq. (10) is expressed as

δE~HF=i(ϵiHFμ)δϱii.\delta\tilde{E}^{\mathrm{HF}}=\sum_{i}(\epsilon^{\mathrm{HF}}_{i}-\mu)\,\delta\varrho_{ii}\,. (12)

The requirement δE~HF0\delta\tilde{E}^{\mathrm{HF}}\geq 0 at the minimum of EHFE^{\mathrm{HF}} derives

{δϱii0for i with ϵiHF>μδϱii0for i with ϵiHF<μ.\left\{\begin{array}[]{ll}\delta\varrho_{ii}\geq 0&\mbox{for $i$ with }\epsilon^{\mathrm{HF}}_{i}>\mu\\ \delta\varrho_{ii}\leq 0&\mbox{for $i$ with }\epsilon^{\mathrm{HF}}_{i}<\mu\end{array}\right.\,. (13)

Combined with (11), Eq. (13) leads to

{ϱiiHF=0for i with ϵiHF>μϱiiHF=1for i with ϵiHF<μ,\left\{\begin{array}[]{ll}\varrho^{\mathrm{HF}}_{ii}=0&\mbox{for $i$ with }\epsilon^{\mathrm{HF}}_{i}>\mu\\ \varrho^{\mathrm{HF}}_{ii}=1&\mbox{for $i$ with }\epsilon^{\mathrm{HF}}_{i}<\mu\end{array}\right.\,, (14)

so that negative δϱii\delta\varrho_{ii} should not be allowed for ϵiHF>μ\epsilon^{\mathrm{HF}}_{i}>\mu and likewise for ϵiHF<μ\epsilon^{\mathrm{HF}}_{i}<\mu. Since ϱii\varrho_{ii} is the occupation number on the s.p. state ii, the lowest NN s.p. states are occupied in |ΦHF|\Phi^{\mathrm{HF}}\rangle unless there exist s.p. levels with ϵiHF=μ\epsilon^{\mathrm{HF}}_{i}=\mu, which could be degenerate. Even if there are several levels having ϵiHF=μ\epsilon^{\mathrm{HF}}_{i}=\mu, the occupation number on them can be set integer, viz., 0 or 11, through a unitary transformation among the degenerate s.p. states. Therefore |ΦHF|\Phi^{\mathrm{HF}}\rangle can be represented by a single Slater determinant, which is formed by {φiHF}\{\varphi^{\mathrm{HF}}_{i}\} with ϱiiHF=1\varrho^{\mathrm{HF}}_{ii}=1. As the ϵiHF=μ\epsilon^{\mathrm{HF}}_{i}=\mu levels do not influence essential points but make arguments lengthy, we shall assume ϵiHFμ\epsilon^{\mathrm{HF}}_{i}\neq\mu for any ii below.

Because of Condition (11), ϱiiHF\varrho^{\mathrm{HF}}_{ii} of Eq. (14) has reached either its maximum or minimum that can be obtained by any unitary transformation, implying that ϱHF\varrho^{\mathrm{HF}} is diagonalized in this representation. When its eigenvalues are only 0 and 11, the density matrix ϱ\varrho is characterized by the idempotency444 By regarding Eq. (15) as a constraint and defining E~~HF:=E~HFkλk(ϱ2ϱ)k\tilde{\tilde{E}}^{\mathrm{HF}}:=\tilde{E}^{\mathrm{HF}}-\sum_{k\ell}\lambda_{k\ell}\,(\varrho^{2}-\varrho)_{k\ell} with the Lagrange multiplier λk\lambda_{k\ell}, the minimum of EHFE^{\mathrm{HF}} can be searched via δE~~HF=0\delta\tilde{\tilde{E}}^{\mathrm{HF}}=0, compatible with the HF equation. ,

ϱ2=ϱ,viz.,mϱkmϱm=ϱk.\varrho^{2}=\varrho\,,\quad\mbox{viz.,}\quad\sum_{m}\varrho_{km}\,\varrho_{m\ell}=\varrho_{k\ell}\,. (15)

A state given by a single Slater determinant |Φ|\Phi\rangle provides ϱk=Φ|aak|Φ\varrho_{k\ell}=\langle\Phi|a^{\dagger}_{\ell}a_{k}|\Phi\rangle that fulfills Eq. (15), as is evident by representing it through the s.p. states {φi;i=1,2,,N}\{\varphi_{i};i=1,2,\cdots,N\} forming |Φ|\Phi\rangle. Conversely, if {ϱk}\{\varrho_{k\ell}\} satisfies Eq. (15), the corresponding state is given by a Slater determinant. This assertion is easily proven by constructing the Slater determinant: take the s.p. states {φi}\{\varphi_{i}\} that diagonalize ϱ\varrho and pick up the states corresponding to the eigenvalue 11. Thus, ϱHF\varrho^{\mathrm{HF}} satisfying Eq. (15) and |ΦHF|\Phi^{\mathrm{HF}}\rangle is represented by a single Slater determinant. Let us denote the full Hilbert space of the NN-particle system by 𝒱full\mathcal{V}_{\mathrm{full}}, and the subspace composed of the states of the Slater determinants by 𝒱idem\mathcal{V}_{\mathrm{idem}}, which are characterized by Eq. (15). It is noted that, although EHF[ϱ]E^{\mathrm{HF}}[\varrho] is defined in 𝒱full\mathcal{V}_{\mathrm{full}}, the wave-function necessarily falls on 𝒱idem\mathcal{V}_{\mathrm{idem}}, as the result of the variation. This fact is equivalent to the conclusion in Ref. Lieb (1981).

III Hohenberg-Kohn theorem

We present the Hohenberg-Kohn (HK) theorem in this section. Whereas it is no more than a summary of many previous works, we state and discuss the HK theorem in a generalized context, preparing to reformulate the KS theory.

Consider a set of variables QAQ^{A}, which are called principal variables in this article. The superscript AA distinguishes the variables, which can be continuous. For brevity, the set of QAQ^{A} is denoted by 𝐐\mathbf{Q}. In the original HK theorem Hohenberg and Kohn (1964), 𝐐\mathbf{Q} is the local density n(𝐫):=Ψ|i=1Nδ(𝐫𝐫i)|Ψn(\mathbf{r}):=\langle\Psi|\sum_{i=1}^{N}\delta(\mathbf{r}-\mathbf{r}_{i})|\Psi\rangle, where ii is the index of particles555 The index ii is commonly used for labeling particles and s.p. states after diagonalization, which correspond to each other. . As mentioned in Introduction, additional variables have been introduced in 𝐐\mathbf{Q} when extending the DFT. In Ref. Gilbert (1975), the HK theorem has extensively been proven for the whole 1-body density matrix, 𝐐={ϱk}\mathbf{Q}=\{\varrho_{k\ell}\}.

The HK theorem (the so-called 2nd theorem) asserts the existence of the functional EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}], whose minimum gives exact g.s. energy and corresponding 𝐐\mathbf{Q}. The theorem was initially proven by showing the one-to-one correspondence of 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}) and the external potential (the 1st HK theorem), which is equivalent to the Legendre transformation. The Legendre transformation was also applied in the effective action formalism, deriving the HK theorem in an elegant manner Fukuda et al. (1994); Valiev and Fernando (1997). However, we follow Levy’s idea of the constrained search Levy (1979), which derives a universal energy functional of 𝐐\mathbf{Q} and provides a basis for the KS theory. The HK energy functional can be defined by

EHK[𝐐]:=minΨ𝐐E[Ψ].E^{\mathrm{HK}}[\mathbf{Q}]:=\min_{\Psi\to\mathbf{Q}}E[\Psi]\,. (16)

Here the expression Ψ𝐐\Psi\to\mathbf{Q} indicates the constraint that |Ψ|\Psi\rangle gives a fixed 𝐐\mathbf{Q}, under which the minimization with respect to |Ψ|\Psi\rangle is carried out. The energy functional of 𝐐\mathbf{Q} is extensively called EDF in this paper, though 𝐐\mathbf{Q} is not necessarily the density. Equation (16) guarantees the uniqueness of the EDF except for degeneracy, as it is defined as the global minimum under Ψ𝐐\Psi\to\mathbf{Q}. For later convenience, the state |Ψ|\Psi\rangle that minimizes the energy for a fixed 𝐐\mathbf{Q} is denoted by |Ψ𝐐HK|\Psi^{\mathrm{HK}}_{\mathbf{Q}}\rangle;

EHK[𝐐](=Ψ𝐐HK|H|Ψ𝐐HK)=E[Ψ𝐐HK].E^{\mathrm{HK}}[\mathbf{Q}]\,\big{(}=\langle\Psi^{\mathrm{HK}}_{\mathbf{Q}}|H|\Psi^{\mathrm{HK}}_{\mathbf{Q}}\rangle\big{)}=E[\Psi^{\mathrm{HK}}_{\mathbf{Q}}]\,. (17)

The minimum of EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] with respect to 𝐐\mathbf{Q} in the NN-particle space gives the exact g.s. energy,

E0=min𝐐NEHK[𝐐].E_{0}=\min_{\mathbf{Q}\to N}E^{\mathrm{HK}}[\mathbf{Q}]\,. (18)

The values of 𝐐\mathbf{Q} that give the minimum are also exact. The HK theorem means that the minimization of (2) can be separated into two steps, (16) and (18).

The representability problem should not be discarded, although this issue is not easy to be explored thoroughly. It is not trivial whether 𝒱full\mathcal{V}_{\mathrm{full}}, the full Hilbert space spanned by |Ψ|\Psi\rangle with NN particles, covers the whole domain of 𝐐\mathbf{Q} (viz., the NN-representability) Coleman (1963). This problem may limit the choice of 𝐐\mathbf{Q}. If 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}), it has been shown Harriman (1981) that there exists |Ψ|\Psi\rangle for any n(𝐫)n(\mathbf{r}) that fulfills n(𝐫)0n(\mathbf{r})\geq 0 and d3rn(𝐫)=N\int d^{3}r\,n(\mathbf{r})=N.

In the 1st HK theorem, n(𝐫)n(\mathbf{r}) is linked to the external s.p. potential via the Legendre transformation. In this case, the vv-representability, i.e., the existence of the s.p. potential conjugate to any physical n(𝐫)n(\mathbf{r}), should be assured. It has been pointed out that vv-representability is not always satisfied Levy (1982). Furthermore, the EDF via the Legendre transformation yields the global minimum only in a restricted region of 𝐐\mathbf{Q}. Examples in the spin DFT are found in Refs. von Barth and Hedin (1972); Capelle and Vignale (2001), and presented in Sec. VI.1. Thus the Legendre transformation derives an EDF that yields the g.s. energy only in a limited domain of 𝐐\mathbf{Q}. Notice that the EDF obtained from the Legendre transformation is not equivalent to EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] of Eq. (16) everywhere. They are different in the domain where the EDF of the Legendre transformation does not give the global minimum.

While the constrained search circumvents this problem, it induces another problem concerning differentiability. To carry out the minimization of Eq. (18), the condition δEHK[𝐐]0\delta E^{\mathrm{HK}}[\mathbf{Q}]\geq 0 (or δEHK[𝐐]=0\delta E^{\mathrm{HK}}[\mathbf{Q}]=0) is imposed, where the expression δ\delta contains differentiation with respect to 𝐐\mathbf{Q}, though we postpone specifying it concretely. Therefore EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] should be differentiable for this practical purpose. As will be shown in Sec. IV, a s.p. potential emerges by the differentiation. Indeed, the functional derivative of the EDF for n(𝐫)n(\mathbf{r}) yields a s.p. potential in the original discussion on the HK theorem. In this respect, the vv-representability remains after the constrained search, transferred to the differentiability of EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}]. The importance of differentiability, or regularity, has been pointed out in Refs. Capelle and Vignale (2001); Kvaal et al. (2014). As illustrated in Appendix B, regularity is often lost if optimization is separated into two steps. The two-step optimization is essential for the universality of EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] but damages regularity as a function (or functional) of QAQ^{A} in most cases. Without this regularity, the procedure of Eq. (18) is impractical.

IV Kohn-Sham theory

The Kohn-Sham (KS) theory is now reformulated and generalized666 Although it is a generalization of the original KS theory, we keep calling it KS theory, as some existing extensions may implicitly be subject to this generalization.. The irregularity of EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] mentioned above may occur as a quantum effect, as in the shell structure of atoms and nuclei. The argument arriving at Eq. (14) is indicative of this problem. Even if the particle number NN were a continuous variable, not restricted to integers, the condition (11) should lead to an irregularity because particles start occupying another s.p. state after one of the ϱii\varrho_{ii}’s reaches unity. This property originating from the Pauli principle is the source of the shell effect. Influence of the shell structure on the regularity of the EDF is discussed in Appendix C. This perception concerning the shell effect motivates us to deal with ϱ\varrho explicitly, so that the EDF could come well-behaved.

If NN is very large and the quantum effects are not strong, the irregularity due to the shell structure may be negligible. It is then possible to directly apply a variational calculation to EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}], deriving the orbital-free DFT, although it is not the current interest.

IV.1 Kohn-Sham equation

In the KS theory Kohn and Sham (1965), which realizes the HK theorem, the energy variation is reduced to equations defining a set of s.p. orbitals, called KS orbitals. This is a striking feature of the KS theory. To disclose what ingredients at minimum enable to solve the many-fermion problem by s.p. equations, the equations are rederived by means of the 1-body density matrix ϱ\varrho.

Impose QA=QA[ϱ]Q^{A}=Q^{A}[\varrho], i.e., all QAQ^{A}’s depend solely on the 1-body density matrix ϱ\varrho, not on 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2). Moreover, QAQ^{A}’s have to be independent of one another as a function of ϱ\varrho. The energy E[Ψ]=E[ϱ,𝒞(2),𝒞(3),,𝒞(N)]E[\Psi]=E[\varrho,\mathcal{C}^{(2)},\mathcal{C}^{(3)},\cdots,\mathcal{C}^{(N)}] is partitioned into two parts, regular and irregular parts with respect to 𝐐\mathbf{Q}. The irregular part is assumed to be represented only with ϱ\varrho, without 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2), and to be differentiable with respect to ϱ\varrho though not with respect to 𝐐\mathbf{Q}. The irregular part may explicitly contain QAQ^{A} without contradiction, because QA=QA[ϱ]Q^{A}=Q^{A}[\varrho]. The partitioning is then expressed as

E[Ψ]=EirrKS[ϱ;𝐐[ϱ]]+Ereg[Ψ].E[\Psi]=E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho;\mathbf{Q}[\varrho]\big{]}+E_{\mathrm{reg}}[\Psi]\,. (19)

This partitioning may correspond to that of the underlying Hamiltonian,

H=Hirr+Hreg;EirrKS[ϱ;𝐐[ϱ]]=Ψ|Hirr|Ψ,Ereg[Ψ]=Ψ|Hreg|Ψ,H=H_{\mathrm{irr}}+H_{\mathrm{reg}}\,;\quad E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho;\mathbf{Q}[\varrho]\big{]}=\langle\Psi|H_{\mathrm{irr}}|\Psi\rangle\,,~{}E_{\mathrm{reg}}[\Psi]=\langle\Psi|H_{\mathrm{reg}}|\Psi\rangle\,, (20)

although it is not requisite.

The individual theory is not defined until fixing 𝐐\mathbf{Q} and the partitioning (viz., EirrKSE^{\mathrm{KS}}_{\mathrm{irr}}). In the standard KS equation, 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}) and EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} corresponds to the kinetic energy term so that the shell effects could be tractable. This also holds in the example of Appendix C. However, the source of the irregularity is not necessarily restricted to the kinetic energy in general cases. The choice of 𝐐\mathbf{Q} and EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} is not always obvious; it may count on physical insight. Still, the structure of the KS theory can be discussed without specifying 𝐐\mathbf{Q} and EirrKSE^{\mathrm{KS}}_{\mathrm{irr}}, as long as the influence of 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2) on them is masked.

With the partitioning of Eq. (19), the result of the constrained search of Eq. (16) is written as

EHK[𝐐]=EirrKS[ϱ[Ψ𝐐HK];𝐐]+EregHK[𝐐];EregHK[𝐐]:=Ereg[Ψ𝐐HK].E^{\mathrm{HK}}[\mathbf{Q}]=E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}];\mathbf{Q}\big{]}+E^{\mathrm{HK}}_{\mathrm{reg}}[\mathbf{Q}]\,;\quad E^{\mathrm{HK}}_{\mathrm{reg}}\big{[}\mathbf{Q}]:=E_{\mathrm{reg}}\big{[}\Psi^{\mathrm{HK}}_{\mathbf{Q}}\big{]}\,. (21)

Whereas EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} in Eq. (21) is not determined until knowing ϱ[Ψ𝐐HK]\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}], it is a difficult task to obtain ϱ[Ψ𝐐HK]\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}], not more accessible than |Ψ𝐐HK|\Psi^{\mathrm{HK}}_{\mathbf{Q}}\rangle. A practical way is to minimize EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} for ϱ\varrho under the constraint that ϱ\varrho should provide a given 𝐐\mathbf{Q}, minϱ𝐐EirrKS[ϱ;𝐐]\min_{\varrho\to\mathbf{Q}}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}]. The resultant ϱ\varrho is not equal to ϱ[Ψ𝐐HK]\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}], and thereby minϱ𝐐EirrKS[ϱ;𝐐]EirrKS[ϱ[Ψ𝐐HK];𝐐]\min_{\varrho\to\mathbf{Q}}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}]\neq E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}];\mathbf{Q}\big{]}. However, this deviation is assumed to be a regular function of 𝐐\mathbf{Q}. Then, the deviation can be incorporated into the regular part of the EDF, deriving the modified partitioning as

EHK[𝐐]=minϱ𝐐EirrKS[ϱ;𝐐]+EregKS[𝐐];EregKS[𝐐]:=EregHK[𝐐]+(EirrKS[ϱ[Ψ𝐐HK];𝐐]minϱ𝐐EirrKS[ϱ;𝐐]).\begin{split}&E^{\mathrm{HK}}[\mathbf{Q}]=\min_{\varrho\to\mathbf{Q}}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}]+E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{Q}]\,;\\ &\quad E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{Q}]:=E^{\mathrm{HK}}_{\mathrm{reg}}[\mathbf{Q}]+\Big{(}E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}];\mathbf{Q}\big{]}-\min_{\varrho\to\mathbf{Q}}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}]\Big{)}\,.\end{split} (22)

Although the principal variables 𝐐\mathbf{Q} should be a function of the density matrix ϱ\varrho, it is convenient to handle them as if they were independent in EirrKSE^{\mathrm{KS}}_{\mathrm{irr}}. To do it, QAQ^{A} is replaced by an auxiliary variable qAq^{A}, with setting qA=QAq^{A}=Q^{A} as a constraint. The EDF within the KS scheme is now defined as follows, with the Lagrange multiplier ΛAKS\Lambda^{\mathrm{KS}}_{A},

EKS[ϱ;𝐪]=EirrKS[ϱ;𝐪]+EregKS[𝐪]AΛAKS(qAQA[ϱ]).E^{\mathrm{KS}}[\varrho;\mathbf{q}]=E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}]+E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{q}]-\sum_{A}\Lambda^{\mathrm{KS}}_{A}\,\Big{(}q^{A}-Q^{A}[\varrho]\Big{)}\,. (23)

This procedure is analogous to the Legendre transformation used in Refs. Fukuda et al. (1994); Valiev and Fernando (1997). Then, after the treatment analogous to Eq. (9),

E~KS:=EKSμ[tr(ϱ)N],\tilde{E}^{\mathrm{KS}}:=E^{\mathrm{KS}}-\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,, (24)

the condition δE~KS[ϱ;𝐪]0\delta\tilde{E}^{\mathrm{KS}}[\varrho;\mathbf{q}]\geq 0 should be fulfilled at the energy minimum, where the variation is performed for ϱ\varrho, 𝐪\mathbf{q} and μ\mu. The variation of EKSE^{\mathrm{KS}} for qAq^{A} yields

ΛAKS=EirrKS[ϱ;𝐪]qA+EregKS[𝐪]qA.\Lambda^{\mathrm{KS}}_{A}=\frac{\partial E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}\,. (25)

Here the differentiation with respect to qAq^{A} is expressed by the partial derivative, but it should read as a functional derivative when AA is continuous. The variation provides s.p. Hamiltonian,

hkKS:=EKSϱk=EirrKS[ϱ;𝐪]ϱk+ΓkKS;ΓkKS:=AΛAKSQAϱk=A(EirrKS[ϱ;𝐪]qA+EregKS[𝐪]qA)QAϱk.\begin{split}h^{\mathrm{KS}}_{k\ell}&:=\frac{\partial E^{\mathrm{KS}}}{\partial\varrho_{k\ell}}=\frac{\partial E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}]}{\partial\varrho_{k\ell}}+\Gamma^{\mathrm{KS}}_{k\ell}\,;\\ \Gamma^{\mathrm{KS}}_{k\ell}&:=\sum_{A}\Lambda^{\mathrm{KS}}_{A}\,\frac{\partial Q^{A}}{\partial\varrho_{k\ell}}=\sum_{A}\bigg{(}\frac{\partial E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}\bigg{)}\,\frac{\partial Q^{A}}{\partial\varrho_{k\ell}}\,.\end{split} (26)

ΓKS\Gamma^{\mathrm{KS}} is the KS potential. It is reasonable to assume that (QA)(Q^{A})^{\ast} is also contained in 𝐐\mathbf{Q} for a complex QAQ^{A}. Then hKSh^{\mathrm{KS}} is hermitian and therefore diagonalizable, having real eigenvalues,

hkKS𝒰iKS=ϵiKS𝒰kiKS,\sum_{\ell}h^{\mathrm{KS}}_{k\ell}\,\mathcal{U}^{\mathrm{KS}}_{\ell i}=\epsilon^{\mathrm{KS}}_{i}\,\mathcal{U}^{\mathrm{KS}}_{ki}\,, (27)

similarly to the HF equation (7). This is the KS equation, and the s.p. state diagonalizing hKSh^{\mathrm{KS}}, φiKS=k𝒰kiKSϕk\varphi^{\mathrm{KS}}_{i}=\sum_{k}\mathcal{U}^{\mathrm{KS}}_{ki}\,\phi_{k}, is the KS orbital. Because of the identical mathematical structure, all the arguments on the solutions of the HF equation in Sec. II hold for those of the KS equation (27) as they are. The resultant 1-body density matrix ϱKS\varrho^{\mathrm{KS}} satisfies Eq. (15), corresponding to the state |ΦKS|\Phi^{\mathrm{KS}}\rangle in which the lowest NN KS orbitals are occupied. However, while all QAQ^{A}’s are physical, individual ϱkKS\varrho^{\mathrm{KS}}_{k\ell} (equivalently, 𝒰kiKS\mathcal{U}^{\mathrm{KS}}_{ki} and φiKS\varphi^{\mathrm{KS}}_{i}) is artificial, losing clear physical meaning, as is recognized from the derivation.

In the conventional formulation of the KS theory, the subspace 𝒱idem\mathcal{V}_{\mathrm{idem}} is considered from the beginning by introducing the non-interacting reference system and the KS orbitals. On the other hand, the functional EKS[ϱ;𝐪]E^{\mathrm{KS}}[\varrho;\mathbf{q}] is here defined in the full Hilbert space 𝒱full\mathcal{V}_{\mathrm{full}}. Hence it can be identified as an effective Hamiltonian, although it depends on 𝐪\mathbf{q} and is not necessarily representable in the second-quantized form. Solutions of the KS equation fall on 𝒱idem\mathcal{V}_{\mathrm{idem}} as a result. Still, it is legitimate to state that the idempotent ϱ\varrho (ϱKS\varrho^{\mathrm{KS}}, to be precise) is implicitly assumed when postulating that EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} does not depend on 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2).

As already mentioned, the HK theorem was extended to the case 𝐐={ϱk}\mathbf{Q}=\{\varrho_{k\ell}\} Gilbert (1975), above which the density-matrix functional theory (DMFT) was exploited Donnely and Parr (1978); Zumbach and Maschke (1985); Schindlmayr and Godby (1995); López-Sandoval and Pastor (2000); Requistl and Pankratov (2008); Pernal and Giesbertz (2016). The present reformulation should be distinguished from the DMFT. It is noticed that EHK[ϱk]E^{\mathrm{HK}}[\varrho_{k\ell}] (i.e., 𝐐={ϱk}\mathbf{Q}=\{\varrho_{k\ell}\}) itself should not be handled in the KS theory because the 1-body density matrix of the KS theory ϱ(=ϱKS)\varrho\,(=\varrho^{\mathrm{KS}}) cannot be physical, as will be discussed in Sec. V. The ensemble representability Valone (1980) is irrelevant to ϱ\varrho here.

If EirrKS[ϱ;𝐪]E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}] consists of two parts, one which linearly depends on ϱ\varrho but does not depend explicitly on 𝐪\mathbf{q} and the other expressed only with 𝐪\mathbf{q},

EirrKS[ϱ;𝐪]=ktkϱk+ΔEirrKS[𝐪],E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{q}]=\sum_{k\ell}t_{k\ell}\,\varrho_{k\ell}+\mathit{\Delta}E^{\mathrm{KS}}_{\mathrm{irr}}[\mathbf{q}]\,, (28)

Eq. (26) becomes

hkKS=tk+ΓkKS;ΓkKS=A(ΔEirrKS[𝐪]qA+EregKS[𝐪]qA)QAϱk.h^{\mathrm{KS}}_{k\ell}=t_{k\ell}+\Gamma^{\mathrm{KS}}_{k\ell}\,;\quad\Gamma^{\mathrm{KS}}_{k\ell}=\sum_{A}\bigg{(}\frac{\partial\mathit{\Delta}E^{\mathrm{KS}}_{\mathrm{irr}}[\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}\bigg{)}\,\frac{\partial Q^{A}}{\partial\varrho_{k\ell}}\,. (29)

This is the case of the standard KS equation (see Sec. VI.1).

IV.2 Kohn-Sham–Bogolyubov-de Gennes equation

Irregularity of EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] may arise via correlations linked to a pair condensate, as in superconductivity. It is then reasonable to include the pairing tensor κ\kappa and κ\kappa^{\ast} in the irregular part of the EDF, and to extend Eq. (19) as

E[Ψ]=EirrKSBdG[ϱ,κ,κ;𝐐[ϱ,κ,κ]]+Ereg[Ψ].E[\Psi]=E^{\mathrm{KSBdG}}_{\mathrm{irr}}\big{[}\varrho,\kappa,\kappa^{\ast};\mathbf{Q}[\varrho,\kappa,\kappa^{\ast}]\big{]}+E_{\mathrm{reg}}[\Psi]\,. (30)

Here QAQ^{A}’s are postulated to depend on (ϱ,κ,κ)(\varrho,\kappa,\kappa^{\ast}). Arguments analogous to Eqs. (21) and (22) yield

EHK[𝐐]=min(ϱ,κ,κ)𝐐EirrKSBdG[ϱ,κ,κ;𝐐]+EregKSBdG[𝐐];EregKSBdG[𝐐]:=Ereg[Ψ𝐐HK]+(EirrKSBdG[ϱ[Ψ𝐐HK],κ[Ψ𝐐HK],κ[Ψ𝐐HK];𝐐]min(ϱ,κ,κ)𝐐EirrKSBdG[ϱ,κ,κ;𝐐]).\begin{split}&E^{\mathrm{HK}}[\mathbf{Q}]=\min_{(\varrho,\kappa,\kappa^{\ast})\to\mathbf{Q}}E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{Q}]+E^{\mathrm{KSBdG}}_{\mathrm{reg}}[\mathbf{Q}]\,;\\ &\quad E^{\mathrm{KSBdG}}_{\mathrm{reg}}\big{[}\mathbf{Q}]:=E_{\mathrm{reg}}\big{[}\Psi^{\mathrm{HK}}_{\mathbf{Q}}\big{]}+\Big{(}E^{\mathrm{KSBdG}}_{\mathrm{irr}}\big{[}\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}],\kappa[\Psi^{\mathrm{HK}}_{\mathbf{Q}}],\kappa^{\ast}[\Psi^{\mathrm{HK}}_{\mathbf{Q}}];\mathbf{Q}\big{]}\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\min_{(\varrho,\kappa,\kappa^{\ast})\to\mathbf{Q}}E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{Q}]\Big{)}\,.\end{split} (31)

As the KS functional in Eq. (23), the Kohn-Sham–Bogolyubov-de Gennes (KSBdG) functional is obtained by regarding qA=QAq^{A}=Q^{A} as a constraint,

EKSBdG[ϱ,κ,κ;𝐪]=EirrKSBdG[ϱ,κ,κ;𝐪]+EregKSBdG[𝐪]AΛAKSBdG(qAQA[ϱ,κ,κ]).E^{\mathrm{KSBdG}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]=E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]+E^{\mathrm{KSBdG}}_{\mathrm{reg}}[\mathbf{q}]-\sum_{A}\Lambda^{\mathrm{KSBdG}}_{A}\,\Big{(}q^{A}-Q^{A}[\varrho,\kappa,\kappa^{\ast}]\Big{)}\,. (32)

In addition, the particle-number condition (8) is considered,

E~KSBdG:=EKSBdGμ[tr(ϱ)N].\tilde{E}^{\mathrm{KSBdG}}:=E^{\mathrm{KSBdG}}-\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,. (33)

By defining

𝖧KSBdG:=(hKSBdGμΔKSBdGΔKSBdG(hKSBdG)+μ),\mathsf{H}^{\mathrm{KSBdG}}:=\begin{pmatrix}h^{\mathrm{KSBdG}}-\mu&\Delta^{\mathrm{KSBdG}}\\ -\Delta^{\ast\mathrm{KSBdG}}&-(h^{\mathrm{KSBdG}})^{\ast}+\mu\end{pmatrix}\,, (34)

where

hkKSBdG:=EKSBdGϱk=EirrKSBdG[ϱ,κ,κ;𝐪]ϱk+ΓkKSBdG;ΓkKSBdG:=AΛAKSBdGQAϱk=A(EregKSBdG[𝐪]qA+EirrKSBdG[ϱ,κ,κ;𝐪]qA)QAϱk,ΔkKSBdG:=EKSBdGκk=EirrKSBdG[ϱ,κ,κ;𝐪]κkA(EregKSBdG[𝐪]qA+EirrKSBdG[ϱ,κ,κ;𝐪]qA)QAκk,ΔkKSBdG:=EKSBdGκk=(ΔkKSBdG),\begin{split}h^{\mathrm{KSBdG}}_{k\ell}&:=\frac{\partial E^{\mathrm{KSBdG}}}{\partial\varrho_{\ell k}}=\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]}{\partial\varrho_{\ell k}}+\Gamma^{\mathrm{KSBdG}}_{k\ell}\,;\\ &\Gamma^{\mathrm{KSBdG}}_{k\ell}:=\sum_{A}\Lambda^{\mathrm{KSBdG}}_{A}\,\frac{\partial Q^{A}}{\partial\varrho_{k\ell}}=\sum_{A}\bigg{(}\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]}{\partial q^{A}}\bigg{)}\,\frac{\partial Q^{A}}{\partial\varrho_{k\ell}}\,,\\ \Delta^{\mathrm{KSBdG}}_{k\ell}&:=-\frac{\partial E^{\mathrm{KSBdG}}}{\partial\kappa^{\ast}_{\ell k}}\\ &=-\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]}{\partial\kappa^{\ast}_{\ell k}}-\sum_{A}\bigg{(}\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]}{\partial q^{A}}\bigg{)}\,\frac{\partial Q^{A}}{\partial\kappa^{\ast}_{k\ell}}\,,\\ \Delta^{\ast\mathrm{KSBdG}}_{k\ell}&:=-\frac{\partial E^{\mathrm{KSBdG}}}{\partial\kappa_{\ell k}}=\big{(}\Delta^{\mathrm{KSBdG}}_{k\ell}\big{)}^{\ast}\,,\end{split} (35)

the KSBdG equation is derived,

𝖧kKSBdG𝖶iKSBdG=εiKSBdG𝖶kiKSBdG,\sum_{\ell}\mathsf{H}^{\mathrm{KSBdG}}_{k\ell}\,\mathsf{W}^{\mathrm{KSBdG}}_{\ell i}=\varepsilon^{\mathrm{KSBdG}}_{i}\,\mathsf{W}^{\mathrm{KSBdG}}_{ki}\,, (36)

in analogy to the HFB equation (63). Above, the relation obtained by the variation with respect to qAq^{A},

ΛAKSBdG=EirrKSBdG[ϱ,κ,κ;𝐪]qA+EregKSBdG[𝐪]qA.\Lambda^{\mathrm{KSBdG}}_{A}=\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{irr}}[\varrho,\kappa,\kappa^{\ast};\mathbf{q}]}{\partial q^{A}}+\frac{\partial E^{\mathrm{KSBdG}}_{\mathrm{reg}}[\mathbf{q}]}{\partial q^{A}}\,. (37)

has been inserted. The identical mathematical structure guarantees that the arguments on the solutions of the HFB equation in Appendix A also apply to those of the KSBdG equation (36).

It is noted that κ\kappa and κ\kappa^{\ast} do not need to be physical in the KSBdG scheme. The KSBdG solution can be exact for systems where the particle number is fixed, provided that EKSBdGE^{\mathrm{KSBdG}} is appropriately determined.

V Discussions on principal variables and energy functional

The reformulation of the KS theory elucidates several aspects of the theory. Some of them are new, not recognized sufficiently, while others have been known but can now be argued in an organized manner. We hereafter focus on the KS equation for the sake of simplicity. Extension to the KSBdG equation will be straightforward.

V.1 Principal variables and density matrix

Whereas the HK theorem in Sec. III does not demand it, all QAQ^{A}’s have been assumed to depend solely on ϱ\varrho in Sec. IV.1, not on 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2). Comments on this point are given.

The KS equation closes within the s.p. degrees of freedom, not leading to hierarchical equations. This remarkable property is actualized because everything in EKSE^{\mathrm{KS}} does not depend on the correlation functions 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2). The principal variables 𝐐\mathbf{Q} should not depend on 𝒞(ν)\mathcal{C}^{(\nu)} for the KS equation to keep mathematical rigor.

One might consider that the KS equation can be derived even when QAQ^{A} depends on 𝒞(ν)\mathcal{C}^{(\nu)} via a constrained search similar to Eq. (16Donnely and Parr (1978). For instance, suppose that QAQ^{A} depends on 𝒞(2)\mathcal{C}^{(2)} as well as on ϱ\varrho. Then EKSE^{\mathrm{KS}} of Eq. (23) also depends on 𝒞(2)\mathcal{C}^{(2)}. Not to lead to hierarchical equations, a new EDF should be constructed via the constrained search min(ϱ,𝒞(2))ϱEKS\min_{(\varrho,\mathcal{C}^{(2)})\to\varrho}E^{\mathrm{KS}}. However, this introduces another step in the minimization process, which likely damages the differentiability of QAQ^{A} for ϱ\varrho, as illustrated in Appendix B. Because the KS theory relies on this differentiability, it is desired that the principal variables do not depend on 𝒞(ν)\mathcal{C}^{(\nu)}. Conversely, if 𝒞(ν)\mathcal{C}^{(\nu)} is initially contained in QAQ^{A} but is finally eliminated in EKSE^{\mathrm{KS}}, one should pay special attention to the differentiability.

V.2 Equational equivalence

It is worth recalling that the KS equation (27) is derived primarily from the variation of EKS[ϱ;𝐪]E^{\mathrm{KS}}[\varrho;\mathbf{q}] with respect to ϱ\varrho. Apart from the physical importance, the variation with respect to qAq^{A} is subsidiary in deriving the equational form, only determining the KS potential ΓKS\Gamma^{\mathrm{KS}} through the Lagrange multiplier ΛAKS\Lambda^{\mathrm{KS}}_{A} as in Eqs. (25) and (26). Indeed, variation of EKSE^{\mathrm{KS}} for ϱ\varrho after inserting qA=QA[ϱ]q^{A}=Q^{A}[\varrho]leads to the same s.p. Hamiltonian as in Eq. (26),

hkKS=ϱkEKS[ϱ;𝐐[ϱ]]=ϱk(EirrKS[ϱ;𝐐[ϱ]]+EregKS[𝐐[ϱ]]).h^{\mathrm{KS}}_{k\ell}=\frac{\partial}{\partial\varrho_{k\ell}}E^{\mathrm{KS}}\big{[}\varrho;\mathbf{Q}[\varrho]\big{]}=\frac{\partial}{\partial\varrho_{k\ell}}\Big{(}E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho;\mathbf{Q}[\varrho]\big{]}+E^{\mathrm{KS}}_{\mathrm{reg}}\big{[}\mathbf{Q}[\varrho]\big{]}\Big{)}\,. (38)

The g.s. energy is given at ϱ=ϱKS\varrho=\varrho^{\mathrm{KS}},

E0=EKS[ϱKS;𝐐[ϱKS]]=EirrKS[ϱKS;𝐐[ϱKS]]+EregKS[𝐐[ϱKS]].E_{0}=E^{\mathrm{KS}}\big{[}\varrho^{\mathrm{KS}};\mathbf{Q}[\varrho^{\mathrm{KS}}]\big{]}=E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho^{\mathrm{KS}};\mathbf{Q}[\varrho^{\mathrm{KS}}]\big{]}+E^{\mathrm{KS}}_{\mathrm{reg}}\big{[}\mathbf{Q}[\varrho^{\mathrm{KS}}]\big{]}\,. (39)

This feature implies that equations equivalent to (27) are obtained by removing or adding some variables to 𝐐\mathbf{Q}, as long as the total EKSE^{\mathrm{KS}} is an identical functional of ϱ\varrho. Consider the case that some of the principal variables are taken out. The new set is denoted by 𝐐={QB}\mathbf{Q}^{\prime}=\{Q^{B^{\prime}}\} and the removed set by 𝐐¯={QB¯}\bar{\mathbf{Q}}=\{Q^{\bar{B}}\}; 𝐐=𝐐𝐐¯\mathbf{Q}=\mathbf{Q}^{\prime}\oplus\bar{\mathbf{Q}}. EregKSE^{\mathrm{KS}}_{\mathrm{reg}} is separated into two parts, EregKS=E¯regKS+EregKSE^{\mathrm{KS}}_{\mathrm{reg}}=\bar{E}^{\mathrm{KS}}_{\mathrm{reg}}+E^{\prime\,\mathrm{KS}}_{\mathrm{reg}}, where E¯regKS\bar{E}^{\mathrm{KS}}_{\mathrm{reg}} includes QB¯Q^{\bar{B}} and EregKSE^{\prime\,\mathrm{KS}}_{\mathrm{reg}} does not. By newly defining EirrKS:=EirrKS+E¯regKSE^{\prime\,\mathrm{KS}}_{\mathrm{irr}}:=E^{\mathrm{KS}}_{\mathrm{irr}}+\bar{E}^{\mathrm{KS}}_{\mathrm{reg}}, the following equality,

EirrKS[ϱ;𝐐[ϱ]]+EregKS[𝐐[ϱ]]=EirrKS[ϱ;𝐐[ϱ]]+EregKS[𝐐[ϱ]],E^{\prime\,\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho;\mathbf{Q}^{\prime}[\varrho]\big{]}+E^{\prime\,\mathrm{KS}}_{\mathrm{reg}}\big{[}\mathbf{Q}^{\prime}[\varrho]\big{]}=E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho;\mathbf{Q}[\varrho]\big{]}+E^{\mathrm{KS}}_{\mathrm{reg}}\big{[}\mathbf{Q}[\varrho]\big{]}\,, (40)

is fulfilled up to the functional form of ϱ\varrho. This relation guarantees the equivalence at the level of the equations as well as the resultant energy. The opposite is true: principal variables may be added without influencing the s.p. equations. We shall call this feature equational equivalence.

Despite the equational equivalence, the choice of 𝐐\mathbf{Q} is physically significant. As already discussed, while the energy EE and the principal variables 𝐐\mathbf{Q} obtained from the KS equation are physical, other quantities do not have strict physical meaning. The equational equivalence indicates that the equational form does not tell what variables are physical. It entirely depends on how the EDF has been constructed. This point matters when the EDF is phenomenologically fixed as in Sec. VI.2.

Several limiting cases are instructive. First, if all QAQ^{A}’s are taken out by setting 𝐐¯=𝐐\bar{\mathbf{Q}}=\mathbf{Q}, 𝐐\mathbf{Q}^{\prime} becomes an empty set \emptyset, leading to EregKS=0E^{\prime\,\mathrm{KS}}_{\mathrm{reg}}=0 and EKS[ϱ;]=EirrKS[ϱ;]E^{\mathrm{KS}}[\varrho;\emptyset]=E^{\prime\,\mathrm{KS}}_{\mathrm{irr}}[\varrho;\emptyset]. In this case, no quantity other than the total energy is considered physical. Second, it is possible to set EirrKS=0E^{\prime\,\mathrm{KS}}_{\mathrm{irr}}=0, opposite to the first limiting case. There are practical cases in which EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} can be represented in terms of certain currents. For instance, in the standard KS theory for electronic systems, the kinetic density of the non-interacting system is expressed as ξ(𝐫)/2m\xi(\mathbf{r})/2m, where

ξ(𝐫):=i=1N[φi(𝐫)][φi(𝐫)].\xi(\mathbf{r}):=\sum_{i=1}^{N}[\nabla\varphi_{i}^{\ast}(\mathbf{r})]\cdot[\nabla\varphi_{i}(\mathbf{r})]\,. (41)

If all those currents are added to 𝐐\mathbf{Q}, we have EirrKS=0E^{\prime\,\mathrm{KS}}_{\mathrm{irr}}=0. Still, we should consider EirrKSE^{\prime\,\mathrm{KS}}_{\mathrm{irr}} to depend on ϱ\varrho, since it is essential to derive the KS equation. An extreme case along this line is that of 𝐐={ϱk}\mathbf{Q}=\{\varrho_{k\ell}\}, the whole 1-body density matrix Donnely and Parr (1978); Zumbach and Maschke (1985); Requistl and Pankratov (2008); Pernal and Giesbertz (2016). Then all components of ϱ\varrho would be considered physical. However, an idempotent ϱ\varrho, which necessarily results from the KS equation, cannot be exact in general, even though there remains room to interpret the s.p. states as those of dressed particles. These limiting cases help to understand issues of representability, as discussed in the subsequent subsection.

V.3 Representability in the KS theory

Significance of representability has been perceived for the HK theorem. The HK theorem in Sec. III was initially asserted by showing the uniqueness of the mapping from the electron density n(𝐫)n(\mathbf{r}) to the external electron-ion potential Uei(𝐫)U_{ei}(\mathbf{r}) Hohenberg and Kohn (1964), which are connected via the Legendre transformation. This mapping should exist for any n(𝐫)n(\mathbf{r}) with n(𝐫)0n(\mathbf{r})\geq 0 and d3rn(𝐫)=N\int d^{3}r\,n(\mathbf{r})=N (the vv-representability). While this vv-representability issue can be circumvented by the constrained search of Eq. (16), it is transferred to the differentiability of the EDF as mentioned in Sec. III, and this problem could arise even at the KS level. Once 𝚲KS={ΛAKS}\mathbf{\Lambda}^{\mathrm{KS}}=\{\Lambda^{\mathrm{KS}}_{A}\} are fixed, the KS equation (27) with the KS Hamiltonian (26) determines ϱKS\varrho^{\mathrm{KS}} and thereby the principal variables 𝐐[ϱKS]\mathbf{Q}[\varrho^{\mathrm{KS}}], defining the mapping from 𝚲KS\mathbf{\Lambda}^{\mathrm{KS}} to 𝐐\mathbf{Q}. The question is in the inverse mapping from 𝐐\mathbf{Q} to 𝚲KS\mathbf{\Lambda}^{\mathrm{KS}}, i.e., whether 𝚲KS\mathbf{\Lambda}^{\mathrm{KS}} exists for any 𝐐\mathbf{Q} in its physical domain. The existence of this inverse mapping may be regarded as vv-representability at the KS level, which has been called ‘non-interacting vv-representability’ in some literature. Equation (25) tells us that the vv-representability at the KS level arrives at the differentiability of EirrKS[ϱ;𝐐]E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}] and EregKS[𝐐]E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{Q}] with respect to QAQ^{A}.

As already mentioned, any 𝐐\mathbf{Q} in the physical domain should be represented by the 1-body density matrix ϱ\varrho that satisfies Eqs. (8) and (11). This condition is called NN-representability and has been proven for the original case 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}) Harriman (1981). For the KS theory to be applicable, any 𝐐\mathbf{Q} in its physical domain should be obtained by an idempotent ϱ\varrho. This condition is stronger than the NN-representability for the HK theorem and the ‘pure-state representability’ in Ref. Valone (1980). We shall call it NN-representability at the KS level. For 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}), the NN-representability is also satisfied at the KS level Harriman (1981).

Two representabilities have been distinguished in the fermionic DFT, the vv- and the NN-representabilities. Both are further distinguished between those at the HK and KS levels. The vv-representability is linked to the differentiability of the EDF, and it is usually weaker at the KS level than at the HK level. On the other hand, the NN-representability at the KS level, which requires the idempotency of the 1-body density matrix, is stronger than that at the HK level.

V.4 Criteria for extendability

The reformulation of the KS theory with the least postulates addresses the criteria for its extendability. The following conditions should be satisfied for the KS equation to be rigorous and practical, and they give the criteria.

  • (a)

    Principal variables 𝐐\mathbf{Q} should be appropriately chosen. They should help to minimize energy and contain all quantities under physical interest other than energy. At the same time, all QAQ^{A}’s should depend only on ϱ\varrho and not on the correlation functions 𝒞(ν)\mathcal{C}^{(\nu)} (ν2\nu\geq 2), or well approximated by functionals only of ϱ\varrho. Moreover, 𝐐\mathbf{Q} should be NN-representable at the KS level in the domain under search, i.e., representable by an idempotent ϱ\varrho with tr(ϱ)=N\mathrm{tr}(\varrho)=N.

  • (b)

    Irregularity in EHK[𝐐]E^{\mathrm{HK}}[\mathbf{Q}] can be removed via EirrKS[ϱ;𝐐]E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}], which is a differentiable functional of ϱ\varrho. Furthermore, EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} should properly be chosen so that EregKS[𝐐]E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{Q}] should be regular for 𝐐\mathbf{Q}. Because EregKSE^{\mathrm{KS}}_{\mathrm{reg}} is defined by (22), this condition usually requires that both EregHK[𝐐]E^{\mathrm{HK}}_{\mathrm{reg}}[\mathbf{Q}] and (EirrKS[ϱ[Ψ𝐐HK];𝐐]minϱ𝐐EirrKS[ϱ;𝐐])\Big{(}E^{\mathrm{KS}}_{\mathrm{irr}}\big{[}\varrho[\Psi^{\mathrm{HK}}_{\mathbf{Q}}];\mathbf{Q}\big{]}-\min_{\varrho\to\mathbf{Q}}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;\mathbf{Q}]\Big{)} are regular for 𝐐\mathbf{Q}.

  • (c)

    The entire EKS[ϱ;𝐐]E^{\mathrm{KS}}[\varrho;\mathbf{Q}] should be fixed before starting computation, which requires that EregKS[𝐐]E^{\mathrm{KS}}_{\mathrm{reg}}[\mathbf{Q}] is constructed appropriately.

Condition (a) is the first step for applying the KS approach. One might think that the greater number of principal variables provide the more accurate description of the system. However, the NN-representability becomes the more difficult to be satisfied. Therefore the set 𝐐\mathbf{Q} should be neither too big nor too small. Significance of (b) is exemplified by the normal-to-superconductor transition. The pair correlation carried by κ\kappa and κ\kappa^{\ast} plays an essential role in this phenomenon, connected to the order parameter. It is difficult to remove the irregularity at the transition only by ϱ\varrho, indicating that the KS theory itself cannot be applied, though possible via the extension to the KSBdG equation777 For the KSBdG scheme, ϱ\varrho in Conditions (a–c) should be read as (ϱ,κ,κ)(\varrho,\kappa,\kappa^{\ast}). . Another example could be the confinement of constituent quarks, in which the 3-body correlation forming the color singlet is essential. It does not seem possible to ascribe irregularity in EHKE^{\mathrm{HK}} to EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} that is a functional only of ϱ\varrho, without using 𝒞(3)\mathcal{C}^{(3)}.

If the underlying Hamiltonian HH is known and is partitioned as in Eq. (20), it may be possible to consider HirrH_{\mathrm{irr}} and HregH_{\mathrm{reg}} unperturbed Hamiltonian and perturbation, respectively. Then, when setting on HregH_{\mathrm{reg}} adiabatically, the wave-function |Ψ|\Psi\rangle changes gradually, as the adiabatic theorem Messiah (1961) tells us. At a glance, it looks reasonable to expect that EregKS[𝐐]E^{\mathrm{KS}}_{\mathrm{reg}}\big{[}\mathbf{Q}] is regular in this case. However, this is not always true because it breaks down for the 𝐐={ϱk}\mathbf{Q}=\{\varrho_{k\ell}\} case discussed in Sec. V.2, in which the NN-representability at the KS level is lost.

Unless all conditions (a–c) are fulfilled, approaches using the KS equation lose firmness or practicality. Conversely, as long as an EDF meets these conditions, the many-body system can be solved via s.p. equations, following the procedure in Sec. IV. Extensions may be adding principal variables, modifying the EDF, or applying to other many-fermion systems than those composed of electrons.

VI Practice of Kohn-Sham theory

We next show how the above arguments on the KS theory are related to practical cases.

VI.1 Many-electron systems: From atoms and molecules to solids

The KS theory has been successful for electronic systems, from atoms and molecules to solids. Under 𝐐=n(𝐫)\mathbf{Q}=n(\mathbf{r}), it has been established as a standard ab initio method for many-electron systems. The Hamiltonian of the system is comprised of the kinetic energy KK, the external potential provided by ions UeiU_{ei}, and the electron-electron interaction VeeV_{ee}. In the context of Eq. (20), these terms are partitioned as888 Although the contribution of UeiU_{ei} is regular for n(𝐫)n(\mathbf{r}), we put it into HirrH_{\mathrm{irr}} so that HregH_{\mathrm{reg}} could be universal. Shifting it to HregH_{\mathrm{reg}} does not influence the resultant equations.,

Hirr=K+Uei,Hreg=Vee.H_{\mathrm{irr}}=K+U_{ei}\,,\quad H_{\mathrm{reg}}=V_{ee}\,. (42)

Correspondingly, the EDF follows Eq. (20). As in Eq. (28), EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} is separated into individual terms of Eq. (42) as

EirrKS[ϱ;n(𝐫)]=EK[ϱ]+Eei[n(𝐫)];EK[ϱ]=12mk|2|kϱk,Eei[n(𝐫)]=d3rUei(𝐫)n(𝐫),\begin{split}E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;n(\mathbf{r})]&=E_{K}[\varrho]+E_{ei}[n(\mathbf{r})]\,;\\ &E_{K}[\varrho]=-\frac{1}{2m}\sum_{k\ell}\langle\ell|\nabla^{2}|k\rangle\,\varrho_{k\ell}\,,\quad E_{ei}[n(\mathbf{r})]=\int d^{3}r^{\prime}\,U_{ei}(\mathbf{r^{\prime}})\,n(\mathbf{r^{\prime}})\,,\end{split} (43)

where EK[ϱ]E_{K}[\varrho] corresponds to the first term of Eq. (28) and Eei[n(𝐫)]=ΔEirrKSE_{ei}[n(\mathbf{r})]=\mathit{\Delta}E^{\mathrm{KS}}_{\mathrm{irr}}. Then EregKSE^{\mathrm{KS}}_{\mathrm{reg}} in Eq. (22) becomes

EregKS[n(𝐫)]=EregHK[n(𝐫)]+(EK[ϱ[Ψn(𝐫)HK]]minϱn(𝐫)EK[ϱ]),E^{\mathrm{KS}}_{\mathrm{reg}}[n(\mathbf{r})]=E^{\mathrm{HK}}_{\mathrm{reg}}[n(\mathbf{r})]+\Big{(}E_{K}\big{[}\varrho[\Psi^{\mathrm{HK}}_{n(\mathbf{r})}]\big{]}-\min_{\varrho\to n(\mathbf{r})}E_{K}[\varrho]\Big{)}\,, (44)

having no contribution from EeiE_{ei}. Since UeiU_{ei} in Eq. (42) carries characteristics of individual systems, EregKSE^{\mathrm{KS}}_{\mathrm{reg}} is expected to be universal. As long as HregH_{\mathrm{reg}} or its exchange-correlation part shown below is perturbative, the regularity of EregKSE^{\mathrm{KS}}_{\mathrm{reg}} is likely satisfied. There have been attempts to extend the theory to include other variables in 𝐐\mathbf{Q} as mentioned in Sec. I, although the representability issues have not been explored sufficiently.

EregKSE^{\mathrm{KS}}_{\mathrm{reg}} is taken as a sum of the direct term of VeeV_{ee} and the rest:

EregKS[n(𝐫)]=Edir[n(𝐫)]+ExcKS[n(𝐫)];Edir[n(𝐫)]=12d3rd3r′′n(𝐫)n(𝐫′′)|𝐫𝐫′′|,ExcKS[n(𝐫)]=d3rxc[n(𝐫)].\begin{split}E^{\mathrm{KS}}_{\mathrm{reg}}[n(\mathbf{r})]&=E_{\mathrm{dir}}[n(\mathbf{r})]+E^{\mathrm{KS}}_{\mathrm{xc}}[n(\mathbf{r})]\,;\\ &E_{\mathrm{dir}}[n(\mathbf{r})]=\frac{1}{2}\int d^{3}r^{\prime}\,d^{3}r^{\prime\prime}\,\frac{n(\mathbf{r}^{\prime})\,n(\mathbf{r}^{\prime\prime})}{|\mathbf{r}^{\prime}-\mathbf{r}^{\prime\prime}|}\,,\quad E^{\mathrm{KS}}_{\mathrm{xc}}[n(\mathbf{r})]=\int d^{3}r^{\prime}\,\mathcal{H}_{\mathrm{xc}}[n(\mathbf{r}^{\prime})]\,.\end{split} (45)

All of the complicated many-body effects are attributed to the exchange-correlation term ExcKSE^{\mathrm{KS}}_{\mathrm{xc}}. It should be noted that the general theory does not demand EregKSE^{\mathrm{KS}}_{\mathrm{reg}} to be represented by an integral of a local function. Indeed, EdirE_{\mathrm{dir}} is an integral containing the densities at distant positions 𝐫\mathbf{r}^{\prime} and 𝐫′′\mathbf{r}^{\prime\prime}, multiplied by the interaction having the non-local form 1/|𝐫𝐫′′|1/|\mathbf{r}^{\prime}-\mathbf{r}^{\prime\prime}|. In contrast, ExcKSE^{\mathrm{KS}}_{\mathrm{xc}} is customarily assumed to be the integral of a local function xc\mathcal{H}_{\mathrm{xc}}, a function depending on the density at a single position 𝐫\mathbf{r}^{\prime}. The kernel xc\mathcal{H}_{\mathrm{xc}} has been determined by the local-density approximation (LDA) Hohenberg and Kohn (1964); Sahni et al. (1988) or the generalized gradient approximation (GGA) Perdew et al. (1996). A method for obtaining xc\mathcal{H}_{\mathrm{xc}} called ‘meta-GGA’ has been developed Tao et al. (2003). A machine-learning-based method has been proposed recently Nagai et al. (2022). While the locality assumption on ExcKSE^{\mathrm{KS}}_{\mathrm{xc}} has helped the KS theory to be practical in electronic systems, it is an approximation whose appropriateness should be scrutinized.

It deserves discussing the spin DFT in the absence of external magnetic field, which may exemplify arguments in Sec. V. In the spin DFT, it is usually considered that the spin density σ(𝐫):=n(𝐫)n(𝐫)\sigma(\mathbf{r}):=n_{\uparrow}(\mathbf{r})-n_{\downarrow}(\mathbf{r}) is the principal variable additional to the density n(𝐫)=n(𝐫)+n(𝐫)n(\mathbf{r})=n_{\uparrow}(\mathbf{r})+n_{\downarrow}(\mathbf{r}). The spin density σ(𝐫)\sigma(\mathbf{r}) appears in the exchange-correlation term in the KS EDF as

EKS[ϱ;n(𝐫),σ(𝐫)]=EK[ϱ]+Eei[n(𝐫)]+Edir[n(𝐫)]+ExcKS[n(𝐫),σ(𝐫)].E^{\mathrm{KS}}[\varrho;n(\mathbf{r}),\sigma(\mathbf{r})]=E_{K}[\varrho]+E_{ei}[n(\mathbf{r})]+E_{\mathrm{dir}}[n(\mathbf{r})]+E^{\mathrm{KS}}_{\mathrm{xc}}[n(\mathbf{r}),\sigma(\mathbf{r})]\,. (46)

It should be kept in mind that the NN-representability has not been proven for the case 𝐐={n(𝐫),σ(𝐫)}\mathbf{Q}=\{n(\mathbf{r}),\sigma(\mathbf{r})\}.

The system can spontaneously be magnetized. Let us go back to the EDF at the HK level. As far as σ(𝐫)/n(𝐫)\sigma(\mathbf{r})/n(\mathbf{r}) is small, we can express the HK EDF as

EHK[n(𝐫),σ(𝐫)]=C0HK[n(𝐫)]+C2HK[n(𝐫)][σ(𝐫)]2+C4HK[n(𝐫)][σ(𝐫)]4.E^{\mathrm{HK}}[n(\mathbf{r}),\sigma(\mathbf{r})]=C_{0}^{\mathrm{HK}}[n(\mathbf{r})]+C_{2}^{\mathrm{HK}}[n(\mathbf{r})]\cdot[\sigma(\mathbf{r})]^{2}+C_{4}^{\mathrm{HK}}[n(\mathbf{r})]\cdot[\sigma(\mathbf{r})]^{4}\,. (47)

We assume C4HK[n]>0C_{4}^{\mathrm{HK}}[n]>0 everywhere, and that C2HK[n]C_{2}^{\mathrm{HK}}[n] changes its sign at a certain value of nn, denoted by ncrn_{\mathrm{cr}}:

{C2HK[n]>0for n<ncrC2HK[n]<0for n>ncr.\left\{\begin{array}[]{ll}C_{2}^{\mathrm{HK}}[n]>0&\mbox{for $n<n_{\mathrm{cr}}$}\\ C_{2}^{\mathrm{HK}}[n]<0&\mbox{for $n>n_{\mathrm{cr}}$}\end{array}\right.\,. (48)

Regarding Eq. (46), the C2C_{2} and C4C_{4} terms mainly correspond to ExcKSE^{\mathrm{KS}}_{\mathrm{xc}}, although EKE_{K} has some contribution due to the Pauli principle, which should be positive for C2C_{2}. Minimizing EHK[n(𝐫),σ(𝐫)]E^{\mathrm{HK}}[n(\mathbf{r}),\sigma(\mathbf{r})] with respect to σ\sigma, we find a non-magnetized phase and a magnetized phase,

minσ(𝐫)EHK[n(𝐫),σ(𝐫)]={C0HK[n(𝐫)]with σ(𝐫)=0(for n(𝐫)<ncr)C0HK[n(𝐫)]|C2HK[n(𝐫)]|24C4HK[n(𝐫)]with [σ(𝐫)]2=|C2HK[n(𝐫)]|2C4HK[n(𝐫)](for n(𝐫)>ncr).\begin{split}&\min_{\sigma(\mathbf{r})}E^{\mathrm{HK}}[n(\mathbf{r}),\sigma(\mathbf{r})]\\ &\quad=\left\{\begin{array}[]{lll}C_{0}^{\mathrm{HK}}[n(\mathbf{r})]&\mbox{with }\sigma(\mathbf{r})=0&\mbox{(for $n(\mathbf{r})<n_{\mathrm{cr}}$)}\\ {\displaystyle C_{0}^{\mathrm{HK}}[n(\mathbf{r})]-\frac{\big{|}C_{2}^{\mathrm{HK}}[n(\mathbf{r})]\big{|}^{2}}{4\,C_{4}^{\mathrm{HK}}[n(\mathbf{r})]}}&\mbox{with }{\displaystyle[\sigma(\mathbf{r})]^{2}=\frac{\big{|}C_{2}^{\mathrm{HK}}[n(\mathbf{r})]\big{|}}{2\,C_{4}^{\mathrm{HK}}[n(\mathbf{r})]}}&\mbox{(for $n(\mathbf{r})>n_{\mathrm{cr}}$)}\end{array}\right.\,.\end{split} (49)

This spontaneous magnetization across n=ncrn=n_{\mathrm{cr}} can be handled in the KS EDF of Eq. (46).

Since the principal variables 𝐐={n(𝐫),σ(𝐫)}\mathbf{Q}=\{n(\mathbf{r}),\sigma(\mathbf{r})\} covers those of the ordinary KS DFT, 𝐐={n(𝐫)}\mathbf{Q}^{\prime}=\{n(\mathbf{r})\}, the EDF of the ordinary KS theory should be derived from the spin DFT, in principle. However, an irregularity enters owing to the transition of the phases across n=ncrn=n_{\mathrm{cr}} in Eq. (49), which is distinguished from the irregularity due to the Pauli principle of Eq. (11). In practice, the HK EDF for 𝐐={n(𝐫)}\mathbf{Q}^{\prime}=\{n(\mathbf{r})\} should be

EHK[n(𝐫)]=minσ(𝐫)EHK[n(𝐫),σ(𝐫)]=C0HK[n(𝐫)]δ(n(𝐫)ncr)|C2HK[n(𝐫)]|24C4HK[n(𝐫)].E^{\mathrm{HK}}[n(\mathbf{r})]=\min_{\sigma(\mathbf{r})}E^{\mathrm{HK}}[n(\mathbf{r}),\sigma(\mathbf{r})]=C_{0}^{\mathrm{HK}}[n(\mathbf{r})]-\delta\big{(}n(\mathbf{r})-n_{\mathrm{cr}}\big{)}\cdot\frac{\big{|}C_{2}^{\mathrm{HK}}[n(\mathbf{r})]\big{|}^{2}}{4\,C_{4}^{\mathrm{HK}}[n(\mathbf{r})]}\,. (50)

The second term of the right-hand side (rhs) implies irregularity with respect to n(𝐫)n(\mathbf{r}). It is difficult to remove the irregularity via EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} of Eq. (43). Instead, we may construct a KS EDF by taking EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} as

EirrKS[ϱ;n(𝐫)]=EK[ϱ]+Eei[n(𝐫)]+C2KS[n(𝐫)][σ(𝐫)]2+C4KS[n(𝐫)][σ(𝐫)]4,E^{\mathrm{KS}}_{\mathrm{irr}}[\varrho;n(\mathbf{r})]=E_{K}[\varrho]+E_{ei}[n(\mathbf{r})]+C_{2}^{\mathrm{KS}}[n(\mathbf{r})]\cdot[\sigma(\mathbf{r})]^{2}+C_{4}^{\mathrm{KS}}[n(\mathbf{r})]\cdot[\sigma(\mathbf{r})]^{4}\,, (51)

where CνKSC_{\nu}^{\mathrm{KS}} (ν=0,2,4\nu=0,2,4) is the σν\sigma^{\nu} term in ExcKS[n,σ]E^{\mathrm{KS}}_{\mathrm{xc}}[n,\sigma] of Eq. (46). The regular part becomes

EregKS[n(𝐫)]=Edir[n(𝐫)]+C0KS[n(𝐫)].E^{\mathrm{KS}}_{\mathrm{reg}}[n(\mathbf{r})]=E_{\mathrm{dir}}[n(\mathbf{r})]+C_{0}^{\mathrm{KS}}[n(\mathbf{r})]\,. (52)

While this EDF well takes account of the irregularity due to the magnetization, σ(𝐫)\sigma(\mathbf{r}) is not regarded as a principal variable. This reminds us of the problem; NN-representability has not been proven for σ(𝐫)\sigma(\mathbf{r}). We may reinterpret the spin DFT, abandoning to keep the whole σ(𝐫)\sigma(\mathbf{r}) as principal variables but regarding the magnetization :=d3rσ(𝐫)\mathcal{M}:=\int d^{3}r\,\sigma(\mathbf{r}) as a principal variable. The KS EDF of Eq. (46) is identified as EKS[ϱ;n(𝐫),]E^{\mathrm{KS}}[\varrho;n(\mathbf{r}),\mathcal{M}], without impairing the NN-representability999 The argument in Ref. Harriman (1981) applies by adopting the s.p. states 1c2ϕ(𝐫)+cϕ(𝐫)\sqrt{1-c^{2}}\,\phi_{\uparrow}(\mathbf{r})+c\,\phi_{\downarrow}(\mathbf{r}) with 2|c|2=1/N2|c|^{2}=1-\mathcal{M}/N. .

VI.2 Many-nucleon systems: Atomic nuclei

The success of the KS theory in many-electron systems has made us wonder whether an ab initio method for nuclear structure calculations could be developed analogously Drut et al. (2010). However, problems remain to be resolved to establish the KS theory in nuclei.

A nucleus consists of two species of particles with nearly equal masses, protons (pp) and neutrons (nn). The particle number of each particle type, NτN_{\tau} (τ=p,n)(\tau=p,n), is conserved. Equation (24) is modified accordingly,

E~KS:=EKSτ=p,nμτ[trτ(ϱ)Nτ],\tilde{E}^{\mathrm{KS}}:=E^{\mathrm{KS}}-\sum_{\tau=p,n}\mu_{\tau}\big{[}\mathrm{tr}_{\tau}(\varrho)-N_{\tau}\big{]}\,, (53)

where trτ\mathrm{tr}_{\tau} is the trace over the s.p. space belonging to the particle type τ\tau. When the local density is taken as a principal variable, it seems reasonable to use nτ(𝐫)n_{\tau}(\mathbf{r}), the density of each particle type.

In self-bound finite systems like atomic nuclei, treatment of the center-of-mass (c.m.) motion is a subtle problem. It seems desirable to construct a theory using the internal density Engel (2007), n~τ(𝐫):=Ψ|iτδ(𝐫(𝐫i𝐑))|Ψ\tilde{n}_{\tau}(\mathbf{r}):=\langle\Psi|\sum_{i\in\tau}\delta\big{(}\mathbf{r}-(\mathbf{r}_{i}-\mathbf{R})\big{)}|\Psi\rangle, where 𝐑\mathbf{R} is the c.m. coordinate. However, 𝐑\mathbf{R} in the δ\delta-function makes n~τ(𝐫)\tilde{n}_{\tau}(\mathbf{r}) a many-body quantity Nakada (2020), prohibiting the KS equation from being applied without losing exactness, relevantly to Condition (a) in Sec. V.4. See also the arguments in Sec. V.1. A rational approximation on n~τ(𝐫)\tilde{n}_{\tau}(\mathbf{r}) should be introduced to apply the KS theory. While discussions in this line have been found in Refs. Barnea (2007); Messud et al. (2009); Messud (2013), it has not fully been resolved, with no well-established practical method. Although this problem is critical in exploiting authentic DFT in nuclei, we proceed without touching on this problem further, leaving it for future works and simply assuming n~τ(𝐫)nτ(𝐫)=Ψ|iτδ(𝐫𝐫i)|Ψ\tilde{n}_{\tau}(\mathbf{r})\approx n_{\tau}(\mathbf{r})=\langle\Psi|\sum_{i\in\tau}\delta(\mathbf{r}-\mathbf{r}_{i})|\Psi\rangle, for the time being.

While approaches analogous to the KS theory have widely been applied to atomic nuclei, proper means to determine the EDFs have yet to be established. Although the density-matrix expansion was proposed from a microscopic standpoint Negele and Vautherin (1972); Campi and Sprung (1972), it has not been applied extensively. Another way is fitting parameters in an EDF or an effective interaction to the experimental data Kortelainen et al. (2012, 2014). We here give a typical form of the nuclear EDF, which is called Skyrme EDF:

ESkyrme=d3rSkyrme[nτ(𝐫),ξτ(𝐫),𝜻τ(𝐫)],E_{\mathrm{Skyrme}}=\int d^{3}r\,\mathcal{H}_{\mathrm{Skyrme}}[n_{\tau}(\mathbf{r}),\xi_{\tau}(\mathbf{r}),\bm{\zeta}_{\tau}(\mathbf{r})]\,, (54)

where

nτ(𝐫)=iτφi(𝐫)φi(𝐫),ξτ(𝐫)=iτφi(𝐫)φi(𝐫),𝜻τ(𝐫)=iiτφi(𝐫)𝝈×φi(𝐫),n_{\tau}(\mathbf{r})=\sum_{i\in\tau}\varphi_{i}^{\ast}(\mathbf{r})\,\varphi_{i}(\mathbf{r})\,,\qquad\xi_{\tau}(\mathbf{r})=\sum_{i\in\tau}\nabla\varphi_{i}^{\ast}(\mathbf{r})\cdot\nabla\varphi_{i}(\mathbf{r})\,,\quad\bm{\zeta}_{\tau}(\mathbf{r})=i\sum_{i\in\tau}\varphi_{i}^{\ast}(\mathbf{r})\,\bm{\sigma}\times\nabla\varphi_{i}(\mathbf{r})\,, (55)

and the index ii in the summation runs over the occupied s.p. states. The parameters contained in Skyrme\mathcal{H}_{\mathrm{Skyrme}} are usually determined by fitting to experimental data on some physical quantities. As pointed out in Sec. V.2, the functional form does not fix principal variables. EirrKSE^{\mathrm{KS}}_{\mathrm{irr}} and EregKSE^{\mathrm{KS}}_{\mathrm{reg}} are not well discriminated in ESkyrmeE_{\mathrm{Skyrme}}. For instance, we have a term with nτ(𝐫)ξτ(𝐫)n_{\tau}(\mathbf{r})\,\xi_{\tau}(\mathbf{r}) in Skyrme\mathcal{H}_{\mathrm{Skyrme}}, which emerges by expanding a finite-range interaction in terms of momentum Vautherin and Brink (1972); Negele and Vautherin (1972), not closely connected to the shell effects. However, if we classify this term into EregKSE^{\mathrm{KS}}_{\mathrm{reg}}, it is reasonable to consider ξτ(𝐫)\xi_{\tau}(\mathbf{r}) to be physical and belong to principal variables. Then the kinetic term becomes physical, which is unacceptable because the true nuclear wave-functions have high-momentum components beyond 𝒱idem\mathcal{V}_{\mathrm{idem}} Mahaux et al. (1985). It is crucial in this respect what quantities are employed for the fitting. Only the quantities in the fitting protocol can be considered principal variables, and physical meaning is questionable for other variables. It can be doubted that even the density nτ(𝐫)n_{\tau}(\mathbf{r}) is truly physical when not involved in the fitting protocol.

As microscopic approaches, the density-matrix expansion has been revived in connection to the nucleonic interaction derived from the chiral perturbation theory Kaiser et al. (2003); Navarro Pérez et al. (2018). Derivation of covariant EDFs based on the Brueckner theory has been advanced Shen et al. (2019). We also mention a few attempts to compromise microscopic and phenomenological ways to determine the EDF Nakada (2020); Baldo et al. (2008).

The total angular momentum JJ should be a good quantum number in nuclei because of the rotational symmetry. It is known that any even-even nucleus has J=0J=0 at its g.s. with no exception. Then the density should be rotationally symmetric, nτ(𝐫)=nτ(r)n_{\tau}(\mathbf{r})=n_{\tau}(r) where r=|𝐫|r=|\mathbf{r}|. However, a number of nuclei are considered to have deformed intrinsic states Bohr and Mottelson (1975). The g.s. of an even-even deformed nucleus has J=0J=0 as a superposition of the intrinsic states rotated with various angles. It is difficult to represent a g.s. of a deformed nucleus within 𝒱idem\mathcal{V}_{\mathrm{idem}}, viz., with a state comprised of the KS orbitals. In other words, the NN-representability at the KS level is questionable. It was argued in Ref. Giraud et al. (2008) that the KS solution would provide an intrinsic state rather than the observed g.s. Though practical, this argument seems to deviate from the variational principle of Eq. (2) and the HK theorem of Eq. (18). In this respect, the nuclear deformation could be an example that Condition (b) in Sec. V.4 does not rigorously hold. This is another problem that has not been resolved entirely. The angular-momentum projection Ring and Schuck (1980) has been developed to restore the rotational symmetry and applied to the solutions of Eq. (7) or (27). The projection has also been discussed in the context of the KS theory Giraud et al. (2008).

VII Summary

The Kohn-Sham (KS) theory provides an attractive and practical platform for ab initio calculations of many-fermion systems, in which a many-fermion problem is reduced to single-particle (s.p.) equations. Intending to expose the minimal composition and supply a guide for extensive application, we have generalized and reformulated the KS theory through the 1-body density matrix ϱk\varrho_{k\ell}, without specifying principal variables for the energy minimization. Taking into account the pairing tensor, we have also mentioned the Kohn-Sham–Bogolyubov-de Gennes equation.

Whereas the two-step optimization of energy under Levy’s constrained search assures that the Hohenberg-Kohn (HK) theorem holds with a universal energy density functional (EDF), it hampers the differentiability of the EDF and motivates the KS procedure. The KS theory is derived by separating the EDF into regular and irregular parts. While the irregular part is not differentiable for the principal variables, the differentiability for the 1-body density matrix ϱ\varrho is assumed. It is noted that the density matrix ϱ\varrho itself is not and should not be a principal variable, introduced only to remove the irregularity in the EDF.

The issues of representability are readdressed in this context. The vv- and NN-representabilities are distinguished between the HK and KS (or the non-interacting representability) levels. The vv-representability at the KS level is reduced to the differentiability with respect to the principal variables for the regular part of the EDF. Since the KS theory relies on variation, differentiability is crucial to justify it. The solution of the KS equation gives idempotent ϱ\varrho, which discriminates the NN-representabilities between the HK and KS levels. The equational equivalence has been pointed out before and after moving some variables from principal to subsidiary, affecting their physical significance. We have also discussed how these issues are related to practical cases. The minimal composition clarifies criteria for the extendability of the KS theory, which will guide further development of an ab initio density-functional theory including other systems, e.g., atomic nuclei.

Acknowledgements.
The author is grateful to T. Kotani for the discussions.

Appendix A Hartree-Fock-Bogolyubov approximation

In this Appendix, we consider the case in which |Ψ|\Psi\rangle does not necessarily have a fixed particle number, while the expectation value is maintained via Eq. (8). By taking into account the contribution of the pairing tensor in addition to ϱ\varrho,

κk:=Ψ|aak|Ψ,κk=Ψ|aka|Ψ,\kappa_{k\ell}:=\langle\Psi|a_{\ell}a_{k}|\Psi\rangle\,,\quad\kappa_{k\ell}^{\ast}=\langle\Psi|a^{\dagger}_{k}a^{\dagger}_{\ell}|\Psi\rangle\,, (56)

the energy is approximated as E[Ψ]EHFB[ϱ,κ,κ]E[\Psi]\approx E^{\mathrm{HFB}}[\varrho,\kappa,\kappa^{\ast}], which derives the Hartree-Fock-Bogolyubov (HFB) approximation. The independent variables are ϱk\varrho_{k\ell} for all (k,)(k,\ell), κk\kappa_{k\ell} and κk\kappa^{\ast}_{k\ell} with k<k<\ell.

By modifying the energy functional analogously to Eq. (9),

E~HFB:=EHFBμ[tr(ϱ)N],\tilde{E}^{\mathrm{HFB}}:=E^{\mathrm{HFB}}-\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,, (57)

variation yields

δE~HFB=kEHFBϱkδϱk+k>(EHFBκkδκk+EHFBκkδκk)μkδϱkkδμ[tr(ϱ)N]=tr[(hHFBμ)δϱ12(ΔHFBδκ+ΔHFBδκ)]δμ[tr(ϱ)N],\begin{split}\delta\tilde{E}^{\mathrm{HFB}}&=\sum_{k\ell}\frac{\partial E^{\mathrm{HFB}}}{\partial\varrho_{\ell k}}\,\delta\varrho_{\ell k}+\sum_{k>\ell}\Big{(}\frac{\partial E^{\mathrm{HFB}}}{\partial\kappa_{\ell k}}\,\delta\kappa_{\ell k}+\frac{\partial E^{\mathrm{HFB}}}{\partial\kappa^{\ast}_{\ell k}}\,\delta\kappa^{\ast}_{\ell k}\Big{)}-\mu\sum_{k}\delta\varrho_{kk}\\ &\quad-\delta\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\\ &=\mathrm{tr}\Big{[}(h^{\mathrm{HFB}}-\mu)\,\delta\varrho-\frac{1}{2}\big{(}\Delta^{\ast\mathrm{HFB}}\,\delta\kappa+\Delta^{\mathrm{HFB}}\,\delta\kappa^{\ast}\big{)}\Big{]}-\delta\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,,\end{split} (58)

where

hkHFB:=EHFBϱk=(hkHFB),ΔkHFB:=EHFBκk=ΔkHFB,ΔkHFB:=EHFBκk=(ΔkHFB).\begin{split}h^{\mathrm{HFB}}_{k\ell}:=\frac{\partial E^{\mathrm{HFB}}}{\partial\varrho_{\ell k}}=\big{(}h^{\mathrm{HFB}}_{\ell k}\big{)}^{\ast}\,,\quad&\Delta^{\mathrm{HFB}}_{k\ell}:=-\frac{\partial E^{\mathrm{HFB}}}{\partial\kappa^{\ast}_{\ell k}}=-\Delta^{\mathrm{HFB}}_{\ell k}\,,\\ &\Delta^{\ast\mathrm{HFB}}_{k\ell}:=-\frac{\partial E^{\mathrm{HFB}}}{\partial\kappa_{\ell k}}=\big{(}\Delta^{\mathrm{HFB}}_{k\ell}\big{)}^{\ast}\,.\end{split} (59)

By doubling the matrix dimension, the HFB Hamiltonian and the generalized density matrix are defined by Ring and Schuck (1980)

𝖧HFB:=(hHFBμΔHFBΔHFB(hHFB)+μ),\mathsf{H}^{\mathrm{HFB}}:=\begin{pmatrix}h^{\mathrm{HFB}}-\mu&\Delta^{\mathrm{HFB}}\\ -\Delta^{\ast\mathrm{HFB}}&-(h^{\mathrm{HFB}})^{\ast}+\mu\end{pmatrix}\,, (60)

and

𝖱:=(ϱκκ1ϱ)=(Ψ|aak|ΨΨ|aak|ΨΨ|aak|ΨΨ|aak|Ψ),viz.,𝖱T=Ψ|(akak)(aa)|Ψ.\mathsf{R}:=\begin{pmatrix}\varrho&\kappa\\ -\kappa^{\ast}&1-\varrho^{\ast}\end{pmatrix}=\begin{pmatrix}\langle\Psi|a^{\dagger}_{\ell}a_{k}|\Psi\rangle&\langle\Psi|a_{\ell}a_{k}|\Psi\rangle\\ \langle\Psi|a^{\dagger}_{\ell}a^{\dagger}_{k}|\Psi\rangle&\langle\Psi|a_{\ell}a^{\dagger}_{k}|\Psi\rangle\end{pmatrix}\,,\quad\mbox{viz.,}\quad\mathsf{R}^{T}=\big{\langle}\Psi\big{|}\begin{pmatrix}a^{\dagger}_{k}\\ a_{k}\end{pmatrix}\begin{pmatrix}a_{\ell}&a^{\dagger}_{\ell}\end{pmatrix}\big{|}\Psi\big{\rangle}\,. (61)

Equation (58) is rewritten as

δE~HFB=12tr(𝖧HFBδ𝖱)δμ[tr(ϱ)N].\delta\tilde{E}^{\mathrm{HFB}}=\frac{1}{2}\,\mathrm{tr}^{\prime}\big{(}\mathsf{H}^{\mathrm{HFB}}\,\delta\mathsf{R}\big{)}-\delta\mu\big{[}\mathrm{tr}(\varrho)-N\big{]}\,. (62)

The expression tr\mathrm{tr}^{\prime} in the first term of the rhs stands for the trace in the space of 𝖧HFB\mathsf{H}^{\mathrm{HFB}} and 𝖱\mathsf{R}, whose dimension is twice the dimension of the s.p. space. The diagonalization of 𝖧HFB\mathsf{H}^{\mathrm{HFB}} via an appropriate unitary transformation 𝖶HFB\mathsf{W}^{\mathrm{HFB}},

𝖧kHFB𝖶iHFB=εiHFB𝖶kiHFB,\sum_{\ell}\mathsf{H}^{\mathrm{HFB}}_{k\ell}\,\mathsf{W}^{\mathrm{HFB}}_{\ell i}=\varepsilon^{\mathrm{HFB}}_{i}\,\mathsf{W}^{\mathrm{HFB}}_{ki}\,, (63)

is the HFB equation. With denoting the vector composing 𝖶kiHFB\mathsf{W}^{\mathrm{HFB}}_{ki} by 𝐰iHFB\mathbf{w}^{\mathrm{HFB}}_{i}, Eq. (63) is expressed as

𝖧HFB𝐰iHFB=εiHFB𝐰iHFB.\mathsf{H}^{\mathrm{HFB}}\,\mathbf{w}^{\mathrm{HFB}}_{i}=\varepsilon^{\mathrm{HFB}}_{i}\,\mathbf{w}^{\mathrm{HFB}}_{i}\,. (64)

A noteworthy property of 𝖧HFB\mathsf{H}^{\mathrm{HFB}} is

Σx𝖧HFBΣx=(𝖧HFB),Σx:=(0110).\Sigma_{x}\,\mathsf{H}^{\mathrm{HFB}}\,\Sigma_{x}=-(\mathsf{H}^{\mathrm{HFB}})^{\ast}\,,\quad\Sigma_{x}:=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\,. (65)

Equation (65) ensures that, if an eigenvalue and associated eigenvector (εi,𝐰i)(\varepsilon_{i},\mathbf{w}_{i}) is a solution of Eq. (64), (εi,Σx𝐰i)(-\varepsilon_{i},\Sigma_{x}\mathbf{w}_{i}^{\ast}) satisfies Eq. (64) as well, similarly to the RPA solution Nakada (2016). Therefore, it is sufficient to solve Eq. (64) for εi0\varepsilon_{i}\geq 0. Application of the unitary transformation 𝖶\mathsf{W} to 𝖱\mathsf{R} yields

𝖱T𝖶T𝖱T𝖶=Ψ|𝖶T(akak)(aa)𝖶|Ψ.\mathsf{R}^{T}\quad\rightarrow\quad\mathsf{W}^{T}\mathsf{R}^{T}\mathsf{W}^{\ast}=\big{\langle}\Psi\big{|}\mathsf{W}^{T}\begin{pmatrix}a^{\dagger}_{k}\\ a_{k}\end{pmatrix}\begin{pmatrix}a_{\ell}&a^{\dagger}_{\ell}\end{pmatrix}\mathsf{W}^{\ast}\big{|}\Psi\big{\rangle}\,. (66)

The unitary transformation 𝖶\mathsf{W} gives a Bogolyubov transformation. Quasiparticle operators are defined by

(αiαi)=𝖶T(akak).\begin{pmatrix}\alpha^{\dagger}_{i}\\ \alpha_{i}\end{pmatrix}=\mathsf{W}^{T}\begin{pmatrix}a^{\dagger}_{k}\\ a_{k}\end{pmatrix}\,. (67)

Because the transformed 𝖱\mathsf{R} is

𝖶𝖱𝖶=(Ψ|αjαi|ΨΨ|αjαi|ΨΨ|αjαi|ΨΨ|αjαi|Ψ),\mathsf{W}^{\dagger}\mathsf{R}\mathsf{W}=\begin{pmatrix}\langle\Psi|\alpha^{\dagger}_{j}\alpha_{i}|\Psi\rangle&\langle\Psi|\alpha_{j}\alpha_{i}|\Psi\rangle\\ \langle\Psi|\alpha^{\dagger}_{j}\alpha^{\dagger}_{i}|\Psi\rangle&\langle\Psi|\alpha_{j}\alpha^{\dagger}_{i}|\Psi\rangle\end{pmatrix}\,, (68)

diagonal elements in the quasiparticle representation 𝖱ii=Ψ|αiαi|Ψ\mathsf{R}_{ii}=\langle\Psi|\alpha^{\dagger}_{i}\alpha_{i}|\Psi\rangle or Ψ|αiαi|Ψ\langle\Psi|\alpha_{i}\alpha^{\dagger}_{i}|\Psi\rangle satisfy 0𝖱ii10\leq\mathsf{R}_{ii}\leq 1, owing to the Pauli principle.

The minimum of EHFBE^{\mathrm{HFB}} under Constraint (8) needs δE~HFB0\delta\tilde{E}^{\mathrm{HFB}}\geq 0. The associated wave-function is denoted by |ΦHFB|\Phi^{\mathrm{HFB}}\rangle, which is constituted by the solutions of Eq. (63). In the vicinity of its solution, δE~HFB\delta\tilde{E}^{\mathrm{HFB}} is written as

δE~HFB=iεiHFBδ𝖱ii,\delta\tilde{E}^{\mathrm{HFB}}=\sum_{i}\varepsilon^{\mathrm{HFB}}_{i}\,\delta\mathsf{R}_{ii}\,, (69)

where the summation runs over ii having εiHFB>0\varepsilon^{\mathrm{HFB}}_{i}>0. Because εiHFB>0\varepsilon^{\mathrm{HFB}}_{i}>0 and 0𝖱ii10\leq\mathsf{R}_{ii}\leq 1, an argument similar to Eqs. (13) and (14) addresses a conclusion that the HFB solution has 𝖱ii=0\mathsf{R}_{ii}=0 (for εiHFB>0\varepsilon^{\mathrm{HFB}}_{i}>0), namely ΦHFB|αiαi|ΦHFB=0\langle\Phi^{\mathrm{HFB}}|\alpha^{\dagger}_{i}\alpha_{i}|\Phi^{\mathrm{HFB}}\rangle=0. Thus |ΦHFB|\Phi^{\mathrm{HFB}}\rangle is the vacuum of quasiparticles defined by Eq. (67) with 𝖶=𝖶HFB\mathsf{W}=\mathsf{W}^{\mathrm{HFB}}. This is characterized by (𝖱HFB)2=𝖱HFB(\mathsf{R}^{\mathrm{HFB}})^{2}=\mathsf{R}^{\mathrm{HFB}}, where 𝖱HFB:=𝖱[ΦHFB]\mathsf{R}^{\mathrm{HFB}}:=\mathsf{R}[\Phi^{\mathrm{HFB}}].

As the vacuum of the quasiparticles, |ΦHFB|\Phi^{\mathrm{HFB}}\rangle is a pair condensate,

|ΦHFBe𝒵|0,𝒵=12kzkaka,|\Phi^{\mathrm{HFB}}\rangle\propto e^{\mathcal{Z}}|0\rangle\,,\quad\mathcal{Z}=\frac{1}{2}\sum_{k\ell}z_{k\ell}\,a^{\dagger}_{k}a^{\dagger}_{\ell}\,, (70)

where |0|0\rangle is the vacuum of the particles: ak|0=0a_{k}|0\rangle=0 for any kk. Even for systems with particle-number conservation, the HFB approximation is valid when the pair correlation is significant while fluctuation of the particle number in |ΦHFB|\Phi^{\mathrm{HFB}}\rangle does not influence seriously. In |ΦHFB|\Phi^{\mathrm{HFB}}\rangle, the component having N(=even)N\,(=\mathrm{even}) particles is

|ΦN𝒵N/2|0.|\Phi_{N}\rangle\propto\mathcal{Z}^{N/2}|0\rangle\,. (71)

The pairing tensor of Eq. (56) could read as

κkΦN2|aak|ΦN.\kappa_{k\ell}\approx\langle\Phi_{N-2}|a_{\ell}a_{k}|\Phi_{N}\rangle\,. (72)

The energy functional of the form EHFB[ϱ,κ,κ]E^{\mathrm{HFB}}[\varrho,\kappa,\kappa^{\ast}] follows if 𝒞kk(2)κkκk\mathcal{C}^{(2)}_{k\ell k^{\prime}\ell^{\prime}}\approx\kappa_{k^{\prime}\ell^{\prime}}^{\ast}\kappa_{k\ell} and 𝒞(ν)\mathcal{C}^{(\nu)}’s are negligible for ν3\nu\geq 3.

Appendix B Two-step minimization and differentiability

This Appendix gives an elementary example of two-step minimization for a pedagogical purpose, illustrating how the constrained search damages differentiability.

Consider a function ϕ(x,y)\phi(x,y) of the two variables xx and yy,

ϕ(x,y)=3x48x3y+6x2(y2d2).(d>0)\phi(x,y)=3x^{4}-8x^{3}y+6x^{2}(y^{2}-d^{2})\,.\quad(d>0) (73)

This ϕ(x,y)\phi(x,y) is obviously analytical everywhere on the xyxy-plane. The three-dimensional landscape of ϕ(x,y)\phi(x,y) is presented in Fig. 1. The following coupled equations give the extrema of ϕ\phi,

ϕx=12x(xy+d)(xyd)=0,ϕy=4x2(2x3y)=0.\begin{split}\frac{\partial\phi}{\partial x}&=12x(x-y+d)(x-y-d)=0\,,\\ \frac{\partial\phi}{\partial y}&=-4x^{2}(2x-3y)=0\,.\end{split} (74)

The solutions are

(x,y)=(0,arbitrary),(+3d,+2d),(3d,2d).(x,y)=(0,\mbox{arbitrary})\,,~{}(+3d,+2d)\,,~{}(-3d,-2d)\,. (75)

By substituting these solutions and comparing the values of ϕ\phi, one immediately obtains

min(x,y)ϕ(x,y)=27d4at(x,y)=(+3d,+2d)and(3d,2d).\min_{(x,y)}\phi(x,y)=-27d^{4}\quad\mbox{at}~{}(x,y)=(+3d,+2d)~{}\mbox{and}~{}(-3d,-2d)\,. (76)

Refer to caption

Figure 1: Landscape of ϕ(x,y)\phi(x,y) defined in Eq. (73), in the region satisfying 5x/d5-5\leq x/d\leq 5 and 5y/d5-5\leq y/d\leq 5, in the unit of d4d^{4}.

The next task is to find the minimum by a two-step procedure. First, ϕ(x,y)\phi(x,y) is minimized ϕ(x,y)\phi(x,y) with respect to xx for individual yy, yielding a function ϕc(y)\phi_{c}(y),

ϕc(y):=minxϕ(x,y).\phi_{c}(y):=\min_{x}\phi(x,y)\,. (77)

This minimization is accomplished via the first equation of Eq. (74),

ϕc(y)=min[ϕc0(y),ϕc+(y),ϕc(y)]={ϕc0(y)(fory3dandy3d)ϕc+(y)(for0y<3d)ϕc(y)(for3d<y<0),\phi_{c}(y)=\min\big{[}\phi_{c0}(y),\phi_{c+}(y),\phi_{c-}(y)\big{]}=\left\{\begin{aligned} \phi_{c0}(y)&\quad(\mbox{for}~{}y\leq-3d~{}\mbox{and}~{}y\geq 3d)\\ \phi_{c+}(y)&\quad(\mbox{for}~{}0\leq y<3d)\\ \phi_{c-}(y)&\quad(\mbox{for}~{}-3d<y<0)\end{aligned}\right.\,, (78)

where

ϕc0(y)=ϕ(0,y)=0,ϕc±(y)=ϕ(y±d,y)=(y±d)3(y3d).\phi_{c0}(y)=\phi(0,y)=0\,,\quad\phi_{c\pm}(y)=\phi(y\pm d,y)=(y\pm d)^{3}(y\mp 3d)\,. (79)

As shown in Fig. 2, ϕc(y)\phi_{c}(y) takes a minimum at y=±2dy=\pm 2d, satisfying

minyϕc(y)=min(x,y)ϕ(x,y)=27d4.\min_{y}\phi_{c}(y)=\min_{(x,y)}\phi(x,y)=-27d^{4}\,. (80)

However, ϕc(y)\phi_{c}(y) is not differentiable at y=0y=0 and ±3d\pm 3d, as seen in Fig. 2. Thus, the differentiability is lost in ϕc(y)\phi_{c}(y) at the global level, though it could hold within limited domains, e.g., 0<y<3d0<y<3d. The violation of differentiability at the global level indicates the non-existence of universal function, as exemplified in Eq. (78). It is problematic to apply variational procedures without knowing the analyticity of ϕc(y)\phi_{c}(y), possibly hampering practicality.

Refer to caption

Figure 2: ϕc(y)\phi_{c}(y) of Eq. (78) (solid line), and ϕc0(y),ϕc±(y)\phi_{c0}(y),\phi_{c\pm}(y) of Eq. (79) (dashed lines).

Appendix C Energy density functional for 1-dimensional harmonic oscillator

While the original HK theorem assures the existence of an EDF in which the energy is represented as a functional only of the local density n(𝐫)n(\mathbf{r}), modification via the KS theory has been needed in its practical application to many-fermion systems. This fact is linked to the shell structure that often emerges in many-fermion systems. To illustrate this situation and help to comprehend the path from the HK theorem to the KS theory, we consider a simple system under the 1-dimensional harmonic oscillator (h.o.) potential.

Let us take the Hamiltonian of the 1-dimensional h.o.,

H=K+U;K=12md2dx2,U(x)=mω22x2.H=K+U\,;\quad K=-\frac{1}{2m}\,\frac{d^{2}}{dx^{2}}\,,~{}U(x)=\frac{m\omega^{2}}{2}x^{2}\,. (81)

This Hamiltonian gives energy levels ϵν=(ν+12)ω\epsilon_{\nu}=(\nu+\frac{1}{2})\omega (ν=0,1,2,)(\nu=0,1,2,\cdots), which can be regarded as orbits. We assume that each level has Ω\Omega-fold degeneracy101010 This is accomplished when the particle has the spin (Ω1)/2(\Omega-1)/2. It can also be connected to a higher-dimensional spherical system by identifying xx as the radial coordinate and averaging out the angular and spin degrees of freedom., and NN non-interacting fermions occupy the orbits. All particles occupy the ν=0\nu=0 orbit at the g.s. for 0<NΩ0<N\leq\Omega, (NΩ)(N-\Omega) particles occupy the ν=1\nu=1 orbit for Ω<N2Ω\Omega<N\leq 2\Omega, and so forth, providing the g.s. energy and density,

EN(0)={12Nω(0<NΩ)[12Ω+32(NΩ)]ω(Ω<N2Ω),nN(0)(x)={mωπNemωx2(0<NΩ)mωπ[Ω+2(NΩ)mωx2]emωx2(Ω<N2Ω).\begin{split}&E^{(0)}_{N}=\left\{\begin{array}[]{ll}\frac{1}{2}N\omega&(0<N\leq\Omega)\\ \big{[}\frac{1}{2}\Omega+\frac{3}{2}(N-\Omega)\big{]}\omega&(\Omega<N\leq 2\Omega)\\ \quad\vdots&\end{array}\right.\,,\\ &\quad n^{(0)}_{N}(x)=\left\{\begin{array}[]{ll}{\displaystyle\sqrt{\frac{m\omega}{\pi}}\,N\,e^{-m\omega x^{2}}}&(0<N\leq\Omega)\\ {\displaystyle\sqrt{\frac{m\omega}{\pi}}\,\big{[}\Omega+2(N-\Omega)m\omega x^{2}\big{]}\,e^{-m\omega x^{2}}}&(\Omega<N\leq 2\Omega)\\ \quad\vdots&\end{array}\right.\,.\end{split} (82)

This system is treated by the EDF under the HK theorem with 𝐐=n(x)\mathbf{Q}=n(x).

The minimization of Eq. (16) can be substituted by the Legendre transformation around the minimum. The main task is constructing the kinetic part of the EDF. The kinetic part of the EDF is desirable not to depend on UU, which is conjugate to 𝐐(=n(x))\mathbf{Q}\,(=n(x)). For 0<NΩ0<N\leq\Omega, the following EDF Stringari and Treiner (1987) is exact,

EHK[n]=𝑑x[18m(n(x))2n(x)+U(x)n(x)].E^{\mathrm{HK}}[n]=\int dx\,\bigg{[}\frac{1}{8m}\,\frac{\big{(}n^{\prime}(x)\big{)}^{2}}{n(x)}+U(x)\,n(x)\bigg{]}\,. (83)

Indeed, the variational equation of EHK[n]E^{\mathrm{HK}}[n] with respect to n(x)n(x) under the constraint 𝑑xn(x)=N\int dx\,n(x)=N,

δE~HK[n]=0;E~HK[n]:=EHK[n]μ[𝑑xn(x)N],\delta\tilde{E}^{\mathrm{HK}}[n]=0\,;\quad\tilde{E}^{\mathrm{HK}}[n]:=E^{\mathrm{HK}}[n]-\mu\bigg{[}\,\int dx\,n(x)-N\bigg{]}\,, (84)

yields

14m[n′′(x)n(x)12{n(x)n(x)}2]U(x)+μ=0,\frac{1}{4m}\bigg{[}\frac{n^{\prime\prime}(x)}{n(x)}-\frac{1}{2}\Big{\{}\frac{n^{\prime}(x)}{n(x)}\Big{\}}^{2}\bigg{]}-U(x)+\mu=0\,, (85)

coinciding with Eq. (82) when U(x)=mω2x2/2U(x)=m\omega^{2}x^{2}/2 and μ=ϵ0\mu=\epsilon_{0}.

The EDF of Eq. (83) is not appropriate as it is in N>ΩN>\Omega. The exact energy and density of Eq. (82) for Ω<N2Ω\Omega<N\leq 2\Omega do not satisfy Eq. (84). On the other hand, it is always possible to obtain an exact functional around a fixed N=N0N=N_{0} in terms of Δn(x):=n(x)nN0(0)(x)\mathit{\Delta}n(x):=n(x)-n^{(0)}_{N_{0}}(x),

EHK[Δn]EN0(0)=𝑑x[18m(Δn(x))2Δn(x)+U(x)Δn(x)],E^{\mathrm{HK}}[\mathit{\Delta}n]-E^{(0)}_{N_{0}}=\int dx\,\bigg{[}\frac{1}{8m}\,\frac{\big{(}\mathit{\Delta}n^{\prime}(x)\big{)}^{2}}{\mathit{\Delta}n(x)}+U(x)\,\mathit{\Delta}n(x)\bigg{]}\,, (86)

because Δn(x)|φνF(x)|2\mathit{\Delta}n(x)\propto|\varphi_{\nu_{\mathrm{F}}}(x)|^{2} (νF\nu_{\mathrm{F}} is the orbit just above N0N_{0}). However, Eq. (86) does not hold across N=ΩN=\Omega; the EDF of (86) is not correct if N0<Ω<NN_{0}<\Omega<N. For Eq. (86) to be valid for N>ΩN>\Omega, nN0(0)(x)n^{(0)}_{N_{0}}(x) and EN0(0)E^{(0)}_{N_{0}} have to be redefined by shifting N0N_{0}. This result provides evidence that the EDF guaranteed by the HK theorem (Eq. (16)) is not regular everywhere. This irregularity comes out because of the shell structure, i.e., because the particles start occupying the ν=1\nu=1 orbit for N>ΩN>\Omega. The source of the irregularity is the kinetic energy.

In contrast, the g.s. energy is immediately written with the 1-body density matrix in the present non-interacting problem. In the coordinate representation, the density matrix is ϱ(x,x):=Ψ|ψ(x)ψ(x)|Ψ\varrho(x,x^{\prime}):=\langle\Psi|\psi^{\dagger}(x^{\prime})\,\psi(x)|\Psi\rangle, where ψ(x)\psi(x) and ψ(x)\psi^{\dagger}(x) are field operators. The energy is exactly expressed as

E[ϱ]=𝑑x𝑑xδ(xx)[12m2xxϱ(x,x)+U(x)ϱ(x,x)].E[\varrho]=\int dx\,dx^{\prime}\,\delta(x-x^{\prime})\bigg{[}\frac{1}{2m}\,\frac{\partial^{2}}{\partial x\,\partial x^{\prime}}\varrho(x,x^{\prime})+U(x)\,\varrho(x,x^{\prime})\bigg{]}\,. (87)

Identifying the kinetic term as the irregular part and handling it as discussed in Sec. IV, the EDF reaches

EKS[ϱ;n]=12m𝑑x𝑑xδ(xx)2xxϱ(x,x)+𝑑xU(x)n(x)=12m𝑑x𝑑xδ′′(xx)ϱ(x,x)+𝑑xU(x)n(x),\begin{split}E^{\mathrm{KS}}[\varrho;n]&=\frac{1}{2m}\int dx\,dx^{\prime}\,\delta(x-x^{\prime})\,\frac{\partial^{2}}{\partial x\,\partial x^{\prime}}\varrho(x,x^{\prime})+\int dx\,U(x)\,n(x)\\ &=-\frac{1}{2m}\int dx\,dx^{\prime}\,\delta^{\prime\prime}(x-x^{\prime})\,\varrho(x,x^{\prime})+\int dx\,U(x)\,n(x)\,,\end{split} (88)

equivalent to Eq. (87). The EDF of (88) is global and differentiable with respect to (ϱ,n)(\varrho,n).

References

  • Parr and Yang (1989) R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules (Oxford Univ. Press, Oxford, 1989).
  • Giuliani and Vignale (2005) G. F. Giuliani and G. Vignale, Quantum Theory of the Electron Liquid (Cambridge Univ. Press, Cambridge, 2005).
  • Engel and Dreizler (2011) E. Engel and R. M. Dreizler, Density Functional Theory (Springer-Verlag, Berlin, 2011).
  • Sahni (2016) V. Sahni, Quantal Density Functional Theory, 2nd ed. (Springer-Verlag, Berlin, 2016).
  • Hohenberg and Kohn (1964) P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).
  • Kohn and Sham (1965) W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).
  • von Barth and Hedin (1972) U. von Barth and L. Hedin, J. Phys. C 5, 1629 (1972).
  • Gunnarsson and Lundqvist (1976) O. Gunnarsson and B. I. Lundqvist, Phys. Rev. B 13, 4274 (1976).
  • Jacob and Reiher (2012) C. R. Jacob and M. Reiher, Int. J. Quant. Chem. 112, 3661 (2012).
  • Vignale and Rasolt (1987) G. Vignale and M. Rasolt, Phys. Rev. Lett. 59, 2360 (1987).
  • Vignale and Rasolt (1988) G. Vignale and M. Rasolt, Phys. Rev. B 37, 10685 (1988).
  • Oliveira et al. (1988) L. N. Oliveira, E. K. U. Gross,  and W. Kohn, Phys. Rev. Lett. 60, 2430 (1988).
  • Lüders et al. (2005) M. Lüders, M. A. L. Marques, N. N. Lathiotakis, A. Floris, G. Profeta, L. Fast, A. Continenza, S. Massidda,  and E. K. U. Gross, Phys. Rev. B 72, 024545 (2005).
  • Donnely and Parr (1978) R. A. Donnely and R. G. Parr, J. Chem. Phys. 69, 4431 (1978).
  • Zumbach and Maschke (1985) G. Zumbach and K. Maschke, J. Chem. Phys. 82, 5604 (1985).
  • Schindlmayr and Godby (1995) A. Schindlmayr and R. W. Godby, Phys. Rev. B 51, 10427 (1995).
  • López-Sandoval and Pastor (2000) R. López-Sandoval and G. M. Pastor, Phys. Rev. B 61, 1764 (2000).
  • Requistl and Pankratov (2008) R. Requistl and O. Pankratov, Phys. Rev. B 77, 235121 (2008).
  • Pernal and Giesbertz (2016) K. Pernal and K. J. H. Giesbertz, Top. Curr. Chem. 368, 125 (2016).
  • Higuchi and Higuchi (2004) M. Higuchi and K. Higuchi, Phys. Rev. B 69, 035113 (2004).
  • Mermin (1965) N. D. Mermin, Phys. Rev. 137, A1441 (1965).
  • Runge and Gross (1984) E. Runge and E. K. U. Gross, Phys. Rev. Lett. 52, 997 (1984).
  • Gross and Kohn (1985) E. K. U. Gross and W. Kohn, Phys. Rev. Lett. 55, 2850 (1985).
  • Vautherin and Brink (1972) D. Vautherin and D. M. Brink, Phys. Rev. C 5, 626 (1972).
  • Skyrme (1959) T. H. R. Skyrme, Nucl. Phys. 9, 615 (1959).
  • Petkov and Stoitsov (1991) I. Z. Petkov and M. V. Stoitsov, Nuclear Density Functional Theory (Oxford Univ. Press, Oxford, 1991).
  • Colò (2020) G. Colò, Adv. Phys.: X 5, 1740061 (2020).
  • Dechargé and Gogny (1980) J. Dechargé and D. Gogny, Phys. Rev. C 21, 1568 (1980).
  • Nakada (2003) H. Nakada, Phys. Rev. C 68, 014316 (2003).
  • Serot and Walecka (1986) B. D. Serot and J. D. Walecka, “The Relativistic Nuclear Many-Body Problem,” in Advances in Nuclear Physics, Vol. 16, edited by J. W. Negele and E. Vogt (Plenum, New York, 1986) pp. 1–320.
  • Ring (2016) P. Ring, “Concept of covariant density functional theory,” in International Review of Nuclear Physics, Vol. 10, edited by J. Meng (World Scientific, Singapore, 2016) pp. 1–20.
  • Thomas (1927) L. H. Thomas, Math. Proc. Camb. Phil. Soc. 23, 542 (1927).
  • Fermi (1928) E. Fermi, Z. Phys. 48, 73 (1928).
  • Deb and Chattaraj (1988) B. M. Deb and P. K. Chattaraj, Phys. Rev. A 37, 4030 (1988).
  • Deb and Chattaraj (1992) B. M. Deb and P. K. Chattaraj, Phys. Rev. A 45, 1412 (1992).
  • Pearson et al. (1993) M. Pearson, E. Smargiassi,  and P. A. Madden, J. Phys.: Cond. Matt. 5, 3221 (1993).
  • Ring and Schuck (1980) P. Ring and P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, New York, 1980).
  • Reichl (1980) L. E. Reichl, A Modern Course in Statistical Physics (Univ. Texas Press, Austin, 1980).
  • Lieb (1981) E. H. Lieb, Phys. Rev. Lett. 46, 457 (1981).
  • Gilbert (1975) T. L. Gilbert, Phys. Rev. B 12, 2111 (1975).
  • Fukuda et al. (1994) R. Fukuda, T. Kotani, Y. Suzuki,  and S. Yokojima, Prog. Theor. Phys. 92, 833 (1994).
  • Valiev and Fernando (1997) M. Valiev and G. W. Fernando,   (1997), arXiv:cond-mat.str-el/9702247 [cond-mat.str-el] .
  • Levy (1979) M. Levy, Proc. Natl. Acad. Sci. USA 76, 6062 (1979).
  • Coleman (1963) A. J. Coleman, Rev. Mod. Phys. 35, 668 (1963).
  • Harriman (1981) J. E. Harriman, Phys. Rev. A 24, 680 (1981).
  • Levy (1982) M. Levy, Phys. Rev. A 26, 1200 (1982).
  • Capelle and Vignale (2001) K. Capelle and G. Vignale, Phys. Rev. Lett. 86, 5546 (2001).
  • Kvaal et al. (2014) S. Kvaal, U. Ekström, A. M. Teale,  and T. Helgaker, J. Chem. Phys. 140, 18A518 (2014).
  • Valone (1980) A. M. Valone, J. Chem. Phys. 73, 1344 (1980).
  • Messiah (1961) A. Messiah, Quantum Mechanics, vol. 2 (Elsevier, Amsterdam, 1961).
  • Sahni et al. (1988) V. Sahni, K.-P. Bohnen,  and M. K. Harbola, Phys. Rev. A 37, 1895 (1988).
  • Perdew et al. (1996) J. P. Perdew, K. Burke,  and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • Tao et al. (2003) J. Tao, J. P. Perdew, V. N. Staroverov,  and G. E. Scuseria, Phys. Rev. Lett. 91, 146401 (2003).
  • Nagai et al. (2022) R. Nagai, R. Akashi,  and O. Sugino, Phys. Rev. Res. 4, 013106 (2022).
  • Drut et al. (2010) J. E. Drut, R. J. Furnstahl,  and L. Platter, Prog. Part. Nucl. Phys. 64, 120 (2010).
  • Engel (2007) J. Engel, Phys. Rev. C 75, 014306 (2007).
  • Nakada (2020) H. Nakada, Int. J. Mod. Phys. E 29, 1930008 (2020).
  • Barnea (2007) N. Barnea, Phys. Rev. C 76, 067302 (2007).
  • Messud et al. (2009) J. Messud, M. Bender,  and E. Suraud, Phys. Rev. C 80, 054314 (2009).
  • Messud (2013) J. Messud, Phys. Rev. C 87, 024302 (2013).
  • Negele and Vautherin (1972) J. W. Negele and D. Vautherin, Phys. Rev. C 5, 1472 (1972).
  • Campi and Sprung (1972) X. Campi and D. W. Sprung, Nucl. Phys. A 194, 401 (1972).
  • Kortelainen et al. (2012) M. Kortelainen, J. McDonnell, W. Nazarewicz, P.-G. Reinhard, J. Sarich, N. Schunck, M. V. Stoitsov,  and S. M. Wild, Phys. Rev. C 85, 024304 (2012).
  • Kortelainen et al. (2014) M. Kortelainen, J. McDonnell, W. Nazarewicz, E. Olsen, P.-G. Reinhard, J. Sarich, N. Schunck, S. M. Wild, D. Davesne, et al., Phys. Rev. C 89, 054314 (2014).
  • Mahaux et al. (1985) C. Mahaux, P. F. Bortignon, R. A. Broglia,  and C. H. Dasso, Phys. Rep. 120, 1 (1985).
  • Kaiser et al. (2003) N. Kaiser, S. Fritsch,  and W. Weise, Nucl. Phys. A 724, 47 (2003).
  • Navarro Pérez et al. (2018) R. Navarro Pérez, N. Schunck, A. Dyhdalo, R. J. Furnstahl,  and S. K. Bogner, Phys. Rev. C 97, 054304 (2018).
  • Shen et al. (2019) S. Shen, H. Liang, W. H. Long, J. Meng,  and P. Ring, Prog. Part. Nucl. Phys. 109, 103713 (2019).
  • Baldo et al. (2008) M. Baldo, P. Schuck,  and X. Viñas, Phys. Lett. B 663, 390 (2008).
  • Bohr and Mottelson (1975) A. Bohr and B. R. Mottelson, Nuclear Structure, vol. 2 (Benjamin, Reading, 1975).
  • Giraud et al. (2008) B. G. Giraud, B. K. Jennings,  and B. R. Barrett, Phys. Rev. A 78, 032507 (2008).
  • Nakada (2016) H. Nakada, Prog. Theor. Exp. Phys. 2016, 063D02 (2016).
  • Stringari and Treiner (1987) S. Stringari and J. Treiner, Phys. Rev. B 36, 8369 (1987).