Although the concept of the uniform electron gas is essential to quantum physics, it has only been defined recently in a rigorous manner by Lewin, Lieb and Seiringer. We extend their approach to include the magnetic case, by which we mean that the vorticity of the gas is also held constant. Our definition involves the grand-canonical version of the universal functional introduced by Vignale and Rasolt in the context of current-density-functional theory. Besides establishing the existence of the thermodynamic limit, we derive an estimate on the kinetic energy functional that also gives a convenient answer to the (mixed) current-density representability problem.

Following Ref. 1, we view the uniform electron gas (UEG) as a continuous system of (interacting or not) electrons with prescribed constant density everywhere. In the non-interacting case, one can readily write down the ground state of the UEG via a fully occupied Fermi sphere. One can consider the classical UEG by neglecting the kinetic energy of the electrons.1 

In contrast, the Jellium is a continuous system of interacting electrons in a constant neutralizing potential.2 In the classical case, it was recently shown that the Jellium and the UEG energies are the same.3 In the quantum case, the problem is still open. We will solely focus on the UEG in this article.

Reference 1 pioneered the use of universal density functionals of density-functional theory (DFT) as a theoretical tool in quantum statistical mechanics, namely for the definition of the UEG. In a later article,4 the same authors established an even more elegant (and equivalent) definition of the UEG, and we will restrict ourselves to this latter approach.

When magnetic fields are present, the situation is much more delicate.5 Even the noninteracting UEG is more difficult to analyze in the magnetic case,6–8 although the ground state can be given exactly. The celebrated Laughlin wave function has also received attention from the mathematics community,9,10 which is a rather precise trial state (with almost constant density in the thermodynamic limit) for the Jellium in a constant magnetic field.

In the first half of this section, we will fix the particle number NN. For simplicity of notation, we disregard spin in this work. The N-particle fermionic Hilbert space is the antisymmetric power HN=NH (endowed with the usual inner product), where H=L2(Rd) is the one-particle Hilbert space. Note that H1=H.

Generally speaking, a density matrix (or state) on a Hilbert space is self-adjoint, non-negative trace class operator that has trace one. We first consider the Hilbert space to be the fermionic N-particle space HN. To make notations more compact, we henceforth restrict our attention to states with finite kinetic energy, which we call H1-regular states. The set of H1-regular (fermionic) N-particle states is defined as
where S1 the trace class. Then clearly 0⩽Γ = Γ and TrHNΓ=1 for all ΓDN. It is not hard to show that the operator Γ can be given by the Hermitian integral kernel Γ:RdN×RdNC (denoted with the same symbol),
so that its action on an element ΦHN is given by ΓΦ(X) = Γ(XY)Φ(Y) dY.
Let A:HNR be a nonnegative quadratic form with domain in HNH1(RdN), and let A:HNHN be its self-adjoint realization. Then for any ΓDN, we define
which is a nonnegative number or +∞. Choosing A = −Δ, i.e., A(Ψ)=Ψ2, we find that TrHN(ΔΓ)=j1μjΨj2< for any ΓDN.
Next, we consider states on Fock space. For convenience, we introduce the notation H0=C. The fermionic Fock space F is then the Hilbert space given by direct sum of the Hilbert spaces Hn, F=H0H1H2 endowed with the usual inner product. A Fock space state Γ is a positive self-adjoint, trace-class operator on F with unit trace: ΓS1(F), 0⩽Γ = Γ and TrFΓ=1. We will not need this generality, however, and it will suffice to consider states Γ which commute with the number operator. It is easy to show by induction that such states are block-diagonal:
where ΓnS1(Hn) is such that 0Γn=Γn. Note in particular that Γ0R+, and that the normalization constraint now reads n0TrFΓn=1. Again, we will further restrict attention to states with finite kinetic energy, which we called H1-regular states. All this leads to the definition
of the set of H1-regular block-diagonal Fock space states. Each Γn may be given by an integral kernel Γn(XnYn), like before (of course these will not be normalized to have unit trace, in general). Note also that by linearity, the one-, and the two-particle reduced density matrices of Γ=Γ0Γ1D are given by
We note in passing that all derived quantities, such as ρΓ, may be defined by linearity.
In this section, we introduce various quantities which may be defined using the one-particle density matrix. The one-particle reduced density matrix γS1(H) of ΓDN is defined as (N times) the partial trace of Γ via duality:
In symbols, this operation is sometimes written as γ=NTrHN1Γ. Note the choice of normalization, so that we have TrHγ=N. It is easy to see that γ can be given by the kernel
It is well-known that γ is a trace class operator on the one-particle Hilbert space H such that 0γ=γ1. By the spectral theorem γ has the decomposition γ = j⩾1λj|φj⟩⟨φj|, which in terms of its kernel can be written as
for some 0⩽λj⩽1 with j⩾1λj = N and {φj}j1H1(Rd)L2-orthonormal. It is clear that γ has finite kinetic energy, i.e., it is H1-regular: j⩾1λjφj2 < ∞.
The density ργL1(Rd) of the one-particle density matrix γ may also be defined via duality, i.e., by requiring that Rdργφ=TrH(γφ) for all φL(Rd). This gives
The kinetic energy TrH(Δγ)j1λjφj2< may be expressed using the kernel γ(x, y) as
Here, the integrand τγ(x)≔xyγ(x, y)|y=x is called the kinetic energy density, and more generally
the kinetic energy tensor. Clearly, 0⩽τγ(x) = τγ(x) and TrR3τγ(x)=τγ(x).
A further useful quantity is the complex current density ζγ:RdCd, which is given by
(1)
A straightforward calculation shows that
(2)
where jγp:RdRd is called the paramagnetic current density.

The following theorem is based on,11–13 the proof may be found toward the end of the paper.

Proposition II.1.

Suppose that γ is fermionic one-particle density matrix composed of sufficiently smooth orbitals. Then the following hold true.

  • |ργ(x)|2+|jγp(x)|2ργ(x)+1dργ(x)Dajγp(x)ργ(x)τγ(x) for all xRd.

  • |jγp(x)||ζγ(x)|τγ(x)1/2ργ(x)1/2 for all xRd, hence jγpL1(Rd)d.

Here, the symmetric-, and antisymmetric derivatives Ds and Da will be used at some occasions, given by
for any differentiable vector field u:RdRd and Du:RdRd×d is its derivative. Further, we denote by A: B = Tr(AB) the Frobenius inner product, and by |A|=A:A the Frobenius norm.
Let us consider the case d = 3. Then Dau=22|rotu| and part (i) improves the usual Hoffmann–Ostenhof bound14 as
(3)
for any H1-regular density matrix γ, where νγ=rotjγpργ is the (gauge-invariant, see below) vorticity. We call the terms on the r.h.s. the von Weizsäcker-, the gauge-, and the vorticity term, respectively. Note that (3) implies that for H1-regular density one-particle matrices γ, we have R3|ργ|2<, R3|jγp|2ργ< and R3ργνγ<, quantities which play important roles later. Finally, we note that for d = 3 the constant in front of vorticity term in (3) can be improved to 1, see Ref. 12.

Remark 1.

Part (ii) of Proposition II.1 implies that the support of jp is contained in the support of ρ, in particular, the quotients jpρ and |jp|2ρ are meaningful on suppρ. Moreover, if τ ≡ 0 in some ball, then by (i), ρ is constant and jp ≡ 0 on that ball.

We now discuss a very important tool in the study of magnetic systems. Given an N-particle state ΓDN, consider the gauge-transformed state
where g is called the gauge function. The following is immediate.

Proposition II.2.
Let γ be a fermionic one-particle density matrix. Then the gauge transformation of γ is given by
In particular, ργ̃=ργ. Furthermore, jγ̃p=jγpργg and
(4)

We did not specify what function space the gauge function g is coming from. In order for the r.h.s. of (4) to make sense, it is enough to assume that gLργ2(Rd)d, using the Cauchy–Schwarz inequality and (3). Here and onwards, Lρp stands for the ρ-weighted Lp space.

Note that the gauge term Rd|jp|2ρ transforms exactly like the kinetic energy (4). More precisely,
therefore the inequality (3) is “consistent” with a gauge transformation in the sense that all the terms involving g will cancel. We call such inequalities gauge-invariant.
Fix NN. Recall that the set of pure (or mixed) state N-representable densities may be described as
according to Refs. 15 and 16. For later purposes, we also introduce the set of Fock space-representable densities as
For density-paramagnetic current density pairs, the situation is more complicated. Consider the set
of all pure state N-representable density-current pairs, and
the set of mixed state N-representable density-current pairs. In this more general setting it is yet unknown whether JNpure is equal to JNmixed or not. Unfortunately, no explicit characterization of either JNpure or JNmixed exists up to date. We can specify a larger set by setting
(5)
so that by (3), JNpure,JNmixedJN. It can be shown that JN is also convex.17 We remark that the finiteness of the vorticity term in (3) may also be added to the definition of JN.

Whether any (ρ,jp)JN can be represented with a Slater determinant ΦHNH1(R3N) such that ρΦ = ρ and jΦp=jp (the determinantal density-current N-representability problem) is still not settled completely. The following result gives a partial answer, however.

Theorem II.3.

(Lieb–Schrader18). Let d = 3 and suppose that (ρ,jp)JN and set ν=rotjpρ. Then the following holds true.

  • (Zero vorticity) If ν = 0, then there exists a Slater determinant ΦHNH1(R3N), such that ρΦ = ρ, jΦp=jp and
    for some constant CN > 0.
  • If N⩾4, then there exists a Slater determinant ΦHN, such that ρΦ = ρ, jΦp=jp. Moreover, if there is a δ > 0 such that the growth conditions
    hold true, thenΦ‖ < . Here, we have set f(x)=(1+x12)1+δ2(1+x22)1+δ2(1+x32)1+δ2.

    Moreover, Ref. 18 also exhibits an example (ρ, jp) in the case N = 2 (and of course ν ≠ 0), when there is no C1-solution to the representability problem. Unfortunately, the restriction N⩾4 and the fact that no control over the kinetic energy is retained in (ii) renders the preceding result inapplicable to us.

The following theorem establishes mixed representability under different assumptions, now for general NN.

Theorem II.4.
(Tellgren–Kvaal–Helgaker19 ). Let d = 3 and suppose that (ρ,jp)JN and that
(6)
Then there exists a constant C > 0 and a fermionic one-particle density matrix γ such that ργ = ρ, jγp=jp and
(7)
Representability holds also true without the integer particle number constraint on the density, i.e., when RdρN, but the representing state will live in the Fock space F. Analogously to the N-particle case, we define the set of Fock space-representable density-current pairs as
Then the conditions of Theorem II.4 provide a smaller set, but we will use an improved version below.

In this section, we define the current-density functionals relevant to our study. These are well-known in CDFT, but we will also employ them in the definition of the uniform electron gas. In this section we restrict ourselves to the physically most relevant 3D case.

We begin by introducing the grand-canonical Vignale–Rasolt functional20 via
(8)
Omitting the Coulomb interaction in the above definition, we obtain the mixed-state kinetic energy functional T(ρ, jp). By definition, T and F is (jointly) convex on J. In our analysis we will consider the indirect energy
(9)
where for any finite measures μ1 and μ2, we define the Coulomb inner product via
and additionally, as an abuse of notation, we set D(μ)≔D(μμ) for the direct energy of μ. If the measures μj come from densities ρj, i.e., dμj = ρj dx (j = 1, 2), then we will use the notations D(ρ1, ρ2) and D(ρ).
For convenience, we introduce the gauge-corrected indirect energy
and similarly T̄(ρ,jp). We will also consider the exchange-correlation functional
which is more common in the physics and chemistry literature.
The reason for the introduction of the “gauge-corrected” quantities is the following. Note that under a gauge transformation jp̃=jpρg, the functional F transforms as
(10)
i.e., like the kinetic energy (see Proposition II.2). Analogous relations holds for T(ρ, jp) and E(ρ, jp). It follows that the gauge-corrected indirect-, and kinetic energy Ē(ρ,jp) and T̄(ρ,jp), and exchange-correlation functional Exc(ρ, jp) are all gauge-invariant.

By replacing the N-particle Hilbert space with the corresponding Fock space, the next result follows along similar lines as the one in Ref. 21.

Theorem II.5.

Let (ρ,jp)J. Then the infima in the definitions (8) of the Vignale–Rasolt functional F and the kinetic energy functional T are both attained.

The grand-canonical Levy–Lieb functional F(ρ) (no paramagnetic current density argument) and the kinetic energy functional T(ρ) is defined similarly to F(ρ, jp) and T(ρ, jp), only without the constraint on jΓp in the infimum.

Proposition II.6.
(Correspondence with the nonmagnetic case). For any ρI, the relations T(ρ, 0) = T(ρ) and F(ρ, 0) = F(ρ) hold true. In particular,
where E(ρ) = F(ρ) − D(ρ) and Exc(ρ) = E(ρ) − T(ρ).

Proof.

By convexity and time-reversal symmetry, we find F(ρ,0)12F(ρ,jp)+12F(ρ,jp)=F(ρ,jp). Hence, F(ρ)F(ρ,0)infjpF(ρ,jp)=F(ρ) as desired. The proof of the relation between the kinetic energy functionals is similar. □

This section is devoted to the precise definition of the various energetic quantities of the 3D uniform electron gas (UEG) with constant density and constant vorticity.

Recall that the vorticity of the ground state of the noninteracting UEG in a constant magnetic field B0 is ν0 = −B0. Physically, we can generate a constant vorticity ν0 in the noninteracting UEG by subjecting it to an external magnetic field −B0. Moreover, linear response theory suggests the same holds true to the first order in the interacting case5. Instead of referring to the external magnetic field, we will simply demand the UEG to have constant density and constant vorticity.

A function 0ηCc(R3) such that suppηB1, R3xη(x)dx=0 and R3η=1 will be called a regularization function. Note that this is a slightly more general definition than the one in Refs. 1 and 4, where it is required that a regularization function be radial. We also introduce the notation ηδ(x) = δ−3η(δ−1x).

Define the indirect gauge-corrected energy per volume of the (interacting) UEG at constant density ρ0R+ and constant vorticity ν0R3 inside a bounded domain ΩR3 smeared with η as
(11)
Note that rot12ν0×x=ν0 and that eΩ,η(0, ν0) = eΩ,η(0, 0) = 0. The smearing of the indicator 1Ω is necessary, since R3|1Ω|2= and this would render the energy E infinite [see (3)]. Recall that Ē contains a term of the form R3|jp|2ρ, which we call the gauge correction term, and it needs to be subtracted otherwise the energy per volume would diverge in the thermodynamic limit. In fact, the gauge correction term grows faster than the volume,
where CΩ,η,ν0>0 is a constant depending only on Ω, η and ν0. Hence, by (3), the energy per volume diverges as → ∞. Notice that replacing 12ρ0ν0×12ρ0ν0×+ρ0g in the definition (11) above would yield the same value for Ē due to the gauge transformation rule (10). In particular, one can choose the Landau gauge instead of the symmetric gauge 12ν0×x above without changing anything.
By replacing Ē with T̄ on the r.h.s. of (11), we can also define the kinetic energy per volume τΩ,η(ρ0, ν0). Finally, we define the exchange-correlation energy per volume
Note that, in this case the gauge correction terms cancel.
Following Fischer, we say that a sequence of domains {ΩN}NNR3 has a uniformly regular boundary if
for some C > 0. For instance, one may take ΩN = NΩ for some convex, bounded and open set.22 

The following theorem is the main result of the paper.

Theorem III.1.
(Uniform electron gas). Let ρ0R+, ν0R3 and let {ΩN}NNR3 be a sequence of bounded domains such thatN| → ∞ and it has a uniformly regular boundary. Let δN > 0 be any sequence such that δN/|ΩN|1/3 → 0 and δNN|1/3 → ∞. Then the following thermodynamic limit exists
and is independent of the choice of the sequencesN}, {δN} and the regularization function η. Analogous results hold true for τUEG(ρ0, ν0) and eUEGxc(ρ0,ν0).
We see using Corollary III.4 (see below) for the upper bound and the Lieb–Thirring and Lieb–Oxford inequalities for the lower bound (see Corollary IV.2 and Theorem IV.3 below),
Note that in eUEG(ρ0, ν0) only the magnitude of ν0 matters due to rotational invariance in the thermodynamic limit [see Lemma IV.5(ii)]. It is clear from Proposition II.6 that we recover the nonmagnetic case by putting ν0 = 0, i.e., eUEG(ρ0, 0) = eUEG(ρ0). Moreover, (3) and the Lieb–Oxford inequality implies that eUEG(ρ0,ν0)cρ0ν0cLOρ04/3, so eUEG(ρ0, ν0) → +∞ as ν0 → ∞ for every fixed ρ0 > 0.

Remark 2.

Unfortunately, we cannot offer a definition of the classical UEG in the magnetic case, akin to Ref. 1 in the nonmagnetic case. The reason is that jp uses the phase of the wavefunction in an essential way, and we have no obvious classical probabilistic interpretation of that.

In this section, we describe a technical tool that is needed in the Proof of Theorem III.1, but it is also of independent interest mainly because it improves previous results on the (mixed) representability problem for density-current density pairs (see Sec. II D).

If ργ and jγp comes from an H1-regular one-particle density matrix γ, we know from (3) that we necessarily have that
Note that if d = 3, the the condition is simply R3ργ|νγ|<, where νγ=rotjγpργ is the vorticity.

Using the method of Proof of Theorem 3 in Ref. 4, we obtain the following.

Theorem III.2.

Fix ρL1(Rd;R+) such that ρL2(Rd)d. Let vLρ2(Rd)d and suppose that v admits a decomposition v = g + w, where gLρ2(Rd), wLρ2(Rd)d, DwLρ1(Rd)d×d.

Then there exists a one-particle density matrix γ with finite kinetic energy, such that ργ = ρ, jγp=jpρv and
(12)
for any ϵ > 0, and for some universal constants κ1, κ2 > 0. Here, the (dimension-dependent) Thomas–Fermi constant cTF is given by cTFdd+24π2|B1|2/d and Cd=1+d34.

The inequality (12) is not gauge-invariant, to see this, take v(x) = (0, x1, 0) (Landau gauge) and w(x)=12e3×x (symmetric gauge), with g(x)=12x1x2 and some appropriate ρ. Then Dav = Daw, Dsv ≠ 0 but Dsw = 0. Clearly, w = v and g = 0 is also a possible choice. The lack of gauge invariance in (12) solely comes from the presence of the symmetric derivative Dsw, which we discuss below. The utility of the formulation of Theorem III.2 is that with a proper choice of gauge, we can hope to make Dsw vanish.

The condition (6) of Theorem II.4 is easily seen to imply DwLρ1(Rd)d×d (up to a gauge) via the Cauchy–Schwarz inequality. We also note that for v = 0, we exactly recover the statement (and proof) of Theorem 3 in Ref. 4, since that construction actually yields a state γ with jγp=0. Based on the kinetic energy of the ground state of the noninteracting electron gas, we conjecture that the optimal constant in 3D is C3 = 1.

The last term on the r.h.s. of (12) has a similar role as the von Weizsäcker term. It can be rewritten as
where we used the fact Dap = 0 for any C2 scalar field p:RdR, due to Young’s theorem. Hence
in the sense of norm equivalence. In 3D, the second term on the r.h.s. is proportional to R3ρ|ν|, which is gauge-invariant.

The first term on the r.h.s. of the preceding equation (with w = wγ)—which we call the strain term—is not necessarily finite for H1-regular γ. However, if we assume that γ is H2-regular (and the gauge function has D2gLρ1(Rd)d×d), then it is not hard to show that the strain term is finite. Recall that the eigenstates of the magnetic Schrödinger Hamiltonian are in H2. Although Theorem III.2 does not provide the complete solution to the mixed density-current N-representability problem (it also does not provide an H2-regular γ without further assumptions, sadly), it seems to provide weak enough assumptions that any practically relevant (ρ, jp) pair verifies it. That being said, it is our belief that the presence of the strain term is only an artifact of the method of the proof of both Theorem II.4 and Theorem III.2. It remains a challenge to devise a gauge-invariant analogue of (12).

Motivated by Theorem III.2, we define the smaller set J̃J, via

With this notation at hand, we can provide an upper bound on the kinetic energy functional T(ρ, jp).

Corollary III.3.

For any (ρ,jp)J̃, we have T(ρ, jp)⩽r.h.s. of (12).

Moreover, using the fact23 that for every one-particle density matrix γ there exists a quasi-free state Γ, such that γΓ = γ, we obtain the following.

Corollary III.4.

For any (ρ,jp)J̃, we have E(ρ, jp)⩽r.h.s. of (12).

In fact, for a quasi-free state Γ, the Coulomb energy of Γ minus D(ρΓ) is always nonpositive.

The rest of the paper is the devoted to proofs. Since Ref. 4 explains technical details in a rather complete manner, we will avoid repetition whenever possible.

  • Step 0. First, define the ground-state one-particle density matrix ft(xy) of the noninteracting uniform electron gas of density t⩾0, whose translation-invariant kernel is given by
    It is immediate that ft is Hermitian and 0ft1 in the sense of operators. Also, its density is ft(0) = t. Moreover, xft(xy)|y=x = yft(xy)|x=y = 0. Furthermore, the kinetic energy density is also constant
  • Step 1. As the first derivative of ft vanishes on the diagonal, this state has zero paramagnetic current density. We would like to construct a state ftu(xy) which has a constant paramagnetic current density proportional to a fixed vector uRd. To achieve this, we simply shift the Fermi sphere |k|2d+2dcTFt2/d by u. More precisely, we consider the translation-invariant kernel ftu(xy), where
    Clearly, ftu is Hermitian and 0ftu1 in the sense of operators. Its density is ftu(0)=t, but now
    (13)
    Therefore, this state has paramagnetic current density
    as desired. Further, the kinetic energy density is again constant, but has an additional term
  • Step 2. We are now ready to construct a fermionic one-particle density matrix γ with prescribed density ρ(x) and paramagnetic current density of the form jp = ρv. Let us first assume that v = w. Following,4 we define24 
    where the weight function η:R+R+ is smooth, and satisfies
    and θ:RdR+ is also smooth and satisfies
    (14)
    These will be determined later in the proof. Then we have
    and
    Furthermore, we find
    as required. Next, we compute the kinetic energy density,
    since the sum of the first order derivatives of ftu cancel on the diagonal, see (13). Here, using the first two relations of (14), we have
    Moreover,
    The last term vanishes because Rdθ(u)du=0, by hypothesis. In summary,
    To optimize η, we simply follow Ref. 4 because there are no mixed terms containing both η and θ. The last term can be bounded by
    Clearly, it is enough to consider radial θ(u) = θ(r) for the optimization of the prefactors involving θ, where r = |u|. After switching to spherical coordinates, we obtain that with
    the bound (∗)⩽δρ(x) + C(δ)ρ(x)|Dv(x)|2 holds true. The function
    (15)
    is admissible, and has
    Hence,
    with the choice δ = |Dv(x)| and Cd=1+d34.
    In summary, we end up with
    (16)
    which completes the proof for v = w.
  • Step 3. For the general case v = −g + w, we apply the gauge transformation g on γ to get γ̃. Proposition II.2 together with (16) (where v is replaced by w) gives
    which finishes the Proof of Theorem III.2.

We start by recalling a simple version of the magnetic Lieb–Thirring inequality, which can be used to derive a gauge invariant bound. Let ALloc2(R3)3 and denote the corresponding magnetic Sobolev space as HA1(R3), see Ref. 25.

Theorem IV.1
(Magnetic Lieb–Thirring kinetic energy inequality).26  Fix ALloc2(R3). For any fermionic one-particle reduced density matrix γ in HA1(R3), the bound
holds true for some universal constant CLT > 0.

Using this, the usual Lieb–Thirring kinetic energy inequality can be cast in a gauge-invariant form.

Corollary IV.2
(Gauge-invariant Lieb–Thirring inequality). For any fermionic one-particle reduced density matrix γ with finite kinetic energy, the bound
holds true for some universal constant cLT > 0.

Proof.
For any ϵ > 0, we set Aϵ=jγpργ+ϵργ. Then AϵLloc2(R3) since R3|jγp|2ργ<. Note that
An application of Fatou’s lemma finishes the proof. □
For any ΓD let us introduce the notations
for the kinetic-, and the Coulomb energy of the state Γ. Of course, T(Γ)=Tr(ΔγΓ), so Corollary IV.2 provides a gauge-invariant lower bound on T(Γ). The following famous estimate on the indirect Coulomb energy will be used repeatedly in the sequel (which is clearly gauge-invariant).

Theorem IV.3.
(Lieb–Oxford inequality). 27–30  There exists a universal constant cLO > 0 such that for any Fock space density matrix Γ (not necessarily fermionic or H1-regular), the bound
holds true.

In this section, we discuss the behavior of the density functionals under various transformations of the density-current density pair.

First, we consider general affine transformations of states. Let Γ=Γ0Γ1D and for any MR3×3 nonsingular and aR3, set T(x) = Mx + a, and define the transformed state ϒTΓ by setting
Then the reduced one-particle density matrix of ϒTΓ is given by γϒTΓ(x,y)=(detM)γΓ(T(x),T(y)), where γΓ denotes the one-particle density matrix of Γ. We have ρϒTΓ(x)=(detM)ρΓ(T(x)) and jϒTΓp(x)=(detM)MjΓp(T(x)). Further, γϒTΓ(2)(x,y;x,y)=(detM)2γΓ(2)(T(x),T(y);T(x),T(y)). We immediately obtain that for any ΓD the following relations hold true:
(17)
(18)
Motivated by the above considerations, for any (ρ,jp)J, we define
(19)
as a slight abuse of notation. Using (17) and (18), and introducing the the gauge-corrected kinetic energy tensor
the transformed gauge-corrected indirect energy may be expressed as
(20)
which follows by a simple reparametrization of the infimum, using the fact that the map Γ↦ϒMΓ is invertible.

In ordinary DFT, the Levy–Lieb functional is invariant under isometries. In CDFT, we have the following transformation rule.

Lemma IV.4.
Let T(x) = Rx + a be an isometry for some R ∈ O(3) and aR3. Then
and

The proof follows from relations analogous to (20). As an application, we deduce how the indirect energy E and eΩ,ηδ(ρ0,ν0), τΩ,ηδ(ρ0,ν0) and eΩ,ηδxc(ρ0,ν0) transform under isometries. The proof uses the gauge transformation rule (10) in an essential way.

Lemma IV.5.

Fix ρ0R+ and ν0R3. Let T(x) = Rx + a be an isometry for some R ∈ O(3) and aR3. Let ΩR3 be a bounded domain with barycenter 0, and let η be a regularization function.

  • The following formula holds true:
    Further, the same relation holds with E replaced by T.
  • The relation eT1(Ω),ηδR(ρ0,ν0)=eΩ,ηδ(ρ0,Rν0) holds true. Moreover, the formula is also valid if e is replaced by τ and exc.

Proof.
First, we show (i). Clearly, T−1(x) = R(xa). Also, D(((1Ωηδ)T)ρ0)=D((1Ωηδ)ρ0). We may write the l.h.s. using Lemma IV.4 as
where we used the identity R(u ×Rv) = Ru ×v. Using the gauge transformation rule ,(10) we obtain
as stated. Here, the cross term disappeared because R3x(1Ωηδ)(x)dx=0, since the smeared set also has barycenter 0, due to R3xη(x)dx=0.
For part (ii), we may write based on part (i) that
where the mixed term in last term vanished as before. □
Choosing M=λ1/31 in (20) and setting
we deduce the following (the lower bound uses the Lieb–Oxford inequality as well).

Theorem IV.6.
For every (ρ,jp)J and 0⩽λ⩽1, we have
(21)
and
(22)

Let C1(12,12)3 denote the origin-centered unit cube, and Δj = TjΔ (j = 1, …, 24) be its tiling with 24 congruent tetrahedra, where Tjx = Rjxzj for some Rj ∈ SO(3) and zjC1. Here, Δ is a “reference” tetrahedron with barycenter 0. Then with C = ℓC1, we write
where > 0 the scale of the tiling. The reason behind the use of this particular tetrahedral tiling is that the Graf–Schenker inequality is employed for the Coulomb interaction, see Refs. 1 and 4 and the original31,32 for details.
The corresponding partition of unity is given by
(23)
The lower bound is based on well-known localization techniques, and for that, smoothing is needed. Let 0ηCc(R3) be such that R3xη(x)dx=0 with suppηB1 and R3η=1. Define
Then suppηδBδ10 and R3ηδ=1. The functions 1Δjηδ form a partition of unity in the following sense:
by (23). The following facts will be used repeatedly. Fix xR3, if suppηδ(x − ·) ∩ (Δj) = ∅, then clearly either supp ηδ(x − ·) ⊂ Δj or else suppηδ(x)(Δj)c. In the former case, (1Δjηδ)(x)=1 and in the latter, (1Δjηδ)(x)=0. Since suppηδ(x)Bδ10(x), we obtain that
(24)
It follows that 1Δjηδ1 on Δj\((Δj)+Bδ10). More conveniently, 1Δjηδ1 on (δj, as this set is clearly contained in the former.

The following trivial observation shows that jp behaves the same way as ρ with respect to localization. Let γ be a one-particle density matrix and fix 0⩽χ⩽1, where χC1(R3), and define γ|χ = χγχ. Then ργ|χ=χ2ργ and jγ|χp=χ2jγp, where we used the fact that χ real-valued.

This implies that we may proceed exactly as in Refs. 1 and 4. We do not, however, discard the kinetic energy term CδT() for the bound on E(ρ, jp), as it will be required to cancel the gauge terms in Theorem IV.10. Consequently, we have

Theorem IV.7.
There is a constant C > 0 such that for any δ > 0 such that 0 < δℓ < 1/C and (ρ,jp)J̃ the bound
holds. Furthermore, for any δ > 0 and > 0 the bound
also holds true.

First, (23) is regularized as follows. For any δ2, consider the functions (1δ/)31Tj(1δ)Δηδ. Here, Tj(1δ)Δ=(δ)RjΔ+zj and
To obtain a true partition of unity, these functions are averaged over the cubic tiling:
(25)
The following result says that the (grand-canonical) Vignale–Rasolt functional decouples the same way as the Levy–Lieb functional.

Proposition IV.8.
For any (ρ,jp)J̃, the bounds
and
hold true.

Proof.
The partition of unity (25) allows us to write our density-current pair as
(26)
where we have set
for any τC, zZ3 and j = 1, …, 24. We now claim that the density-current pairs (ρτ,z,j,jτ,z,jp) are representable, i.e., (ρτ,z,j,jτ,z,jp)J̃. In fact, it is clear that ρτ,z,jI, and with v=jpρ and
there holds R3ρτ,z,j|vτ,z,j|2R3ρ|v|2<. Also, R3ρτ,z,j|Dvτ,z,j|R3ρ|Dv|<.

The rest of the proof is analogous to that of Proposition 1 in Ref. 4. We note that in the construction of the trial state Theorem II.5 is used. □

Next, we state a similar bound for the indirect energy defined in 9.

Theorem IV.9.
For any (ρ,jp)J̃, 0<δ<2 and 0<α12, the bound
holds for a universal constant C. Here, 1α1+αdtt41α1+αdtt411α1+αdtt4.

Proof.
We have
(27)
According to the Proof of Proposition 1 in Ref. 4 (and Lemma 1 ibid.), replacing tℓ, δ and ρρR (and jpR(jpR)), and averaging, the preceding quantity may be bounded as follows
for 0 < δ < /2. By applying the same replacements and averagings in (27), we may use the preceding estimate and we end up with
Recalling Lemma IV.4 finishes the proof. □

We now prove that the thermodynamic limit for the indirect-, and the kinetic energy density of the uniform electron gas defined in Sec. III A exists when the domain is taken to be a tetrahedron.

Theorem IV.10

(Convergence rate for tetrahedra). Fix ρ0R+ and ν0R3. Let η and η̃ be regularization functions. Let e̲UEG(ρ0,ν0) and ēUEG(ρ0,ν0) denote the liminf and the limsup of eΔ,ηδ(ρ0,ν0) as ℓδ → ∞ and δ/ → 0.

  • For /δ sufficiently large, we have
  • We have for all 0<α12,
  • If ρ01/3C, then
  • The thermodynamic limit
    exists and is independent of the choice of the regularization function η.

Proof.
Proof of (i). The upper bound in (i) is proved by applying the lower bound of Theorem IV.7 with the choice
(28)
where ℓ′. Recalling (11), we get
(29)
Here, (III) can be bounded simply by Cρ0(1+δ1+δ3ρ0), using the trivial estimate:
(30)
valid for any α⩾1 due to the fact that 1Δηδ1.
Considering (I), we can proceed as follows. Define the index set
which collects all the small tetrahedra R(Δj + z + τ) that possibly intersect with the big smeared tetrahedron supp(1Δηδ). Furthermore, define
which contains small smeared tetrahedra which are well inside ′Δ. See Fig. 1.
Then, we may decompose the double sum in (I) into a sum over J0 and one over J\J0. The latter part is treated around (36), and the former part reads
(31)
Note first that 1Δjηδ=(1Δηδ)Tj,1, where Tj,(x) = Rjx + zj is the isometry that puts Tj,(Δ) = Δj. Setting
Using Lemma IV.5 (i), we have
(32)
Further, the same relation holds with E replaced by T.
In other words, we expressed the energy in any (smeared) tetrahedron well inside ′Δ in terms of the energy in the (smeared) reference tetrahedron Δ, plus a gauge term, which is position-dependent and grows faster than the volume. More concretely, 31 can be written as
(33)
where
The term involving CδT() can be bounded using (3) (discarding all but the gauge term) as
(34)
The gauge correction term (II) in (29) can be rewritten using the partition of unity
see (23), as
(35)
Again, this sum may be split into two parts: the J0 part can be written as
where in the last step we used once more that the smeared tetrahedron has barycenter 0, due to R3xη(x)dx=0. Note that the last term cancels exactly with the second term of (33). In total, the J0 part of (I) + (II) may be bounded by [using also (34)]
Next, we need to bound the part J \ J0 of (I) corresponding to tetrahedra at the boundary. We use the Lieb–Oxford inequality and (3) (again dropping all but the gauge term) to get
(36)
Again, we see that the second term cancels exactly with the remaining, J \ J0 part of the gauge correction term (II), see (35). The first term may be bounded by
Let us bound the quantities M and M. If SO(3)eΔ,η̃δ(ρ0,Rν0)dR0, then we can use the trivial bound M/|′Δ|⩽1. The boundary tetrahedra indexed by J \ J0 are at a distance O(+δ+δ) from (′Δ), and hence they fill a volume of M=O(2(+δ+δ)). But then M⩾|′Δ| − Cℓ2( + δ + δ′), so we arrive at the bound
(37)
where we have set
(38)
Next, we use Corollary III.4 to bound part of the l.h.s. of the preceding inequality. To this end, note that v(x)=1supp(1Δ*ηδ)(x)12ν0×x is clearly divergence-free and has R3ρ|Dv|=|Δ|22ρ0|ν0|. We have
where we also used the fact that 13R3|1Δηδ|2=O1δ. We obtain
After taking the liminf as ′ → ∞ and dividing by (1 − Cσδ/), we arrive at
as desired.
Proof of (ii). The proof of the lower bound is similar. We use the upper bound of Theorem IV.9 with the same choice (28). We have for any 0<δ̃<2 and 0<α12,
(39)
where we used the notation β=1δ̃/. The smearing parameter δ̃ is chosen so that δ̃=βδ, i.e., δ̃=δ/(1+δ/).
Define J and J0 analogously. First, we consider the summation over J0 in the first term. Using tTj(1δ̃/)Δ=tβRjΔ+tzj, we may write
where T is defined as above except with tℓ in place of . Here, we used the scaling relation
(40)
valid for all a > 0. We have
(41)
where in the first step we used Lemma IV.5, and (21) in the second (with λ = β−3 and dropping the gauge terms).
Furthermore, note that the function ν0E((1tΔη̃tδ)(ρ0,12ρ0Rν0×)) is convex for all fixed ρ0 and R, using the convexity of F. Hence, we find
where used 0<β<12. Here, we used that E(ρ, 0) = E(ρ), see Proposition II.6.
The gauge correction term in (39) may be written similarly to (35). The J0 part reads
We see that the second term here cancels with the last term of (41).
Next, we consider the J\J0 part. We use Corollary III.4 to bound the first term of (39),
For the J\J0 part of the gauge correction term, we use the partition of unity with holes to cancel the gauge term in the previous estimate. In total, we have obtained
(42)
The proof of (ii) is finished after taking the limsup as δ′ → ∞, δ′/′ → 0.
Proof of (iii). Finally, to show (iii), we substitute tℓ and δ in (37), apply 1α1+αdtt4 to the inequality and use (ii), with the roles of η and η̃ interchanged,
Here, we have modified σ in suitable way [cf. (38)]. The choice δ=1/3ρ04/9 leads to
and setting =()3/5ρ02/15 results in
using ρ01/3C.
Proof of (iv). To show that the limit exists, it is enough to prove that ēUEG(ρ0,ν0)e̲UEG(ρ0,ν0). But this follows from (i) combined with (iii). It remains to prove that eUEG(ρ0, ν0) is independent of the choice of the regularization function. Let ẽUEG(ρ0,ν0) denote eUEG(ρ0, ν0), but with ηδ replaced with η̃δ. Using (iii) and then (i), we find that the following chain of inequalities hold true,
from which we deduce ẽUEG(ρ0,ν0)=eUEG(ρ0,ν0). □

By omitting the Coulomb interaction, we find the following.

FIG. 1.

The tetrahedra corresponding to R=1 and τ = 0 included in the sets J0 and J\J0 are depicted in blue and red, respectively. The tiling is composed of translates of Δj.

FIG. 1.

The tetrahedra corresponding to R=1 and τ = 0 included in the sets J0 and J\J0 are depicted in blue and red, respectively. The tiling is composed of translates of Δj.

Close modal

Proposition IV.11.

Fix ρ0R+ and ν0R+. Let η and η̃ be regularization functions. Let τ̲UEG(ρ0,ν0) and τ̄UEG(ρ0,ν0) denote the liminf and the limsup of τΔ,ηδ(ρ0,ν0) as ℓδ → ∞ and δ/ → 0.

  • For , we have
  • We have
  • The thermodynamic limit
    exists and is independent of the choice of the regularization function η.

The proof is very similar to the one in Ref. 4. Let {ΩN} be a sequence of domains such that |ΩN| → ∞, |ΩN + Br|⩽CrN|2/3 for r⩽|ΩN|1/3/C and δN/|ΩN|1/3 → 0, δNN|1/3 → ∞.

We can define the index sets J and J0 for the domain ΩN in an obvious manner,
which collects all the small tetrahedra R(Δj + z + τ) that possibly intersect the smearing of ΩN. Furthermore, define
which contains small smeared tetrahedra which are well inside ΩN.
The argument for the tetrahedron ′Δ works exactly the same for the general domain ΩN, we only need to bound the quantities M and M to obtain the analog of (37). If eΩN,δN(ρ0,ν0)0, then we can use the trivial bound M/|ΩN|⩽1. The boundary tetrahedra indexed by J \ J0 are at a distance r = + δ + δN and hence, by the uniform regularity hypothesis, they fill a volume of
for sufficiently large N, since r⩽|ΩN|1/3/C by the convergence δN/|ΩN|1/3 → 0. But then M⩾|ΩN| − C( + δ + δ′)|ΩN|2/3, so we arrive at the bound
Taking δ fixed and = |ΩN|1/6, the r.h.s. goes to eUEG(ρ0, ν0) as δN/|ΩN|1/3 → 0.
For the upper bound, we use the analog of (42), with α = 1/2,
Using now δNN|1/3 as well as δN/|ΩN|1/3 → 0, and the choice = |ΩN|1/6, δ = |ΩN|−1/12, we see that the r.h.s. goes to eUEG(ρ0, ν0), and the proof is finished.

The proof for the kinetic energy per volume τΩN,δN(ρ0,ν0) is very similar. The limit for eΩN,δNxc(ρ0,ν0) is obtained by subtracting τΩN,δN(ρ0,ν0) from eΩN,δN(ρ0,ν0) and taking the limit term-by-term.

Proof of Proposition II.1.

Part (i). Following Ref. 12, it is convenient to consider the so-called intrinsic kinetic energy tensor,
(43)
where πxixζγ(x)ργ(x). We immediately have that 0⩽ωγ(x) = ωγ(x). Also, ωγ is invariant under gauge transformations. For clarity, we omit the subscripts γ. Note first that
(44)
We find
which upon expansion of the r.h.s. and using Theorem 6.17 in Ref. 25,
yields the desired bound.

Part (ii). Follows from the Cauchy–Schwarz inequality. □

Proof of (3). Let γ(x,y)=j1λjφj(x)φj(y)̄, for some 0⩽λj⩽1 with j⩾1λj = N and {φj}j1H1(Rd)L2-orthonormal. Let {φjn}Cc(R3) be such that φjn=1 and φjnφj in H1 as n. Denoting ρn(x)=j1λj|φjn(x)|2, we have that ρρn0 and ρρn0 by an argument similar to Proof of Theorem 1.3 in Ref. 16. It is easy to see that we also have jnpρnjpρ0. Using Proposition II.1 (i), we may write
where jnp and νn are defined in terms of the φjn. Using expansion (44),
we obtain the desired bound. □

A.L. and M.A.Cs. acknowledge funding through ERC StG REGAL under Agreement No. 101041487. M.A.Cs. is grateful to Mathieu Lewin and Giovanni Vignale for helpful discussions. A.L., M.A.Cs. and E.I.T. acknowledge funding from RCN through the Hylleraas Center under Agreement No. 262695. E.I.T. acknowledges funding from RCN under Agreement No. 287950. A.L. and M.A.Cs. also received partial funding from the Norwegian Research Council through Grant Nos. 287906 (CCerror) and 262695 (CoE Hylleraas Center for Quantum Molecular Sciences).

The authors have no conflicts to disclose.

Mihály A. Csirik: Writing – original draft (equal). Andre Laestadius: Writing – original draft (equal). Erik I. Tellgren: Writing – original draft (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
M.
Lewin
,
E. H.
Lieb
, and
R.
Seiringer
, “
Statistical mechanics of the uniform electron gas
,”
J. Ec. Polytech.-Math.
5
,
79
116
(
2018
).
2.
E. H.
Lieb
and
H.
Narnhofer
, “
The thermodynamic limit for jellium
,”
J. Stat. Phys.
12
,
291
310
(
1975
).
3.
M.
Lewin
,
E. H.
Lieb
, and
R.
Seiringer
, “
Floating Wigner crystal with no boundary charge fluctuations
,”
Phys. Rev. B
100
,
035127
(
2019
).
4.
M.
Lewin
,
E. H.
Lieb
, and
R.
Seiringer
, “
The local density approximation in density functional theory
,”
Pure Appl. Anal.
2
,
35
73
(
2020
).
5.
G.
Giuliani
and
G.
Vignale
,
Quantum Theory of the Electron Liquid
(
Cambridge University Press
,
2005
).
6.
I.
Fushiki
,
E. H.
Gudmundsson
,
C. J.
Pethick
, and
J.
Yngvason
, “
Matter in a magnetic field in the Thomas-Fermi and related theories
,”
Ann. Phys.
216
,
29
72
(
1992
).
7.
E. H.
Lieb
,
J. P.
Solovej
, and
J.
Yngvason
, “
Asymptotics of heavy atoms in high magnetic fields: I. Lowest Landau band regions
,”
Commun. Pure Appl. Math.
47
,
513
591
(
1994
).
8.
E. H.
Lieb
,
J. P.
Solovej
, and
J.
Yngvason
, “
Asymptotics of heavy atoms in high magnetic fields: II. Semiclassical regions
,”
Commun. Math. Phys.
161
,
77
124
(
1994
).
9.
N.
Rougerie
,
S.
Serfaty
, and
J.
Yngvason
, “
Quantum Hall phases and plasma analogy in rotating trapped bose gases
,”
J. Stat. Phys.
154
,
2
50
(
2014
).
10.
N.
Rougerie
, “
On the Laughlin function and its perturbations
,”
Sémin. Laurent Schwartz—EDP Appl.
,
2
17
(
2019
).
11.
J. F.
Dobson
, “
Spin-density functionals for the electron correlation energy with automatic freedom from orbital self-interaction
,”
J. Phys.: Condens. Matter
4
,
7877
7890
(
1992
).
12.
S.
Sen
and
E. I.
Tellgren
, “
A local tensor that unifies kinetic energy density and vorticity in density functional theory
,”
J. Chem. Phys.
149
,
144109
(
2018
).
13.
A.
Laestadius
,
E. I.
Tellgren
,
M.
Penz
,
M.
Ruggenthaler
,
S.
Kvaal
, and
T.
Helgaker
, “
Kohn–Sham theory with paramagnetic currents: Compatibility and functional differentiability
,”
J. Chem. Theory Comput.
15
,
4003
4020
(
2019
).
14.
M.
Hoffmann-Ostenhof
and
T.
Hoffmann-Ostenhof
, “‘
Schrödinger inequalities’ and asymptotic behavior of the electron density of atoms and molecules
,”
Phys. Rev. A
16
,
1782
(
1977
).
15.
J. E.
Harriman
, “
Orthonormal orbitals for the representation of an arbitrary density
,”
Phys. Rev. A
24
,
680
(
1981
).
16.
E. H.
Lieb
, “
Density functionals for Coulomb systems
,”
Int. J. Quantum Chem.
24
,
243
277
(
1983
).
17.
A.
Laestadius
, “
Density functionals in the presence of magnetic field
,”
Int. J. Quantum Chem.
114
,
1445
1456
(
2014
).
18.
E. H.
Lieb
and
R.
Schrader
, “
Current densities in density-functional theory
,”
Phys. Rev. A
88
,
032516
(
2013
).
19.
E. I.
Tellgren
,
S.
Kvaal
, and
T.
Helgaker
, “
Fermion N-representability for prescribed density and paramagnetic current density
,”
Phys. Rev. A
89
,
012515
(
2014
).
20.
G.
Vignale
and
M.
Rasolt
, “
Current- and spin-density-functional theory for inhomogeneous electronic systems in strong magnetic fields
,”
Phys. Rev. B
37
,
10685
10696
(
1988
).
21.
S.
Kvaal
,
A.
Laestadius
,
E.
Tellgren
, and
T.
Helgaker
, “
Lower semicontinuity of the universal functional in paramagnetic current–density functional theory
,”
J. Phys. Chem. Lett.
12
,
1421
1425
(
2021
).
22.
C.
Hainzl
,
M.
Lewin
, and
J. P.
Solovej
, “
The thermodynamic limit of quantum Coulomb systems Part I. General theory
,”
Adv. Math.
221
,
454
487
(
2009
).
23.
V.
Bach
,
E. H.
Lieb
, and
J. P.
Solovej
, “
Generalized Hartree–Fock theory and the Hubbard model
,”
J. Stat. Phys.
76
,
3
89
(
1994
).
24.

Note that this state reduces to the cited one by taking v ≡ 0 and θ = δ0.

25.
E. H.
Lieb
and
M.
Loss
,
Analysis
(
American Mathematical Society
,
2001
), Vol.
14
.
26.
E. H.
Lieb
and
R.
Seiringer
,
The Stability of Matter in Quantum Mechanics
(
Cambridge University Press
,
2010
).
27.
E. H.
Lieb
, “
A lower bound for Coulomb energies
,”
Phys. Lett. A
70
,
444
446
(
1979
).
28.
E. H.
Lieb
and
S.
Oxford
, “
Improved lower bound on the indirect Coulomb energy
,”
Int. J. Quantum Chem.
19
,
427
439
(
1981
).
29.
G.
Kin-Lic Chan
and
N. C.
Handy
, “
Optimized Lieb–Oxford bound for the exchange-correlation energy
,”
Phys. Rev. A
59
,
3075
(
1999
).
30.
M.
Lewin
,
E. H.
Lieb
, and
R.
Seiringer
, “
Improved Lieb-Oxford boudn on the indirect and exchange energies
,”
Lett. Math. Phys.
112
(
5
),
92
(
2022
).
31.
G. M.
Graf
and
D.
Schenker
, “
On the molecular limit of Coulomb gases
,”
Commun. Math. Phys.
174
,
215
227
(
1995
).
32.
G. M.
Graf
, “
An electrostatic inequality with applications to the constitution of matter
,”
J. Équations Dériv. Partielles
1996
,
11
.