Skip to Main Content
Skip Nav Destination

The adiabatic connection formalism yielding an exact expression to exchange–correlation functional, has been instrumental in building DFT approximations. Here, we derive and analyze the density-fixed adiabatic connection (AC), which gives an exact expression of the DFT’s exchange–correlation functional. We give an outline of the exact features of the key AC quantities, yielding physical constraints for constructing approximations to the exchange–correlation functional. Lastly, inspired by recent progress in the AC developments in wave function theories, we discuss the similarities and differences between the AC in DFT and wave function formalisms.

Over the last half-century, density functional theory (DFT) has made an unparalleled impact on a variety of scientific and technological disciplines. Its remarkable balance between computational efficiency and predictive accuracy has rendered it the most widely-used electronic structure method with estimates suggesting that a staggering one-third of the world’s supercomputing power is deployed to run DFT simulations.1  For a range of systems and chemical reactions, current DFT methodology is now capable of producing results that can rival in terms of the accuracy far more expensive wave function simulations. In the era of artificial intelligence, machine learning models trained on DFT results have significantly augmented the power of DFT. These advancements now enable simulations on larger time and length scales than were previously feasible.

A vast majority of DFT simulations employ the Kohn–Sham DFT (KS DFT) scheme. Within this practical framework, the exchange–correlation (XC) functional is the sole energy component that requires approximation. Even though it does not constitute the major component of the total electronic energy, the XC functional encodes the coulombic, fermionic, and quantum-mechanical nature of electrons. Approaches for developing XC approximations span a wide spectrum, from rigorous derivations firmly rooted in quantum mechanical principles to empirical fits and extensive databases. Even machine learning-derived functional forms for the XC have begun to compete with their human-designed counterparts.2–9  In this chapter, we take a step back to the foundational concepts of the XC functional, centering our discussion around the density-fixed AC formalism – a framework that offers an exact formula for the XC functional. We note here that computing the exact XC functional via the AC formalism carries a heavier computational burden than solving the Schrödinger equation. Nevertheless, as a powerful tool for both, constructing and analysing XC approximations, the density-fixed AC formalism played a pivotal role in the development of DFT over the past 30 years. Specifically, this approach provides a number of valuable constraints for XC functional approximations and a range of approximations to the XC functional have been built by directly approximating the AC path.10–15  The AC formalism has been also used to rationalize parameters used in hybrid12,16,17  and double hybrid functionals.18–20  Furthermore, by studying the strong interaction limit of the AC new ingredients for the construction of the XC approximation have been designed.

The idea of the adiabatic connection (AC) formalism is to employ a continuous path between the non-interacting KS system and the exact many-electron quantum system while keeping the ground-state electronic density function fixed.21–23  This path is obtained by linking the non-interacting KS system with the physically interacting one through scaling of the electron–electron interaction. As a result, the AC formalism allows the derivation of the exact XC functional using only the electron–electron repulsion operator, even though the XC functional contains both, electron–electron repulsion component and a kinetic energy component.

In Section 2, we derive the AC based expression of the XC functional. It is an integral, where the integrand is a parameter dependent global density functional. In Section 3, we list the exact known conditions on the AC integrand. We provide an overview of constructing approximations to the XC functionals based on the AC framework in Section 4. Then, in Section 5, we elaborate briefly on the connection between the global and local AC integrands. Section 6 showcases the strong interaction limit in DFT and its link to the AC. Finally, in Section 7, we review the similarities and differences between the DFT’s adiabatic connection and that used in the Møller–Plesset theory.

In Levy’s constrained search formalism,24  the ground-state energy and electronic density function n of a many-electron quantum system in an external potential v is given by:
(1)
where F[n] is the universal HK functional:
(2)
with T ˆ the kinetic energy operator and V ˆ ee the electron–electron (Coulomb) repulsion operator. In eqn (2), the minimization is performed over all antisymmetric wave functions integrating to the given density n. On the other hand, the minimization in eqn (1) is performed over all N-representable densities.24  In the KS framework, F[n] is commonly partitioned in the following way:
(3)
where Ts[n] is the KS non-interacting kinetic energy, U[n] is the Hartree (mean-field) energy and Exc[n] is the exchange–correlation energy.
The adiabatic connection formula for the XC functional has the following expression:21,22 
(4)
where Wλ[n] is the global (integrated over all space) AC integrand:
(5)
and Ψλ[n] is a wave function that minimizes the λ dependent universal HK functional yielding n(r).
To prove eqn (4), we employ the constrained–search approach on a many-electron quantum system that is governed by the following Hamiltonian:
(6)
where V ˆ λ is an operator representing a one-body local potential vλ(r) which ensures that Ψλ yields the same n(r) of the physical system (λ = 1) for every λ ≥ 0. Thus, at λ = 1, vλ(r) reduces to the usual external (nuclear) potential. The use of eqn (6), relies on the v-representability assumption, i.e. existence of V ˆ λ for each λ. The above Hamiltonian, Ĥλ, reduces to the KS non-interacting Hamiltonian in the case of λ = 0 and corresponds to the interacting system for λ = 1.
Using the HK DFT formalism, it is possible to define a universal functional associated to the system of eqn (6) for each value of λ,
(7)
where Ψλ[n] is the minimiser of the above formula for Fλ[n]. It corresponds to the wave function describing a constrained ground-state of the λ dependent system governed by Ĥλ.
Simply utilizing the fact that the integral of a derivative gives the difference of the integrand end points, we have the following equality:
(8)
The left-hand side of eqn (8) consists of two specific universal functionals,
according to eqn (3), and,
because minimising the kinetic energy for a constrained wave function is equivalent to the non-interacting kinetic energy of n. Thus, the F1[n] − F0[n] difference is equal to the sum of the Hartree and XC functionals (according to eqn (3)),
(9)
Furthermore, we also have the following integral equality:
(10)
Eqn (10) is derived with the stationarity condition inside the minimisation of Fλ[n] with respect to any variation in λ. It reads
(11)
which follows from Ψ λ [ n ] | T ˆ + λ V ˆ ee | Ψ λ [ n ] achieving its minimum at λ′ = λ, while Ψλ keeps the density constant.
By combining eqn (9) and (10), we obtain:
(12)
Subtracting the Hartree energy, U[n], on both sides of eqn (12) yields the formula of the exchange–correlation energy based on the adiabatic connection framework (see eqn (4))

It is interesting to observe the global AC integrand Wλ[n] at specific λ points. For instance, when λ = 0, then W0[n] is equal to the exact exchange EX[n]. When λ = 1, Ψ1[n] is the minimizer for F1[n] = F[n], which corresponds to the constrained full interacting ground-state wave function. Hence, W1[n] is the electron–electron repulsion energy of the many-electron quantum system without Coulomb interaction, which is equivalent to the XC energy from which we substract the difference between the true and non-interacting kinetic energy.

Fig. 1 illustrates the nature of the global AC integrand. Integration between λ = 0 and λ = 1 corresponds to the shaded area below the function, which also stands for the exact XC energy according to eqn (4). The figure illustrates the monotonically decreasing nature of Wλ[n], and also illustrates the strong coupling limit when λ → ∞. More properties of Wλ[n] are listed in the following section.

Fig. 1

An illustration of the adiabatic connection integrand curve with a shaded region that represents the exchange–correlation energy.

Fig. 1

An illustration of the adiabatic connection integrand curve with a shaded region that represents the exchange–correlation energy.

Close modal

The AC formalism provides a rigorous way for the development of new classes of density functional approximations. Thus, we want to emphasize in this chapter the exact features of the AC integrand, because they are most relevant for the construction of exchange–correlation functionals in DFT. The following is a list of several essential conditions:

  • It has been proven that the AC integrand, Wλ[n], monotonically decreases with respect to λ,25,26  i.e.
    (13)
    Along with the assumption of a monotonically decreasing nature, it is generally taken that the function Wλ[n] always exhibits convexity. This assumption is supported by numerical evidence, but a formal proof is still lacking to the best of our knowledge. It is worth noting that, theoretically, convexity could be violated in cases of phase transitions along the adiabatic connection path, where the AC curves are likely piecewise convex. For instance, the AC corresponding to the uniform electron gas at a density lower than that of the ferromagnetic transition might exhibit a slight discontinuity. However, the density constraints imposed on Wλ[n] drastically reduce the likelihood of such occurrences.
  • A specific scaling property is satisfied by the AC integrand linking AC with the uniform coordinate scaling:27 
    (14)
    where nγ(r) is a scaled density:
    (15)
    for γ > 0, which conserves the number of electrons. When γ increases, the density shrinks and when it decreases, the density spreads out. Using this uniform scaling relation one can also recover Wλ[n] from Exc[n] with
    (16)
    The exact relation of eqn (16) applying to the exact XC functional has also been used to analyse the approximations and see how for small systems (e.g., hydrogen dimer along its dissociation curve) the AC curves arising from approximate XC functionals (e.g., PBE28 ), deviate from the exact counterpart calculated directly from eqn (5).15,29,30  Also, the behaviour of the XC functional in the high- (γ → 0) and low-density limits25,31  (γ → ∞) are known:
    (17)
    Overall, from eqn (14), (16) and (17) we can see that the uniform density scaling and the adiabatic connection approaches in DFT are deeply interlinked.
  • Using the differentiability with respect to λ allows an expansion of the global AC integrand in the non-interacting limit case (small λ). It is given by:32 
    (18)
    where the AC integrand for λ = 0 corresponds to the exact exchange functional W0[n] = Ex[n], which is given in terms of the occupied KS orbitals:
    (19)
    Furthermore, the slope of Wλ is given by two times the correlation energy of the second-order Görling–Levy perturbation theory (GL2):32,33  W 0 [ n ]=2 E c GL2 [ n ] . This correlation energy can be also expressed in terms of occupied and unoccupied KS orbitals and their energies:
    (20)
    where the indices i, j and a, b run over occupied and non-occupied KS orbitals respectively. The second term in eqn (20) accounts for the difference between the local KS potential, v x KS [ n ] ( r ) = n ( r ) | r r | d r , and the non-local Hartree–Fock (HF) exchange potential, v x HF [ n ] ( r ) = E x [ n ] n ( r ) , evaluated with KS orbitals.
  • A similar expansion of the AC integrand exists in the strongly-interacting limit, i.e. λ → ∞:34,35 
    (21)
    Thereby, W[n] and W [ n ] have a completely different structure to the previous functionals, W0[n] and W 0 [ n ] . They are highly non-local with respect to the density and a strictly-correlated electrons (SCE) approach34  has been designed for their evaluation. Note that W [ n ] is denoted here as the second term in the expansion after W[n], at large λ. In fact, W [ n ] does not correspond to the derivative of Wλ in the λ → ∞ limit. Its denotation resembles its role in the strong coupling limit, analogous to the one of W 0 [ n ] in the weak coupling limit.34,36 
  • The Lieb–Oxford inequality31,37–40  provides a bound from below for the AC integrand:
    (22)

    The exact value for the CLO constant is not known, but it is rigorously proven to lie in between 1.4442 and 1.5765.41–43  More generally, C LO n 4/3 ( r ) d r bounds from below the indirect energy (electron–electron repulsion minus the Hartree energy) of any correctly normalized and antisymmetric Ψ[n]. Letting Ψ[n] be Ψλ[n], we obtain eqn (22). In view of the first AC constraint we listed here, W[n] will be the smallest value for the l.h.s. of eqn (22). With this in mind, finding lower bounds for the optimal constant CLO reduces to searching for a density n for which the ratio between W[n] and n 4/3 ( r ) d r ,44,45  will be the greatest. This procedure has been indeed applied for searching the optimal CLO and understanding for what type of densities CLO is large.39,40,44–46 

Usually, the development of approximations to Wλ[n] aim to obey as many of the exact constraints listed in Section 315  as possible, and then utilize the density functional approximations (DFAs) formulas for an evaluation. One of the first eminent AC-hybrid models was introduced by Becke who applied a linear model for the AC integrand.47  In fact, Levy provided a proof showing the linearity of Wλ[n]27  for small enough λ. There are many more approximation models for the AC integrand, such as the [1/1] Padé approximant,10  coming from the MCY family of functionals.48  Teale et al. employed ab initio arguments to propose some models for Wλ[n].26  Additionally, the interaction strength interpolation (ISI) class of functionals, firstly introduced by Seidl and co-workers49,50  and later revised,15,34,35  is a group of functionals based on an interpolation between the weak and strong coupling limit of AC DFT. By including the information from both weak- and strong-coupling regimes, the ISI framework circumvent the prevalent bias towards weak correlation, a characteristic feature of a vast majority of XC approximations. A potential problem of the ISI functionals is their lack of size consistency, which, however, can be easily corrected by using the size-consistency correction of ref. 51 when there are no degeneracies. Within the DFT realm, the ISI functionals have been tested against several chemical data sets, showing that they perform reasonably well for interaction energies (energy differences).51–53  However, their application has proven to be far more successful outside the DFT realm, particularly when used in conjunction with Hartree–Fock, as opposed to DFT orbitals. These successes and underlying theoretical framework will be discussed in Section 7.

Instead of modelling the global AC integrand (integrated over all space) Wλ[n], one would have more flexibility in the construction of XC functional when modelling the local AC integrand.15  This local AC integrand, denoted as wλ(r), is the λ-dependent XC energy density per particle. In this way, wλ(r) integrates back to the global AC integrand Wλ[n] when multiplied by the density:
(23)
Unfortunately, there is no unique expression for the local AC integrand as adding any function to wλ(r) which integrates to zero when multiplied by the density would yield the same functional Wλ[n]. Therefore, we have to be careful when specifying a definition (gauge) for wλ(r). A commonly accepted gauge for the local AC integrand is the one of the electrostatic potential of the XC hole.15,54–56  In the literature, there are also other definitions for the local AC integrand (see ref. 46, 57 and 58), some of which we will briefly discuss later.
Using the gauge of the electrostatic potential of the XC hole leads to the following expression for the local AC integrand:
(24)
where h xc λ (r, r′) is the XC hole associated to the wave function Ψλ through
(25)
with P 2 λ ( r , r ) being the pair-density that reads (in its spinless form)
(26)
Since the XC hole is the only quantity in the local AC integrand that depends on λ, the resulting formula for the XC functional with the corresponding global AC integrand reads
(27)
where
is the coupling-constant-integrated (λ-averaged) XC hole. It leads to the interpretation of Exc[n] as the electrostatic interaction energy of an electron with its coupling-constant-integrated correlation hole. As a result, the XC functional takes the following form:
(28)
where
(29)
is the XC energy density per particle, which is also a functional of the density.

Approximations to XC functionals have been built by constructing εxc[n](r) through an interpolation at each point in space between w0(r) and w(r) (local interpolations).

Compared to global interpolations, these local interpolations usually deliver superior results for small chemical systems, as noted in previous studies.15,30  However, they typically do not fix all the shortcomings found with global interpolations.30  Despite these observations, the accuracy of XC functionals based on local interpolation between w0(r) and w(r) is largely unexplored and will likely benefit from the use of specific machine learning techniques.59  Specifically, combining this local AC interpolation with deep learning-based ‘local hybrids’59  holds promise for improving XC approximations, particularly when it comes to treating the long standing problem of DFT’s strong correlation.

In Fig. 2, we illustrate the advantage of the use of a specific definition of wλ(r) given in terms of the electrostatic potential of the XC hole (that of eqn (24)) over other definitions. In the top panel of Fig. 2, we show w0(r) and w(r) given by eqn (24) for the beryllium atom. We can see that w(r) is always below w0(r). Thus, the following behaviour of the global AC integrand W0[n] ≥ W[n] arising from eqn (13) also holds locally here w0(r) ≥ w(r), provided that the definition of eqn (24) for w(r) is used. Having w(r) always lying below w0(r) makes it easier to build interpolation models between these two quantities, as the resulting model for εxc[n](r) that would be obtained via such interpolation would lie exactly between w(r) and w0(r). To show that w0(r) ≥ w(r) can be easily violated within other local AC definitions despite W0[n] ≥ W[n] holding, we consider the following virial definitions of w0(r) and w(r):
(30)
(31)
where v0(r) and v(r) are functional derivatives of W0[n] and W[n] w.r.t. n, respectively. The results for w 0 virial ( r ) and w virial ( r ) for the beryllium atom are shown in the bottom panel of Fig. 2. From there we can see that the behaviour of w 0 virial ( r ) and w virial ( r ) is far less predictable than that of w0(r) and w(r) shown in the top panel of Fig. 2. First, the shape of w 0 virial ( r ) and w virial ( r ) is very different. Second, there is a region where w 0 virial ( r ) is below w virial ( r ) . In addition to the virial gauge of eqn (30), similar unpredictable trends have been observed for other gauges (i.e. definition of local AC integrands) in the literature.46  Echoing the observations made in Fig. 2, it was determined that only when the wλ(r) adheres to the definition given in eqn (24), w(r) does maintain a similar shape to w0(r) and consistently falls below the latter. These findings substantiate the idea that the specific definition of the local AC integrand given by eqn (24) is likely the optimal choice for modeling the XC functional when it involves interpolation between w0(r) and w(r).
Fig. 2

Top panel: w0(r) and w(r) given in terms of eqn (24) for the beryllium atom. Bottom panel: w0(r) and w(r) given in terms of eqn (30) for the beryllium atom.

Fig. 2

Top panel: w0(r) and w(r) given in terms of eqn (24) for the beryllium atom. Bottom panel: w0(r) and w(r) given in terms of eqn (30) for the beryllium atom.

Close modal

In previous two sections we discussed strategies for building approximations for the XC exchange functional by using the information from both the weak- and strong-coupling limit of DFT [e.g., w0(r) and w(r) shown in the top panel of Fig. 2]. The key idea of this approach is that our description of the physical case (λ = 1) can be significantly improved if we have knowledge about both λ = 0 and λ = ∞ scenarios. This becomes particularly relevant when trying to describe systems characterised by strong correlation, whose treatment remained (one of) the biggest challenges in the field of DFT development.60–63  In this section, we will briefly describe the main characteristics of the strong-interaction limit of DFT, which corresponds to the λ → ∞ limiting case of our density-fixed adiabatic connection (see ref. 64 for a more comprehensive review of the topic).

As said, the strongly interacting limit of DFT corresponds to the situation when λ tends to infinity. Thus, mathematical description of this limit requires finding Ψ that integrates to n and minimizes T ˆ + λ V ˆ ee when λ is set to ∞. It has been shown that the problem is equivalent to finding Ψ, which integrates to n and minimizes solely V ˆ ee . The strictly correlated electron approach (SCE)64–66  has been developed to give a candidate for such Ψ minimizer. The SCE functional reads as:
(32)
and in terms of the λ-dependent universal functional (eqn (7)) it is given by:
(33)
While it has been proven that the SCE state provides the absolute minimum on the r.h.s. of eqn (33) exclusively for N = 2 (across any dimensionality), with N the number of electrons, it is anticipated to yield solutions that are extremely close to this true minimum for any value of N.64  Note also that the SCE functional of eqn (33) is a natural counterpart to the KS non-interacting kinetic energy functional of eqn (2). This is because the SCE functional is obtained by minimizing the electron repulsion energy alone, while the KS functional is obtained by minimizing the kinetic energy alone. The SCE functional can be computed from:
(34)
where fi(r) are the co-motion functions, the objects that give the positions of N − 1 electrons when one electron is at position r. Thus, in the SCE state, if a reference electron is located at the position r, the remaining electrons are exclusively found at the positions defined by fi(r) and nowhere else. Detailed discussions on the mathematical properties of co-motion functions, their construction methods, and associated challenges can be found in ref. 64.
What matters most for us in the context of the large λ limit of density is the end point of the global AC integrand, W[n]. In terms of the SCE functional, it is given by:
(35)
Furthermore, w(r) that we used several times in this chapter and showed in Fig. 2, can be computed in terms of the co-motion function from:64,67 
(36)
where vH(r) is the Hartree potential.

The construction of co-motion functions necessary for SCE solutions has introduced certain integrals of the density, which subsequently offer unique ingredients for the construction of the XC functional.64  These new ingredients are very different from the conventional ones used for building XC approximations (i.e., semilocal ingredients and KS orbitals). Indeed, these novel definitions have been leveraged to develop new functionals that approximate both, λ → ∞ and physical regimes. All these new functionals strive to tackle the strong correlation problem in DFT. In the years and decades ahead, it will be interesting to see to what extent these new SCE-based mathematical features can address the strong correlation issue and to construct the next generation of DFT methods. Particularly, the new ingredients that emerge in the SCE limit could be used as novel features for machine learning the XC functional (Fig. 3).

Fig. 3

DFT and MP AC (correlation) adiabatic connection curves for the hydride ion (H). The data have been taken from ref. 68.

Fig. 3

DFT and MP AC (correlation) adiabatic connection curves for the hydride ion (H). The data have been taken from ref. 68.

Close modal
The DFT AC framework introduced in this chapter defines the exchange–correlation energy within KS DFT. However, the correlation energy is in a traditional quantum chemical sense defined as the difference between the true energy and Hartree–Fock (HF) energy. Using the AC formalism, the Møller–Plesset adiabatic connection (MPAC),69–72  was introduced where a different non-interacting system, HF system, is coupled to the full interacting system,
(37)
which has the standard Coulomb J ˆ = J ˆ [ n HF ] and exchange operators K ˆ = K ˆ [ { ϕ i HF } ] . Using the Hellmann–Feynman theorem, we find the following definition of the correlation energy,
where W c , λ HF is the HF-based adiabatic connection integrand,
(38)
The small λ expansion of this AC is now the standard Møller–Plesset pertubation theory,
(39)

Although the basic principle behind the two adiabatic connections is the same, there are some important differences. Firstly, the density constrain that characterizes the DFT AC is absent in the MP AC, which recovers the HF density, nHF, at λ = 0 and the physical density, n, at λ = 1. Naively, one would expect the electrons to be unbounded, however the scaling with λ of the Hartree potential ( J ˆ [ n HF ] = v H [ n HF ] ) creates an increasingly negative energy well which binds the electrons.73 

The strong interacting limit was uncovered as,71 
(40)
(41)
(42)
(43)
where the electrostatic energy, Eel, can be understood as a Wigner crystal of N point charges bound by −vH,
(44)
and is linked to W via the following inequality,70 
(45)
The subleading order terms are accurate variational estimates both of which are also functionals of nHF and depend respectively on the minimizing positions, {r1rN}, and the minimizing positions at the nuclei, r Z k .71  Notice that the exchange operator, although scaled by λ, only contributes at subleading orders.70,73  This directly effects the subleading order terms by adding a spin dependence,72  which differs from the DFT AC where the spin only comes in at e λ .74  Gradient expansion approximations have recently also been derived for the first two orders73  based on an analysis coming from particle number scaling that study the large Z limit.40,75–77 

Similar to the functionals of Section 4, interpolation schemes have been derived for the MP AC.78,79  The main difference is that the MP AC is not convex, because MP2 (only the first term of eqn (20)) underestimates the correlation energy forcing a concave curve for small λ, whereas at larger λ the curve was proven to be convex. To circumvent this issue, two SPL functionals,49,78,79  called SPL2, were combined resulting in a functional with more freedom to approximate this feature correctly. To mimic the effect of the large corrections from subleading orders, two fitting parameters were added to the inequality of eqn (45). Afterwards, the point-charge-plus-continuum model36  is used to approximate W. Other interpolation schemes have also been introduced, but we will focus on SPL2.

Recently,79  cost saving strategies and more accurate MP2 forms were used to not only decrease the cost but also increase the accuracy. In principle we are only allowed to use the standard MP2 correlation energy as the first order approximation, however it was shown that other forms of MP2 can improve on the previous iterations for a variety of different systems. As an example, Fig. 4 shows the dissocation curves of the benzene dimer for CCSD(T) (black), a SPL2 form (orange), MP2 (blue), and B2PLYP-D3 (red), where it shows that SPL2 is the only curve that improves MP2 towards CCSD(T) whereas the hybrid functional is significantly less accurate. Especially for halogen bonded complexes with strong charge transfer natures, the new SPL2 functional can improve by a factor of 3 to 5 on dispersion corrected hybrids and double hybrids, reaching a near CCSD(T) accuracy for a large range of non-covalant interactions.

Fig. 4

The dissociation curves of benzene (π-stacked) dimer obtained from MP2, SPL2, B2PLYP-D3 and CCSD(T) as a reference. R/Re denotes the distance between the benzene fragments in a dimer relative to their equilibrium distance.

Fig. 4

The dissociation curves of benzene (π-stacked) dimer obtained from MP2, SPL2, B2PLYP-D3 and CCSD(T) as a reference. R/Re denotes the distance between the benzene fragments in a dimer relative to their equilibrium distance.

Close modal

With the help of very accurate wave function methods, the characteristic of the global AC curves for small systems is attainable.80,81  A similar result is obtained for the local AC functions.15,82  These observations are useful for assessing the accuracy of new approximation models for the XC functional. In addition to the density fixed AC reviewed here, several other alternatives have been proposed: the AC based on the range-separation of electronic interaction,83  the kinetic-energy based adiabatic connection84–86  and adiabatic connections for strictly correlated electrons.87,88 

1
N.E.R.S.C. Center
, NERSC Computational Resources Report, 2022.
2
Schmidt
J.
,
Benavides-Riveros
C. L.
,
Marques
M. A. L.
,
J. Phys. Chem. Lett.
,
2019
, vol.
10
pg.
6425
3
Dick
S.
,
Fernandez-Serra
M.
,
Nat. Commun.
,
2020
, vol.
11
pg.
3509
4
Nagai
R.
,
Akashi
R.
,
Sugino
O.
,
npj Comput. Mater.
,
2020
, vol.
6
pg.
43
5
Bogojeski
M.
,
Vogt-Maranto
L.
,
Tuckerman
M. E.
,
Müller
K. R.
,
Burke
K.
,
Nat. Commun.
,
2020
, vol.
11
pg.
5223
6
Kalita
B.
,
Li
L.
,
McCarty
R. J.
,
Burke
K.
,
Acc. Chem. Res.
,
2021
, vol.
54
pg.
818
7
Margraf
J. T.
,
Reuter
K.
,
Nat. Commun.
,
2021
, vol.
12
pg.
344
8
Pederson
R.
,
Kalita
B.
,
Burke
K.
,
Nat. Rev. Phys.
,
2022
, vol.
4
pg.
357
9
Huang
B.
,
von Rudorff
G. F.
,
von Lilienfeld
O. A.
,
Science
,
2023
, vol.
381
pg.
170
10
Ernzerhof
M.
,
Chem. Phys. Lett.
,
1996
, vol.
263
pg.
499
11
Levy
M.
,
March
N. H.
,
Handy
N. C.
,
J. Chem. Phys.
,
1996
, vol.
104
5
pg.
1989
12
Burke
K.
,
Ernzerhof
M.
,
Perdew
J. P.
,
Chem. Phys. Lett.
,
1997
, vol.
265
pg.
115
13
Becke
A. D.
,
J. Chem. Phys.
,
1993
, vol.
98
pg.
5648
14
Mori-Sanchez
P.
,
Cohen
A. J.
,
Yang
W. T.
,
J. Chem. Phys.
,
2006
, vol.
124
pg.
091102
15
Vuckovic
S.
,
Irons
T. J. P.
,
Savin
A.
,
Teale
A. M.
,
Gori-Giorgi
P.
,
J. Chem. Theory Comput.
,
2016
, vol.
12
6
pg.
2598
16
Perdew
J. P.
,
Ernzerhof
M.
,
Burke
K.
,
J. Chem. Phys.
,
1996
, vol.
105
22
pg.
9982
17
Görling
A.
,
Levy
M.
,
J. Chem. Phys.
,
1997
, vol.
106
7
pg.
2675
18
Grimme
S.
,
J. Chem. Phys.
,
2006
, vol.
124
3
pg.
034108
19
Zhang
Y.
,
Xu
X.
,
Goddard
W. A.
,
Proc. Natl. Acad. Sci. U. S. A.
,
2009
, vol.
106
13
pg.
4963
20
Sharkas
K.
,
Toulouse
J.
,
Savin
A.
,
J. Chem. Phys.
,
2011
, vol.
134
pg.
064113
21
Langreth
D. C.
,
Perdew
J. P.
,
Solid State Commun.
,
1975
, vol.
17
pg.
1425
22
Gunnarsson
O.
,
Lundqvist
B. I.
,
Phys. Rev. B: Condens. Matter Mater. Phys.
,
1976
, vol.
13
pg.
4274
23
Perdew
J. P.
,
Langreth
D. C.
,
Phys. Rev. B: Condens. Matter Mater. Phys.
,
1977
, vol.
15
pg.
2884
24
Levy
M.
,
Proc. Natl. Acad. Sci. U. S. A.
,
1979
, vol.
76
12
pg.
6062
25
Levy
M.
,
Perdew
J. P.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1985
, vol.
32
4
pg.
2010
26
Teale
A. M.
,
Coriani
S.
,
Helgaker
T.
,
J. Chem. Phys.
,
2010
, vol.
132
pg.
164115
27
Levy
M.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1991
, vol.
43
pg.
4637
28
Perdew
J. P.
,
Burke
K.
,
Ernzerhof
M.
,
Phys. Rev. Lett.
,
1996
, vol.
77
pg.
3865
29
Fuchs
M.
,
Niquet
Y. M.
,
Gonze
X.
,
Burke
K.
,
J. Chem. Phys.
,
2005
, vol.
122
pg.
094116
30
Vuckovic
S.
,
Irons
T. J. P.
,
Wagner
L. O.
,
Teale
A. M.
,
Gori-Giorgi
P.
,
Phys. Chem. Chem. Phys.
,
2017
, vol.
19
pg.
6169
31
Levy
M.
,
Perdew
J. P.
,
Phys. Rev. B: Condens. Matter Mater. Phys.
,
1993
, vol.
48
pg.
11638
32
Görling
A.
,
Levy
M.
,
Phys. Rev. B: Condens. Matter Mater. Phys.
,
1993
, vol.
47
pg.
13105
33
Görling
A.
,
Levy
M.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1994
, vol.
50
pg.
196
34
Gori-Giorgi
P.
,
Vignale
G.
,
Seidl
M.
,
J. Chem. Theory Comput.
,
2009
, vol.
5
4
pg.
743
35
Liu
Z. F.
,
Burke
K.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
2009
, vol.
79
6
pg.
064503
36
Seidl
M.
,
Perdew
J. P.
,
Kurth
S.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
2000
, vol.
62
pg.
012502
37
Lieb
E. H.
,
Oxford
S.
,
Int. J. Quantum Chem.
,
1981
, vol.
19
pg.
427
38
Perdew
J. P.
,
Ruzsinszky
A.
,
Sun
J.
,
Burke
K.
,
J. Chem. Phys.
,
2014
, vol.
140
18
pg.
18A533
39
Seidl
M.
,
Benyahia
T.
,
Kooi
D. P.
and
Gori-Giorgi
P.
, 2022, arXiv:2202.10800.
40
Daas
K. J.
,
Kooi
D. P.
,
Benyahia
T.
,
Seidl
M.
,
Gori-Giorgi
P.
,
J. Chem. Phys.
,
2023
, vol.
159
23
pg.
234114
41
Lewin
M.
,
Lieb
E. H.
,
Seiringer
R.
,
Phys. Rev. B
,
2019
, vol.
100
3
pg.
035127
42
Cotar
C.
and
Petrache
M.
, arXiv, 2017, preprint, arxiv:1707.07664,
43
Lewin
M.
,
Lieb
E. H.
,
Seiringer
R.
,
Lett. Math. Phys.
,
2022
, vol.
112
5
pg.
92
44
Räsänen
E.
,
Seidl
M.
,
Gori-Giorgi
P.
,
Phys. Rev. B: Condens. Matter Mater. Phys.
,
2011
, vol.
83
pg.
195111
45
Seidl
M.
,
Vuckovic
S.
,
Gori-Giorgi
P.
,
Mol. Phys.
,
2016
, vol.
114
pg.
1076
46
Vuckovic
S.
,
Levy
M.
,
Gori-Giorgi
P.
,
J. Chem. Phys.
,
2017
, vol.
147
21
pg.
214107
47
Becke
A. D.
,
J. Chem. Phys.
,
1993
, vol.
98
pg.
1372
48
Mori-Sánchez
P.
,
Cohen
A. J.
,
Yang
W.
,
J. Chem. Phys.
,
2006
, vol.
125
pg.
201102
49
Seidl
M.
,
Perdew
J. P.
,
Levy
M.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1999
, vol.
59
1
pg.
51
50
Seidl
M.
,
Perdew
J. P.
,
Kurth
S.
,
Phys. Rev. Lett.
,
2000
, vol.
84
22
pg.
5070
51
Vuckovic
S.
,
Gori-Giorgi
P.
,
Della Sala
F.
,
Fabiano
E.
,
J. Phys. Chem. Lett.
,
2018
, vol.
9
11
pg.
3137
52
Fabiano
E.
,
Gori-Giorgi
P.
,
Seidl
M.
,
Della Sala
F.
,
J. Chem. Theory Comput.
,
2016
, vol.
12
10
pg.
4885
53
Giarrusso
S.
,
Gori-Giorgi
P.
,
Della Sala
F.
,
Fabiano
E.
,
J. Chem. Phys.
,
2018
, vol.
148
13
pg.
134106
54
Becke
A. D.
,
J. Chem. Phys.
,
2005
, vol.
122
pg.
064101
55
Perdew
J. P.
,
Staroverov
V. N.
,
Tao
J.
,
Scuseria
G. E.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
2008
, vol.
78
5
pg.
052513
56
Vuckovic
S.
,
Wagner
L. O.
,
Mirtschink
A.
,
Gori-Giorgi
P.
,
J. Chem. Theory Comput.
,
2015
, vol.
11
7
pg.
3153
57
Burke
K.
,
Cruz
F. G.
,
Lam
K. C.
,
J. Chem. Phys.
,
1998
, vol.
109
pg.
8161
58
Cruz
F. G.
,
Lam
K. C.
,
Burke
K.
,
J. Phys. Chem. A
,
1998
, vol.
102
25
pg.
4911
59
Kirkpatrick
J.
,
McMorrow
B.
,
Turban
D. H.
,
Gaunt
A. L.
,
Spencer
J. S.
,
Matthews
A. G.
,
Obika
A.
,
Thiry
L.
,
Fortunato
M.
,
Pfau
D.
et al.,
Science
,
2021
, vol.
374
6573
pg.
1385
60
Burke
K.
,
J. Chem. Phys.
,
2012
, vol.
136
15
pg.
150901
61
Cohen
A. J.
,
Mori-Sánchez
P.
,
Yang
W.
,
Chem. Rev.
,
2012
, vol.
112
1
pg.
289
62
Pribram-Jones
A.
,
Gross
D. A.
,
Burke
K.
,
Rev. Phys. Chem.
,
2015
, vol.
66
1
pg.
283
63
Sim
E.
,
Song
S.
,
Vuckovic
S.
,
Burke
K.
,
J. Am. Chem. Soc.
,
2022
, vol.
144
15
pg.
6625
64
Vuckovic
S.
,
Gerolin
A.
,
Daas
T. J.
,
Bahmann
H.
,
Friesecke
G.
,
Gori-Giorgi
P.
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
,
2023
, vol.
13
pg.
e1634
65
Seidl
M.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1999
, vol.
60
6
pg.
4387
66
Seidl
M.
,
Gori-Giorgi
P.
,
Savin
A.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
2007
, vol.
75
4
pg.
042511
67
Mirtschink
A.
,
Seidl
M.
,
Gori-Giorgi
P.
,
J. Chem. Theory Comput.
,
2012
, vol.
8
9
pg.
3097
68
Vuckovic
S.
,
Fabiano
E.
,
Gori-Giorgi
P.
,
Burke
K.
,
J. Chem. Theory Comput.
,
2020
, vol.
16
7
pg.
4141
69
Pernal
K.
,
Int. J. Quantum Chem.
,
2018
, vol.
118
1
pg.
e25462
70
Seidl
M.
,
Giarrusso
S.
,
Vuckovic
S.
,
Fabiano
E.
,
Gori-Giorgi
P.
,
J. Chem. Phys.
,
2018
, vol.
149
24
pg.
241101
71
Daas
T. J.
,
Grossi
J.
,
Vuckovic
S.
,
Musslimani
Z. H.
,
Kooi
D. P.
,
Seidl
M.
,
Giesbertz
K. J. H.
,
Gori-Giorgi
P.
,
J. Chem. Phys.
,
2020
, vol.
153
21
pg.
214112
72
Daas
K. J.
,
Klute
E.
,
Seidl
M.
,
Gori-Giorgi
P.
,
J. Phys. Chem. A
,
2024
, vol.
128
20
(pg.
4138
-
4149
)
73
Daas
T. J.
,
Kooi
D. P.
,
Grooteman
A. J. A. F.
,
Seidl
M.
,
Gori-Giorgi
P.
,
J. Chem. Theory Comput.
,
2022
, vol.
18
3
pg.
1584
74
Grossi
J.
,
Kooi
D. P.
,
Giesbertz
K. J. H.
,
Seidl
M.
,
Cohen
A. J.
,
Mori-Sánchez
P.
,
Gori-Giorgi
P.
,
J. Chem. Theory Comput.
,
2017
, vol.
13
12
pg.
6089
75
Perdew
J. P.
,
Constantin
L. A.
,
Sagvolden
E.
,
Burke
K.
,
Phys. Rev. Lett.
,
2006
, vol.
97
pg.
223002
76
Elliott
P.
,
Burke
K.
,
Can. J. Chem.
,
2009
, vol.
87
10
pg.
1485
77
Argaman
N.
,
Redd
J.
,
Cancio
A. C.
,
Burke
K.
,
Phys. Rev. Lett.
,
2022
, vol.
129
pg.
153001
78
Daas
T. J.
,
Fabiano
E.
,
Della Sala
F.
,
Gori-Giorgi
P.
,
Vuckovic
S.
,
J. Phys. Chem. Lett.
,
2021
, vol.
12
20
pg.
4867
79
Daas
K. J.
,
Kooi
D. P.
,
Peters
N. C.
,
Fabiano
E.
,
Sala
F. D.
,
Gori-Giorgi
P.
,
Vuckovic
S.
,
J. Phys. Chem. Lett.
,
2023
, vol.
14
38
pg.
8448
80
Colonna
F.
,
Savin
A.
,
J. Chem. Phys.
,
1999
, vol.
110
pg.
2828
81
Teale
A. M.
,
Coriani
S.
,
Helgaker
T.
,
J. Chem. Phys.
,
2009
, vol.
130
pg.
104111
82
Irons
T. J.
,
Teale
A. M.
,
Mol. Phys.
,
2015
, vol.
114
pg.
484
83
Gori-Giorgi
P.
,
Savin
A.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
2006
, vol.
73
pg.
032506
84
Savin
A.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1995
, vol.
52
pg.
R1805
85
Levy
M.
,
Görling
A.
,
Phys. Rev. A: At., Mol., Opt. Phys.
,
1995
, vol.
52
pg.
R1808
86
Teale
A. M.
,
Helgaker
T.
,
Savin
A.
,
J. Am. Chem. Soc.
,
2016
, vol.
63
1
pg.
121
87
Liu
Z. F.
,
Burke
K.
,
J. Chem. Phys.
,
2009
, vol.
131
pg.
124124
88
Gori-Giorgi
P.
,
Seidl
M.
,
Vignale
G.
,
Phys. Rev. Lett.
,
2009
, vol.
103
pg.
166402
Close Modal

or Create an Account

Close Modal
Close Modal