We show that the leading term in the strong-interaction limit of the adiabatic connection that has as weak-interaction expansion the Møller-Plesset perturbation theory can be fully determined from a functional of the Hartree-Fock density. We analyze this functional and highlight similarities and differences with the strong-interaction limit of the density-fixed adiabatic connection case of Kohn-Sham density functional theory.

The adiabatic connection formalism has always been a powerful theoretical tool to build approximations for the exchange-correlation (XC) energy in density functional theory (DFT), in particular, for hybrid1–3 and double-hybrid functionals,4–7 but also for other kinds of XC functionals.8–15 In the Kohn-Sham (KS) framework, the density fixed adiabatic connection can be defined via the λ-dependent hamiltonian16,17

(1)

where T^ is the kinetic energy operator for the N electrons, V^ee is their mutual Coulomb repulsion, and V^λ[ρ]=i=1Nvλ(ri,[ρ]) is the one body potential that makes the ground-state wavefunction of Eq. (1), ΨλDFT, yield the density ρ(r) ≡ ρλ=1(r) for all values of λ. From Eq. (1), one can derive an exact formula for the KS DFT XC energy16,17

(2)

where

(3)

with U[ρ] being the Hartree energy. The coupling-constant integrand of Eq. (3) has the known small-18 and large-λ expansions19 

(4)
(5)

where ExDFT is the exact KS exchange energy [the same expression as in Hartree-Fock (HF) theory, but using KS orbitals] and EcGLn is the nth term in the Görling-Levy perturbation series.18,20 The expansion for large λ of Eq. (5) has as leading term the functional WDFT[ρ], given by the minimum possible expectation value of the electron-electron repulsion in a given density ρ(r)21–23 

(6)

while the next leading term, determined by WDFT[ρ], corresponds to the potential energy of zero-point vibrations around the manifold corresponding to the support of the minimizing probability density in Eq. (6).19 While for the leading term there are rigorous proofs,22,23 this next term is only a very reasonable conjecture that has been confirmed numerically in simple cases where it was possible to compute the exact integrand WλDFT[ρ].24,25

Mixing KS DFT with Hartree-Fock (HF) ingredients is an approximation strategy that has a long history in chemistry, already starting with hybrids1–3,26–29 and double hybrids,4–7 but also by simply inserting the HF density into a given approximate XC density functional.30–35 Very recently, it has also been observed that rather accurate interaction energies,36,37 particularly for non-covalent complexes,38 can be obtained from models for WλDFT[ρ] that interpolate between the two limits of Eq. (4)—retaining only the first term, GL2, in the GL series—and of Eq. (5), using HF densities and orbitals as input, i.e., by constructing de facto an approximate resummation of the Møller-Plesset (MP) series, a procedure that lacks so far a theoretical justification. Motivated, in particular, by these last findings, we analyze in this communication the Hartree-Fock adiabatic connection [see Eq. (7)] whose Taylor expansion around λ = 0 is the MP series [see Eq. (10)] and show that the leading term in the λ expansion is determined by a functional of the HF density [see Eqs. (14) and (15)]. We also highlight similarities and differences with the DFT case, showing that the large λ expansion in HF theory has a structure similar to the one of Eq. (5).

We keep the notation general, as only few key properties of the HF operators are important here. We consider the adiabatic connection (see, e.g., Ref. 39)

(7)

with V^ext being the (nuclear) external potential and Ĵ = Ĵ[ρHF] and K^=K^[{ϕiHF}] being the standard HF Coulomb and exchange operators, which are fixed once for all in the initial HF calculation, and do not depend on λ, but only on the HF density ρHF and occupied HF orbitals {ϕiHF}. In the ground state ΨλHF of ĤλHF, the density ρλ(r) changes with λ: ρλ=0(r) is the HF density ρHF(r), and ρλ=1(r) is the exact physical density ρ(r). Note that Teale et al.40 have analyzed a related adiabatic connection, in which the external potential is kept fixed; in that framework in the limit λ, all the electrons but one escape to infinity. From Eq. (7), the Hellmann-Feynman theorem yields the exact formula

(8)

for the XC energy in the HF framework, with

(9)

Equation (9) has been defined to allow for a direct comparison with the DFT WλDFT[ρ] of Eqs. (2) and (3), with Wλ=0HF=ExHF, and for small λ

(10)

where EcMPn the nth term in the MP series. As is well-known (see, e.g., Refs. 41 and 42), the radius of convergence of the MP series is typically smaller than 1. Here we ask the question: What happens toΨλHFandWλHFas λ → ∞? After answering this theoretical question, we will discuss its actual relevance for constructing approximations.

When λ becomes very large, the term λ(V^eeĴK^) in Eq. (7) becomes more and more important, and we argue that the wavefunction ΨλHF should end up minimizing this term alone, similarly to the DFT case21 of Eq. (6). The difference here is that the minimizer is not constrained to yield a fixed density and the operator to be minimized also contains ĴK^. We further argue that the expectation value of K^ is subleading with respect to the one of V^eeĴ, i.e., we argue that

(11)

Before we shall support this conjecture with a variational argument, we discuss its consequences.

If Eq. (11) holds, then ΨλHF for λ ends up minimizing the even simpler operator λ(V^eeĴ)

(12)
(13)

The “asymptotic hamiltonian” H^HF=V^eeĴ[ρHF] is completely specified by the HF density ρHF(r), since N = ∫drρHF(r) and Ĵ[ρ]=i=1NvH(ri;[ρ]), with vH(r;[ρ])ρ(r)|rr|dr. Consequently, also the minimizer in Eqs. (12) and (13) is specified solely by ρHF

(14)

and the minimum in Eq. (13) is a functional of ρHF

(15)

The minimizer in Eq. (12) could be not unique, but this does not affect the value of the minimum, which is the object of the present investigation. The functional Eel[ρ]=minΨΨ|V^eeĴ[ρ]|Ψ+U[ρ] has a simple classical interpretation: Since H^HF=V^eeĴ[ρHF] is a purely multiplicative operator

(16)

the square modulus |ΨHF|2 of its minimizing wave function is a distribution in R3N that is zero wherever H^HF as a function of r1, …, rN does not assume its global minimum (if it were otherwise, it would not be optimal as we could always lower the energy by increasing the weight of the wave function in the global minimum of H^HF). In other words,

(17)

is the minimum total electrostatic energy of N equal classical point charges (−e) in a positive background with continuous charge density (+e)ρ(r). The term U[ρ], inherited from Eq. (9), represents the background-background repulsion.

Strictly speaking, the minimizer ΨHF is not in the space of allowed wavefunctions so that the minimum is actually an infimum, similarly to the DFT case.22,23,43

Equations (14) and (15) comprise a central result of this work: they show that the strong-interaction limit of the HF adiabatic connection can be determined from a functional of the HF density, providing some theoretical justification for resumming the MP series by using a DFT-like expansion at large λ with functionals of ρHF,36–38 although, as we will discuss, there are still several points to be addressed.

We now analyze the functionals ΨHF[ρ] and Eel[ρ], comparing them with the DFT case. As λ, ĤλDFT of Eq. (1) tends to λH^DFT[ρ], with21 

(18)

Comparing H^DFT[ρ] with H^HF[ρ] of Eq. (16), we see that both hamiltonians consist of the electron-electron repulsion operator and of an attractive one-body potential. In the HF case, the attractive potential is −vH(r, [ρ]), which is, for typical Hartree-Fock densities, strong enough to create a classical bound crystal. To be more precise, −vH(r, [ρ]) is more attractive than the one-body potential v(r, [ρ]). In fact, the potential v(r, [ρ]), which has been studied in several studies,21,44–46 is generated by a charge that integrates to N − 121,45

(19)

while the attractive potential −vH(r, [ρ]) is generated by the given density ρ(r), which integrates to N. For finite systems, the state ΨHF[ρ] is thus more compact than the state ΨDFT[ρ]: this is due to the density constraint in the DFT adiabatic connection, which forces ΨDFT[ρ] to have the given quantum mechanical density ρ(r).21,46

We note in passing that, for given occupied HF orbitals, we have the chain of inequalities

(20)

The first one, WHFEel[ρHF], is trivial since WHF=Eel[ρHF]+2ExHF and ExHF0. The second inequality holds for any density ρ(r), Eel[ρ]WDFT[ρ]. To prove it, we introduce the bifunctional W[ρ,v]

(21)

for which we have

(22)

and, from the dual formulation of WDFT[ρ],21,47,48

(23)

which clearly completes the proof.

As promised, we provide a variational argument to support the assumption of Eq. (11), sketching the main points and leaving a more detailed treatment to a subsequent paper. We start by considering the global minimum R̲min{r1min,,rNmin} of the function H^HF of Eq. (16) and construct the simple trial wavefunction

(24)

where Gα(r)=α3/4π3/4eα2|r|2, with α being a λ-dependent variational parameter that goes to infinity for large λ, α(λ) ∼ λq with q > 0. By construction, when α (i.e., when λ), we have that

(25)

where ΨHF was introduced in Eq. (12) (in the case of degeneracy, we can select one of the minimizers, since here we only want to obtain an upper bound to the lowest eigenvalue of HλHF). We now analyze, for large α, the expectation value on ΨλT of each term appearing in ĤλHF of Eq. (7), obtaining

(26)
(27)
(28)
(29)

where t, h, and k are all positive numbers. This is obvious for t, but it is also true for k because the expectation of K^ is positive for any wavefunction Ψ, as K^ has a negatively definite kernel. The fact that the expectation value of K^ on ΨλT vanishes as α−1 for large α is due to the non-locality of K^, which samples the gaussians in the bra and in the ket in different points of space, and due to the regularity properties of the HF orbitals (which have no delta-function singularities). The positivity of h in Eq. (27) can be proven by expanding H^HF around R̲min up to the second order, which gives a positive definite hessian matrix.

Putting together Eqs. (26)–(29) and replacing α with λq, we find that, for large λ, the expectation value of ĤλHF on ΨλT behaves asymptotically as

(30)

With t, h, and k being positive, we see that the best variational choice to make the next leading term after O(λ) increase with the lowest possible power of λ is q = 1/2, as conjectured in Eq. (11). Although ΨλT of Eq. (24) is not antisymmetric, we can always properly antisymmetrize it, which only leads to corrections O(eα) in the computation of the expectation values, similarly to the DFT case.25 

Thus, we have explicitly constructed a variational wavefunction that yields the minimum possible value for the leading term O(λ) in the expectation of ĤλHF. In fact, since Eel[ρHF] − U[ρHF] is the global minimum of the multiplicative operator V^eeĴ, there is no wavefunction that can yield a lower expectation for this operator. Moreover, since K^ is positive definite, the best we can do is to make its expectation zero when λ, which our wavefunction is able to do.

This variational argument also shows that the next leading term in WλHF should be order λ−1/2, similarly to the DFT case of Eq. (5). A quantitative estimate of this next leading term could be in principle obtained by using the normal modes around the minimum of V^eeĴ: a unitary transformation from the ririmin to the normal mode coordinates ξ1, , ξ3N that diagonalize the hessian of H^HF at R̲min leads to a set of uncoupled harmonic oscillators whose spring constants scale with λ

(31)

with ωα2 being the eigenvalues of the hessian of H^HF at R̲min. The ground-state of ĤλZP is obtained by occupying the lowest state of each oscillator, with the product state

(32)

This wavefunction should provide the minimum possible expectation, to order λ1/2, of T^+λ(V^eeĴ). However, since λK^ is of the same order λ1/2, we cannot exclude at this point that the minimization of the full T^+λ(V^eeĴK^) could lead to a different set of occupied oscillator states. This investigation will be the object of future studies. From our present treatment, we have so far

(33)

with

(34)
(35)

where WK,HF is due to the effect of λK^ at orders λ1/2 in ĤλHF and is a functional of the occupied HF orbitals. Equation (34) should be exact while Eq. (35) is for now a conjecture. We also see that both WHF and WHF have a part that is a functional of the HF density only and a part that is a functional of the occupied HF orbitals. In both cases, the part that is a density functional has an origin similar to the one of the DFT functionals of Eq. (5), being, respectively, a classical electrostatic energy and the potential energy of zero-point oscillations around a classical minimum. The parts that need the knowledge of the occupied HF orbitals do not appear in the DFT case. This structure should be exact, although the detailed form of WHF might include a different set of occupied oscillator states.

Although the λ limit of WλHF has a structure similar to the one of DFT, there are many differences that need to be kept in mind. Both WλDFT[ρ] and WλHF are decreasing functions of λ

(36)

but WλDFT[ρ] for λ ≥ 0 is believed to be convex or at least piecewise convex (if there are crossings of states), while WλHF is for sure not always convex. In fact, the MP2 correlation energy usually underestimates (in absolute value) the total correlation energy EcHF, implying that WλHF for 0 < λ ≪ 1 must run below its tangent; thus, WλHF usually starts concave for small λ and then needs to change convexity to tend to the finite asymptotic value WHF for large λ. Moreover, while the density constraint of the DFT adiabatic connection usually mitigates the crossing of states, the HF adiabatic connection might have jumps or kinks as λ is increased. A simple example is the N = 1 case, for which WλHF=U[ρHF] for 0 ≤ λ ≤ 1, while for λ > 1, the curve starts to decrease, tending, as λ to a well defined value, with the electrostatic energy determined by the configuration in which the electron is sitting in the minimum of −vH(r, [ρ]).

In conclusion, we have shown that by looking at the λ limit of the HF adiabatic connection, we recover functionals of the HF density, revealing a new intriguing formal link between HF and DFT. However, we should also stress that the use of models for WλHF taken from DFT, although somehow justified by our analysis, should at this stage still be taken with some caution. The empirical observation so far37,38 is that these models are not accurate for total energies, but work rather well for interaction energies, with a small variance, particularly for non-covalent complexes.37,38 This point requires further investigation, which will be the object of a paper in preparation, where WλHF will be computed and analyzed for various systems, and will be compared against various models. We will also evaluate and further analyze WHF and WHF, making a detailed numerical comparison with the corresponding DFT functionals. We can already remark that the difference between WHF and WDFT can be big. For example, for the He atom, we have WHF4.347 Ha, while WDFT1.50 Ha. We will also study whether it is possible to extract a model for the self-energy in the strong-coupling limit, to be used in the context of Green’s function approaches,49–52 and to analyze in the same spirit adiabatic connections appearing in other theories.39,53

This work was supported by the European Research Council under H2020/ERC Consolidator Grant corr-DFT (Grant No. 648932). S.V. acknowledges financial support from NWO through Rubicon (Grant No. 019.181EN.026).

1.
A. D.
Becke
,
J. Chem. Phys.
98
,
1372
(
1993
).
2.
A. D.
Becke
,
J. Chem. Phys.
98
,
5648
(
1993
).
3.
J. P.
Perdew
,
M.
Ernzerhof
, and
K.
Burke
,
J. Chem. Phys.
105
,
9982
(
1996
).
4.
K.
Sharkas
,
J.
Toulouse
, and
A.
Savin
,
J. Chem. Phys.
134
,
064113
(
2011
).
5.
S.
Grimme
,
J. Chem. Phys.
124
,
034108
(
2006
).
6.
L.
Goerigk
and
S.
Grimme
,
J. Chem. Theory Comput.
7
,
291
(
2010
).
7.
N. Q.
Su
and
X.
Xu
,
J. Chem. Phys.
140
,
18A512
(
2014
).
8.
M.
Ernzerhof
,
Chem. Phys. Lett.
263
,
499
(
1996
).
9.
M.
Seidl
,
J. P.
Perdew
, and
S.
Kurth
,
Phys. Rev. Lett.
84
,
5070
(
2000
).
10.
P.
Mori-Sanchez
,
A. J.
Cohen
, and
W. T.
Yang
,
J. Chem. Phys.
125
,
201102
(
2006
).
11.
A. D.
Becke
,
J. Chem. Phys.
138
,
074109
(
2013
).
12.
S.
Vuckovic
,
T.
Irons
,
A.
Savin
,
A. M.
Teale
, and
P.
Gori-Giorgi
,
J. Chem. Theory Comput.
12
,
2598
(
2016
).
13.
S.
Vuckovic
,
T. J. P.
Irons
,
L. O.
Wagner
,
A. M.
Teale
, and
P.
Gori-Giorgi
,
Phys. Chem. Chem. Phys.
19
,
6169
(
2017
).
14.
H.
Bahmann
,
Y.
Zhou
, and
M.
Ernzerhof
,
J. Chem. Phys.
145
,
124104
(
2016
).
15.
S.
Vuckovic
and
P.
Gori-Giorgi
,
J. Phys. Chem. Lett.
8
,
2799
(
2017
).
16.
D. C.
Langreth
and
J. P.
Perdew
,
Solid State Commun.
17
,
1425
(
1975
).
17.
O.
Gunnarsson
and
B. I.
Lundqvist
,
Phys. Rev. B
13
,
4274
(
1976
).
18.
A.
Görling
and
M.
Levy
,
Phys. Rev. B
47
,
13105
(
1993
).
19.
P.
Gori-Giorgi
,
G.
Vignale
, and
M.
Seidl
,
J. Chem. Theory Comput.
5
,
743
(
2009
).
20.
A.
Görling
and
M.
Levy
,
Phys. Rev. A
50
,
196
(
1994
).
21.
M.
Seidl
,
P.
Gori-Giorgi
, and
A.
Savin
,
Phys. Rev. A
75
,
042511/12
(
2007
).
23.
C.
Cotar
,
G.
Friesecke
, and
C.
Klüppelberg
,
Arch. Ration. Mech. Anal.
228
,
891
(
2018
).
24.
L.
Cort
,
D.
Karlsson
,
G.
Lani
, and
R.
van Leeuwen
,
Phys. Rev. A
95
,
042505
(
2017
).
25.
J.
Grossi
,
D. P.
Kooi
,
K. J.
Giesbertz
,
M.
Seidl
,
A. J.
Cohen
,
P.
Mori-Sánchez
, and
P.
Gori-Giorgi
,
J. Chem. Theory Comput.
13
,
6089
(
2017
).
26.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
,
J. Chem. Phys.
118
,
8207
(
2003
).
27.
Y.
Zhao
and
D. G.
Truhlar
,
Acc. Chem. Res.
41
,
157
(
2008
).
28.
J.
Jaramillo
,
G. E.
Scuseria
, and
M.
Ernzerhof
,
J. Chem. Phys.
118
,
1068
(
2003
).
29.
A. V.
Arbuznikov
and
M.
Kaupp
,
Chem. Phys. Lett.
440
,
160
(
2007
).
30.
P. M.
Gill
,
B. G.
Johnson
,
J. A.
Pople
, and
M. J.
Frisch
,
Int. J. Quantum Chem.
44
,
319
(
1992
).
31.
N.
Oliphant
and
R. J.
Bartlett
,
J. Chem. Phys.
100
,
6550
(
1994
).
32.
M.-C.
Kim
,
E.
Sim
, and
K.
Burke
,
J. Chem. Phys.
134
,
171103
(
2011
).
33.
M.-C.
Kim
,
E.
Sim
, and
K.
Burke
,
J. Chem. Phys.
140
,
18A528
(
2014
).
34.
M.-C.
Kim
,
E.
Sim
, and
K.
Burke
,
Phys. Rev. Lett.
111
,
073003
(
2013
).
35.
E.
Sim
,
S.
Song
, and
K.
Burke
,
J. Phys. Chem. Lett.
9
,
6385
(
2018
).
36.
E.
Fabiano
,
P.
Gori-Giorgi
,
M.
Seidl
, and
F.
Della Sala
,
J. Chem. Theory Comput.
12
,
4885
(
2016
).
37.
S.
Giarrusso
,
P.
Gori-Giorgi
,
F.
Della Sala
, and
E.
Fabiano
,
J. Chem. Phys.
148
,
134106
(
2018
).
38.
S.
Vuckovic
,
P.
Gori-Giorgi
,
F.
Della Sala
, and
E.
Fabiano
,
J. Phys. Chem. Lett.
9
,
3137
(
2018
).
39.
K.
Pernal
,
Int. J. Quantum Chem.
118
,
e25462
(
2018
).
40.
A. M.
Teale
,
S.
Coriani
, and
T.
Helgaker
,
J. Chem. Phys.
130
,
104111
(
2009
).
41.
J.
Olsen
,
O.
Christiansen
,
H.
Koch
, and
P.
Joergensen
,
J. Chem. Phys.
105
,
5082
(
1996
).
42.
B.
Forsberg
,
Z.
He
,
Y.
He
, and
D.
Cremer
,
Int. J. Quantum Chem.
76
,
306
(
2000
).
43.
C.
Cotar
,
G.
Friesecke
, and
C.
Klüppelberg
,
Comm. Pure Appl. Math.
66
,
548
(
2013
).
44.
45.
S.
Vuckovic
,
M.
Levy
, and
P.
Gori-Giorgi
,
J. Chem. Phys.
147
,
214107
(
2017
).
46.
S.
Giarrusso
,
S.
Vuckovic
, and
P.
Gori-Giorgi
,
J. Chem. Theory Comput.
14
,
4151
(
2018
).
47.
G.
Buttazzo
,
L.
De Pascale
, and
P.
Gori-Giorgi
,
Phys. Rev. A
85
,
062502
(
2012
).
48.
S.
Vuckovic
,
L. O.
Wagner
,
A.
Mirtschink
, and
P.
Gori-Giorgi
,
J. Chem. Theory Comput.
11
,
3153
(
2015
).
50.
J.
Schirmer
and
G.
Angonoa
,
J. Chem. Phys.
91
,
1754
(
1989
).
51.
W.
Tarantino
,
P.
Romaniello
,
J. A.
Berger
, and
L.
Reining
,
Phys. Rev. B
96
,
045124
(
2017
).
52.
P. F.
Loos
,
P.
Romaniello
, and
J. A.
Berger
,
J. Chem. Theory Comput.
14
,
3071
(
2018
).
53.