<<

Energy of non-relativistic many-particle quantum states from

non-commuting Coulomb field and momentum operators

Purnima Ghale∗

Department of Mechanical Science and Engineering,

University of Illinois at Urbana-Champaign

1206 W Green St., Urbana, IL, 61801 and

Cornell High Energy Synchrotron Source

Wilson Lab, Synchrotron Drive, Ithaca, NY, 14850

(Dated: March 24, 2021)

Abstract

We derive a functional form for the energy of interacting many-particle systems from first principles.

The effective zero-point energy of a Coulomb-interacting quantum system can be considered in terms of

fluctuations around its classical electrostatic interaction and fluctuations of momentum. This perspective

provides a controlled and intuitive way to realize the Hohenberg-Kohn theorem. Specifically, an uncertainty

equation that relates the interaction energy, kinetic energy and local charge density is derived; it is then

combined with the Lieb-Thirring bound on kinetic energy. The combination of these two constraints results

in the functional form for total energy of interacting many particle systems that applies for bosons as well as

fermions, with appropriate consideration for (anti)symmetry of wavefunctions under coordinate exchange. arXiv:2010.01656v2 [.comp-ph] 23 Mar 2021 In the uniform density limit, the functional form is consistent with (QMC) data for

the bosonic fluid, as well as fermionic systems – paramagnetic, ferromagnetic and BCC Wigner crystalline.

[email protected]

1 I. INTRODUCTION

We consider the many-body Schrodinger¨ equation with the Hamiltonian given by

2 ! X ∇α X 1 1 ˆ H Ψ = − + + Vext Ψ (1) 2 2 |xα − xβ| α α6=β

2 ˆ ∇α ˆ 1 where T = − and Vee = denote kinetic and electron-electron interaction energies re- 2 |xα−xβ | spectively, with Ψ(x1, . . . , xN ) as the (anti)symmetric wavefunction under coordinate exchange.

The Hamiltonian contains instantaneous interaction between particles at xα and xβ, in Coulomb

gauge: many-particle coordinates, (x1, . . . , xN ), are used but the temporal components of (t1, x1) are ignored; instead a common time coordinate applies for the entire system that we heuristically assign to (t, r) where r is the spatial coordinate on which the charge density, n(r), and exter- nal potential, Vext(r), are defined. We further note that Equation 1 does not contain transverse electromagnetic fields or operators that couple spin degrees of freedom.

Since Dirac expressed the energy of interaction as the sum of Hartree and exchange terms [1], it is generally assumed that the physics of the interacting many-electron system cannot be further simplified, and that one needs to proceed with approximate methods. Bohm and Pines sought to understand interacting many-particle systems in a series of papers that focused on charge fluc- tuations, treating the electromagnetic field in a covariant manner [2–5], but while this approach provided many insights, it did not solve the many-electron problem. Hohenberg and Kohn[6], fol- lowed by Kohn and Sham [7], then reframed the many-electron problem into a single-particle effective potential, motivating Density Functional Theory. In parallel, numerical simulations of many-particle wavefunctions with Quantum Monte Carlo [8, 9] have informed standardized solutions to the many-electron problem in the uniform density limit. Fundamentally, the non- relativistic many-electron problem sits at the intersection of two obstructions. First, the number

2 of possible many-body antisymmetric wavefunctions is too large and antisymmetry impedes effi-

cient sampling in Quantum Monte Carlo via the sign-problem [10, 11]. Second, in the strongly

interacting case, the Hamiltonian is not separable into a sum of independent quasiparticle Hamil-

tonians; non-separability indicates chaos [12] and if one starts with a Kohn-Sham ansatz, strong

interactions result in an explosion in the number of Feynmann diagrams required. Perhaps, there

is some physics we have missed.

In considering this problem, we first briefly focus on Density Functional Theory (DFT) [6, 13–

16], where the ground state total energy is expressed as expectation values of the operators

Z min ˆ ˆ min E[n] = hΨ|H |Ψi = hΨn |T + Vee|Ψn i + Vext(r)n(r) (2)

min where Ψn denotes the (anti)symmetric wavefunction that results in the prescribed charge den- sity n(r) and minimizes E[n]. Under the Kohn-Sham ansatz, a Slater determinant is constructed from the eigenstates of an effective single-particle Hamiltonian. For a given charge density,

ˆ ˆ n(r), bounds on hVeei and hT i [15, 17–21] can inform a hierarchy of functional forms for the ˆ ˆ exchange-correlation energy [22–26], but the behavior of the sum hΨ|T + Vee|Ψi and the effect

of (anti)symmetric Ψ remains unknown. As a result, although the Hohenberg-Kohn theorem im-

plies a mapping between many-particle quantum systems and scalar charge density fields, and the

development of density functionals has led to accurate schemes [23], the physical mechanism by

which nature constrains the energy of many-particle states via the charge density has not been

proposed.

Our work aims to understand interacting, non-relativistic, many-particle systems by first con-

sidering this fundamental gap, and is a variation on the simple harmonic oscillator. In particular,

ˆ pˆ2 xˆ2 consider the ground state of a quantum harmonic oscillator H = 2 + 2 ; its classical counter-

3 part is the solution with hxi = 0 and hpi = 0. The zero-point energy of this system is hHi =

∆p2/2 + ∆x2/2, which is constrained by the uncertainty relation, [p, x] = i~ =⇒ ∆p∆x ≥ ~/2.

Similarly, for non-relativistic, interacting, many-particle systems, understanding the quantum me- chanical interaction, hVeei, in terms of quantum fluctuations around the electrostatic one, and considering the necessary commutation relations, is key to our understanding. Thus, we obtain a fundamentally intuitive way to realize the Hohenberg-Kohn theorem, and the resulting functional

ˆ ˆ form for hT + Veei is consistent with Quantum Monte Carlo simulations in the uniform density limit [9]. The derivation also provides a way to correct small but systematic biases due to assumed

N and V dependence in those simulations, and makes explicit the importance of two length scales

2/3 – V , which is dependent on the volume, and rs, which is dependent on the charge density – on ˆ ˆ our understanding of hT + Veei of many-particle systems.

Note, however, that the problems of electrostatics and its eventual quantization in field theory

– namely self-interaction due to point charges and negative normalization of vacuum states cor- responding to the scalar potential of Aµ = (V (r), A(r, t)) – also plague our solution as they are inherited by the many-particle Schrodinger equation in Equation 1. A covariant treatment of the electromagnetic field in the strongly interacting limit, should be our ultimate goal, but is currently beyond the scope of this work.

Section II presents the theoretical derivation and the functional form for non-uniform density, which is then generalized to the uniform density case. Section III then fits the functional form in the uniform density limit to Quantum Monte Carlo (QMC) data from [9], and investigates the efficacy of this functional form in capturing the energies as well as energy-differences of quantum states.

4 II. THEORETICAL RESULTS

A. Individual terms in the Hamiltonian

Let us start by setting the total momentum, hpˆi = 0, i.e. the many-electron system is not under-

going translation – this is not essential but simplifies our derivation. Particle indistinguishability

then implies[20, see p. 320] hpˆ1,ii = hpˆ2,ii = ... = hpˆN,ii = 0, and the kinetic energy is the sum

of fluctuations of momentum, N 3 2 X X ∆pα,i hTˆi = (3) 2 α=1 i=1

where α denotes particle labels and i denotes xyz-directions.

The scalar potential at r is given by Vˆ (r) = PN 1 Ψ(x ,..., x ), and its gradient, α=1 |r−xα| 1 N ˆ E(r) = −∇rV (r). Using what we know from classical electrostatics, i.e. the square of the

Coulomb electric field denotes the energy density of interaction between charged particles, let us

expand the quantum product

Z Z ˆ ˆ X 1 1 hΨ|E(r) · E(r)|Ψi = hΨ|∇r · ∇r |Ψi (4a) |r − xα| |r − xβ| r α,β r

2 1 integrating by parts in r, setting the surface term to zero, and using the identity, ∇r |r−x| =

−4πδ(r − x) in 3 dimensions results in

Z Z ˆ δ(r − x1) hΨ|E · E(r)|Ψi = 8π hΨ|Vee|Ψi + 4πN hΨ| |Ψi (4b) r |r − x1|

4πN R 1 with the last term simplifying to a constant, √ 3 2 . A similar analysis for a classical product ( 2π) k |k|

5 of operators shows

Z Z 1 n(r)n(r0) hΨ|E(r)|Ψi · hΨ|E(r)|Ψi = 8π 0 (4c) r r,r0 2 |r − r |

Taking the difference between Eqs 4b and 4c, the left-hand-side denotes fluctuations of the

R P3 2 Coulomb field, r j=1 ∆Ej (r), and rearrangement leads to

Z 0 Z Z ˆ 1 n(r)n(r ) 1 X 2 N 1 3 hVeei = 0 + ∆Ej (r) − √ 3 2 d k (4d) 0 2 |r − r | 8π  |k| r,r r j 2 2π

where the last term is due to self-interaction – we get this term whenever we interchange point-

interactions, 1 , by fields in classical or quantum theory. We can now express the expectation |xα−xβ |

value of the Hamiltonian in Equation 1 as E[n] = Eclassical + Equantum where

Z 1 Z n(r)n(r0) N Z 1 E = V (r)n(r) + − (5a) classical ext 0 √ 3 2 r 2 r,r0 |r − r | 2 2π k |k|

2 Z X ∆pα,j 1 X E = + ∆E2(r) (5b) quantum 2 8π j α,j r j

B. Relationship between operators

Next, we focus on Equantum by inspecting the relationship between the momentum and electric

field operators via commutation

∂E (r)  1  [ˆp , Eˆ (r)]Ψ = ~ j Ψ = −~ −∂2 Ψ j,α j rj (6) i ∂xj,α i |r − xα|

6 where applying the Cauchy-Schwartz inequality following Robertson[27], results in a relationship

~ 2 1 between fluctuations of the kind, ∆pj,α∆Ej(r) ≥ hΨ| − ∂ |Ψi . Using |g(x)| ≥ g(x), 2 rj |r−xα|

2 1 summing over three dimensions, and using ∇ |r−x| = −4πδ(r − x), eventually leads to the P3 relation j=1 ∆pj,α∆Ej(r) ≥ 2π hnˆα(r)i. Summation over particle labels, assuming isotropy of fluctuations, and using particle indistinguishability we get

3N∆p∆E(r) ≥ 2πn(r) (7)

which has not been derived before. Equation 7 suggests that by fixing n(r), we are essentially real- izing a generalized uncertainty constraint on the quantum fluctuations of the many-particle system, albeit an uncertainty relation that is far more limited. In particular, the regime of applicability is limited to non-relativistic systems with many-particle wavefunctions Ψ(x1, . . . , xN ), with charge density (and the corresponding electrostatic field) defined over the spatial field, r. Note also that the application of the Cauchy-Schwartz inequality assumes that the operators are positive definite, but we know that vacuum states of scalar components of the electromagnetic potential in Coulomb gauge have negative normalization in field theory [28]. This may not be a problem, however, if we consider that the scalar potential takes the place of Aµ=0 in the electromagnetic potential, while the momentum appears as pµ=1−3 in the relativistic four-momentum, suggesting that if both oper- ators are collapsed to the non-relativistic limit c → ∞, they require different sign-rules to define inner products. More work is necessary to ensure that it is mathematically rigorous. Finally, spin and transverse components of the electromagnetic field need to be included in the future, and will likely result in current density functionals.

7 C. Energy of many-particle systems

Let us now consider how the kinetic energy of a system is constrained by the charge density.

ˆ Lieb has shown that bounds on hVeei by themselves do not depend on the (anti)symmetry of

the wavefunction under coordinate exchange[18], but the kinetic energy is constrained by the

exchange symmetry (and spin-degeneracy) of the many-body wavefunction. In particular,[21]:

Z 5/3 hT i ≥ Cke n (r) (8a) r

Z 2 5/3 hT i = 3N∆p /2 ≥ Cke n (r) (8b) r

The value of Cke is a constant for fermionic systems (the constant itself depends on the internal symmetry, for example, spin-degeneracy, of the wavefunction), but for bosonic wavefunctions,

−2/3 Cke has an additional factor of ∝ N . The same notation, Cke(N), is used for now in all cases.

Instead of using the Lieb-Thirring bound on kinetic energy, one could also compute the kinetic energy from Kohn-Sham wavefunctions and use that to inform momentum fluctuations and thus develop meta functionals in the Jacob’s ladder of density functionals [22].

Collecting terms in E[n] = Equantum[n] + Eclassical[n] that we have derived so far, we get

Z π R n2 Z 1 Z n(r)n(r0) N Z 1 E[n] = C n5/3(r)+ + V (r)n(r)+ − (9) ke R 5/3 ext 0 √ 3 2 2CkeN n 2 |r − r | 2 2π k |k|

where the first term is due to the constraint on kinetic energy, and the second term arises from

the constraint on the local electric field. The remaining terms are from Eclassical[n]. For later ˆ ˆ comparison with Quantum Monte Carlo data in [9], we consider parts that correspond to hT + Veei,

8 R i.e. excluding r Vext(r)n(r); then, in the uniform density limit,

π n2V n2 Z 1 N Z 1 hTˆ + Vˆ i = C n5/3V + + d3rd3r0 − (10a) ee ke 5/3 0 √ 3 2 2CkeN n V 2 |r − r | 2 2π k |k|

N π n2 n2 Z 1 N Z 1 = C n5/3 + + d3rd3r0 − (10b) ke 5/3 0 √ 3 2 n 2CkeN n 2 |r − r | 2 2π k |k|

Above, note that it is usually possible to either cancel out V from the numerator and denominator,

R R 1 or use n = N/V , with the exception of a factor in the third term, r r0 |r−r0| ; we denote this

4π 3 1 quantity as A(V ). Using the definition of the Wigner-Seitz radius, rs, as 3 rs = n , then results in

the expression:

N Z 1 π  3 1/3  3 2/3  3 2 = − + + NC + A(V )/2 (10c) √ 3 2 3 ke 3 3 2 2π k |k| 2NCke 4πrs 4πrs 4πrs

1/3 2 π(3/4π) 1 2/3 1 A(V )(3/4π) 1 = c1(N) + + NCke(3/4π) 2 + 6 (10d) 2NCke rs rs 2 rs

Rearranging Equation 10d in decreasing powers of rs and encapsulating N and V dependent terms as c(N) or c(V ), we get:

ˆ ˆ uniform 1 1 1 hT + Veei [n] = c1(N) + c2(N) + c3(N) 2 + c4(V ) 6 (10e) rs rs rs

The functional form thus presented arises primarily from the relationship between operator fluctu- ations. The effect of (anti)symmetry of the many-body wavefunction under exchange is included in the value (and N-dependence) of Cke, which in turn affects c2(N) and c3(N). Considering the

form of Equation 10e in powers of rs, one is struck by the simplicity of the formula, as well as

1 the realization that the rs dependence includes an unexpected 6 term – Wigner crystals, in par- rs

9 1 ticular, are predicated on the assumption that the kinetic energy of fermionic systems scales as 2 , rs while the potential energy scales as 1 [29]. Our work is consistent with [29], if we remember a rs

subtle variation: the inclusion of the effects of the uniform external potential on the charge den- R sity, Vext(r)n(r) as well as the interaction between classical background charges, not included in Equation 1 [30], are necessary for the total energy to reduce to the traditional scaling of 1 and rs

1 ˆ ˆ 1 R 1 2 terms. Otherwise, hT + Veei contains an additional 6 0 0 term, that while dimension- rs rs r,r |r−r | ally consistent with 1 , cannot be reduced to it. It is worthwhile to consider the dependencies in rs ˆ ˆ hT + Veei because a non-uniform classical background could give rise to a uniform charge density.

ˆ ˆ Therefore, considering only the quantum system defined by hT + Veei, the two length scales

3 are characteristic: rs determined by n, and L determined by volume V = L , which are also

evident in the scaling of single-particle momentum fluctuations of bosonic vs. fermionic systems.

One can show by hand using the Lieb-Thirring bound on kinetic energy [21] that, in the uniform

2 Cke density limit, momentum fluctuations of bosons are constrained by ∆pα ≥ V 2/3 , while momentum

2 1 fluctuations of fermionic wavefunctions are constrained by ∆p ≥ Cke 2 . α rs

III. NUMERICAL IMPLICATIONS

We present numerical implications of this work in the context of Quantum Monte Carlo (QMC)

simulations in [9], to investigate both the strengths and possible improvements of our approach.

Other QMC simulations are not used as they can introduce small and systematic biases due to

finite-size effects.

10 A. Parameterization procedure

Finite-particle simulations in [9] use E(N) = E0 + E1/N + E2/∆N to extrapolate to infinite particles, with the third term absent for bosonic systems. Our work suggests, instead, that expec-

ˆ ˆ tation value of each term in hT + Veei scales differently with N and V . However, since QMC simulations in [9] are the current standard in electronic structure theory, we test the consistency of the functional form in the uniform density limit by fitting the values of {c1, . . . , c4} against ener- gies in [9]. Four states are considered: the bosonic fluid (BF), paramagnetic fermion fluid (PMF), ferromagnetic fluid (FMF), and BCC Wigner crystals (BCC).

The proposed functional form is physically well-motivated; as such, we consider whether it is consistent not only with energies but energy differences as well, and whether it fails where it should, i.e. when relevant physics such as spin-orbit coupling is missing. Energy differences between paramagnetic fermion and bosonic fluid for example, would be consistent if,

PMF BF PMF BF EQMC − EQMC ≈ E[n] − E[n]

PMF PMF BF BF Rearrangement results in EQMC −E[n] ≈ EQMC −E[n] . Let the left and right-hand-sides be represented by e(PMF ) and e(BF ) respectively, we can then fit the functional form from

Equation 10e to Quantum Monte Carlo data by minimizing a single cost function written as the

2-norm of deviations from reference QMC data,

||[e(BF ), e(PMF ), e(FMF ), e(BCC)]||2

11 added to the sum of differences between deviation of reference QMC data

||e(BF ) − e(PMF )||2 + ||e(BF ) − e(FMF )||2 + all non-repeating pairs

by using the fmincon numerical optimizer set to – zero initial parameter-guesses; “active-set” optimizer algorithm; machine-tolerance for cost-function tolerance; 106 function evaluations; 103 optimizer evaluations – within the MATLAB Optimizer Toolbox [31]. Note that these heuristics do not guarantee a unique or globally optimal solution, but are helpful to confirm that the functional form derived in this work are consistent with QMC data.

B. Results: energies and coefficients

Figure 1 shows the efficacy of the functional form derived in this work; discrete points show

Quantum Monte Carlo data from [9] while continuous plots represent parametric fits to the func- tional form. Each discrete point also has an associated error bar indicating deviations of the func- tional form from QMC values, which we qualitatively note, are small.

Table I presents the coefficients obtained by the optimization procedure outlined in Sec- tion III A. Note that the sign of c2, coefficient of the term which accounts for the quantum fluctu- ation of the Coulomb electric field, is negative – this is expected as our treatment includes only the zero-th component of the electromagnetic potential. Furthermore, we note that this parame- terization is not unique, as it is possible to find other combinations of parameters {c1, . . . c4} that match QMC data to within chemical accuracy of 1.5 mHa, but rely on the details of numerical optimizer setting. Equally important but less well-understood are the small but systematic biases due to extrapolation of finite size QMC simulation to infinite particles; simulations in [9] consider number dependence to be the most significant concern in terms of error.

12 0 0.8

0.6 -0.1 0.4

-0.2 0.2

0 -0.3 -0.2

-0.4 -0.4 0 50 100 150 200 0 20 40 60 80 100

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0

-0.2 -0.2

-0.4 -0.4 0 20 40 60 80 100 0 50 100 150 200

FIG. 1. hTˆ + Vˆeei of quantum many-body systems in the uniform density limit for four different quantum states. Discrete points in each plot denote Quantum Monte Carlo (QMC) data from [9] while the continuous lines represent parametric fits to the functional form in Equation 10e. Each discrete point also has an associated error bar that denotes deviation from the QMC values, which are within chemical accuracy (see Figure 2).

state c1 c2 c3 c4 BF -0.0042 -1.2300 0.7783 -1.8025 PMF -0.0039 -1.2630 2.5674 -0.1264 FMF -0.0038 -1.2890 3.4322 2.7008 BCC -0.0043 -1.2252 2.0053 x

TABLE I. Parameters obtained by fitting Equation 10e to Quantum Monte Carlo simulations[9]. We note that these are not unique coefficients, but can be affected by systematic biases due to finite-size effects in [9], as well as the optimizer settings used – as we have used a numerical optimizer to fit parameters, it is not a guarantee that they have converged to the global minimum. Finally, note that c4 for the BCC Wigner case appears to have no effect on the optimization procedure, likely due to the lack of data-points in the high density region where c4 is relevant. We refrain from adding additional data from other simulations, as they also bring small but systematic biases.

Next, Figure 2 shows the quantitative deviation of the functional form from reference QMC

13 data: for all four quantum states, deviations are within 1.5 mHa, or within chemical accuracy.

Furthermore, errors for each state are in-tandem with the others, both in sign as well as in the

magnitude of error. This suggests that the functional form might possess enough sensitivity to

capture phase transitions, which is explored in the next section. Note that while chemical accuracy

requires deviations to be within 1.5 mHa, quantum phase transitions require sharper margins as

energy differences between quantum phases can be in the order of meV.

1.5

1

0.5

0

-0.5

-1

-1.5 100 101 102

FIG. 2. Deviations of the functional form with respect to QMC data for different systems can be minimized in-tandem. This is useful because it reveals systematic deviations that affect all quantum states. Here, we note that deviations from QMC data are less than 1.5 mHa, and that they can be minimized in-tandem. This suggests that the functional form is capable of reproducing the physics of Quantum Monte Carlo simulations – energies as well as energy differences.

C. Results: phase transitions

Figures 3a-d show the energy differences between different pairs of quantum states, ∆E [meV].

Continuous curves denote results from the functional form while discrete points denote corre- sponding QMC data. Each subfigure contains two panels – the first one presents the difference in energy, ∆E, while the second panel shows log10 |∆E| that amplifies any changes. A vertical line

14 through both panels indicates a zero-temperature transition predicted by QMC simulations [9].

5 5 0

-5 0

-10 -5 -15 0 50 100 150 200 0 50 100 150 200

105

100 100

0 50 100 150 200 0 50 100 150 200

(a) BF-BCC transitions. (b) PMF-BCC transitions.

0 5

-5 0

-10 -5

-15 0 20 40 60 80 100 0 50 100 150 200

100

-1 100 10

0 20 40 60 80 100 0 50 100 150 200

(c) PMF-FMF transitions. (d) FMF-BCC transitions.

FIG. 3. Energy differences between different pairs of wavefunctions obtained from the functional fits in this work (dashed curve) and QMC data in [9] (discrete points). A vertical line denotes the transition rs predicted by [9]. In addition to ∆E, log10(|∆E|) are presented to explore singularities due to the functional form. In (a-b), note that as expected, our work agrees best with QMC data when spin-orbit coupling is less relevant. Overall, error-bars or quantitative deviations from QMC data are quite small, but between fermionic systems, the functional form predicts additional transitions at low rs that are unsupported by QMC. This could be the effect of strong (non-perturbative) electron-electron interactions, or systematic effects due to the minimum fermion nodes required to model them[11]. Alternatively, this could be mediated by spin-orbit coupling, or due to effects of empirical fitting. More detailed theoretical and computational work is necessary.

Overall, given that energy differences are in the meV-range, the functional form agrees well with QMC data – with an important caveat. The best agreement between QMC data and our functional form is found for the BF-BCC transition shown in Figure 3a: the functional form

15 correctly predicts the transition rs and agrees qualitatively with predictions in [9]. Transitions

between fermionic states, presented in Figures 3b-d, however, are more complicated: although

values of ∆E generally agree quantitatively with QMC data, the functional form predicts extra

phase-transitions at high densities that are not supported by the QMC data in [9].

Specifically, Figure 3b shows energy differences between the paramagnetic and BCC Wigner

crystal states, where we note excellent quantitative agreement. QMC data indicate a transition

near rs ≈ 100, agreeing with the functional form, but our work predicts an additional transition

near rs < 50, which is neither supported nor simulated by QMC data. A similar situation occurs in

Figure 3c, which shows energy differences between the paramagnetic and ferromagnetic fermionic

fluids, where we note larger error-bars since spin-orbit coupling is missing. Again, while QMC

data and our functional form agree more-or-less on the transition near rs > 65, they do not agree on the transition at low rs. Finally, for completeness, we also include Figure 3d which shows energy differences between the ferromagnetic fermion fluid and the BCC Wigner crystal – QMC data in [9] do not mention a transition involving these two states and larger quantitative errors appear due to exclusion of spin-orbit coupling.

Thus, while quantitative agreement with QMC data is promising, disagreement on phase tran- sitions between fermionic systems at low rs (or high density), needs further consideration. One ˆ ˆ possibility is that the functional form of this work is fitted to values of hT + Veei and therefore

1 contains 6 , leading to physically correct nonlinearities at low rs, which are nevertheless cancelled rs by background interaction in case of jellium. Secondly, the inclusion of spin-orbit coupling might not only improve quantitative results on the PMF-FMF transition, but could also smooth the low rs behavior of ∆E. In addition, other causes such as artifacts of empirical fitting and subtler effects of strong electron-electron interaction in QMC simulations, need to be considered. For this, more detailed theoretical and computational work is necessary.

16 IV. CONCLUSION

In this work, we have shown that it is possible to derive the energy of interacting non-relativistic quantum systems, despite the non-separability of the Hamiltonian and the antisymmetry-induced sign problems. The key is to observe that the electromagnetic field through which particles in- teract, is itself a quantum mechanical quantity, with its own fluctuations, and as a result, fixing a local distribution of charge density necessitates wavefunctions to follow a generalized uncer- tainty relation. Note, however, that we are bound to the Coulomb gauge and the treatment of the electromagnetic field is not covariant. Thus, more work is necessary to ultimately resolve interact- ing systems. Numerically, the functional form shows promising agreement with Quantum Monte

Carlo data in the uniform density limit. The agreement is remarkable because energy differences in the meV-regime are hard to match, and here we have used a relatively simple and physically trans- parent functional with four coefficients for each quantum state. Nevertheless, some complications appear in the high density limit, in that the functional form predicts additional phase transitions.

More work is necessary to fully understand these discrepancies with respect to Quantum Monte

Carlo simulations.

Within the realm of chemistry and materials science, more work is necessary to apply the find- ings of this work, and the limits where the anisotropy of fluctuations may be important. A plethora of challenges remain – such as the thermodynamic limit of total energy in strongly interacting many-particle systems, the implications of this work in the context of periodicity, and current and spin density functional theories. Finally, even within Density Functional Theory of finite sys- tems, while the expression for the total energy has now been derived, obtaining the single particle exchange-correlation potential, Vxc, could be challenging in the non-uniform density case.

17 Acknowledgements This material is based in part upon work supported by the Department of Energy, National Nuclear Security Administration, under Award Number DE-NA0002374, in collaboration with Prof. Harley T. Johnson at the University of Illinois. Discussions with John- son, David Ceperley, Andre Schleife and Lucas Wagner at University of Illinois are all gratefully acknowledged. This work is also based in part upon research conducted at the Center for High

Energy X-ray Sciences (CHEXS) which is supported by the National Science Foundation un- der award DMR-1829070. Discussions with Jacob Ruff at the Cornell High Energy Synchrotron

Source (CHESS) are gratefully acknowledged.

[1] Paul AM Dirac. Note on exchange phenomena in the thomas atom. In Mathematical Proceedings of

the Cambridge Philosophical Society, volume 26, pages 376–385. Cambridge University Press, 1930.

[2] David Bohm and David Pines. A collective description of electron interactions: I. Magnetic interac-

tions. Physical Review, 82(5):625, 1951.

[3] David Pines and David Bohm. A collective description of electron interactions: II. Collective vs

individual particle aspects of the interactions. Physical Review, 85(2):338, 1952.

[4] David Bohm and David Pines. A collective description of electron interactions: III. Coulomb interac-

tions in a degenerate electron gas. Physical Review, 92(3):609, 1953.

[5] David Pines. A collective description of electron interactions: IV. Electron interaction in metals.

Physical Review, 92(3):626, 1953.

[6] Pierre Hohenberg and Walter Kohn. Inhomogeneous electron gas. Physical Review, 136(3B):B864,

1964.

[7] LJ Sham and Walter Kohn. One-particle properties of an inhomogeneous interacting electron gas.

Physical Review, 145(2):561, 1966.

18 [8] D Ceperley. Ground state of the fermion one-component plasma: A Monte Carlo study in two and

three dimensions. Physical Review B, 18(7):3126, 1978.

[9] David M Ceperley and BJ Alder. Ground state of the electron gas by a stochastic method. Physical

Review Letters, 45(7):566, 1980.

[10] David M Ceperley. Fermion nodes. Journal of Statistical Physics, 63(5-6):1237–1267, 1991.

[11] DK Sunko. Natural generalization of the ground-state slater determinant to more than one dimension.

Physical Review A, 93(6):062109, 2016.

[12] Fritz Haake. Quantum Signatures of Chaos, volume 54. Springer Science & Business Media, 2013.

[13] Walter Kohn and Lu Jeu Sham. Self-consistent equations including exchange and correlation effects.

Physical Review, 140(4A):A1133, 1965.

[14] Mel Levy. Universal variational functionals of electron densities, first-order density matrices, and nat-

ural spin-orbitals and solution of the v-representability problem. Proceedings of the National Academy

of Sciences U.S.A., 76(12):6062–6065, 1979.

[15] Elliott H Lieb. Density functionals for couiomb systems. International Journal of Quantum Chemistry,

24:243, 1983.

[16] Mel Levy and John P Perdew. Hellmann-feynman, virial, and scaling requisites for the exact uni-

versal density functionals. shape of the correlation potential and diamagnetic susceptibility for atoms.

Physical Review A, 32(4):2010, 1985.

[17] Elliott H Lieb. A lower bound for coulomb energies. Physical Letters A, 70(5-6):444–446, 1979.

[18] Elliott H Lieb and Stephen Oxford. Improved lower bound on the indirect coulomb energy. Int. J.

Quant. Chem., 19(3):427–439, 1981.

[19] Garnet KL Chan and Nicholas C Handy. Optimized Lieb-Oxford bound for the exchange-correlation

energy. Physical Review A, 59(4):3075, 1999.

19 [20] Elliott H Lieb. Kinetic energy bounds and their application to the stability of matter. In Inequalities,

Selecta of Elliot H Lieb, Eds. Loss and Ruskai, pages 317–328. Springer, 2002.

[21] Elliott H Lieb and Walter E Thirring. Inequalities for the moments of the eigenvalues of the

Schrodinger Hamiltonian and their relation to Sobolev inequalities. In The Stability of Matter: From

Atoms to Stars, pages 135–169. Springer, 1991.

[22] John P Perdew and Karla Schmidt. Jacob’s ladder of density functional approximations for the

exchange-correlation energy. In AIP Conference Proceedings, volume 577, pages 1–20. AIP, 2001.

[23] Jianwei Sun, Adrienn Ruzsinszky, and John P Perdew. Strongly constrained and appropriately normed

semilocal density functional. , 115(3):036402, 2015.

[24] Jeng-Da Chai and Martin Head-Gordon. Long-range corrected hybrid density functionals with

damped atom–atom dispersion corrections. Physical Chemistry Chemical Physics, 10(44):6615–6620,

2008.

[25] Axel D Becke. Density-functional exchange-energy approximation with correct asymptotic behavior.

Physical Review A, 38(6):3098, 1988.

[26] Jianmin Tao, John P Perdew, Viktor N Staroverov, and Gustavo E Scuseria. Climbing the density

functional ladder: Nonempirical meta–generalized gradient approximation designed for molecules

and solids. Physical Review Letters, 91(14):146401, 2003.

[27] Howard Percy Robertson. The uncertainty principle. Physical Review, 34(1):163, 1929.

[28] Lewis H Ryder. Quantum field theory. Cambridge university press, 1996.

[29] Eugene Wigner. Effects of the electron interaction on the energy levels of electrons in metals. Trans-

actions of the Faraday Society, 34:678–685, 1938.

[30] Assuming the Born-Oppenheimer approximation.

[31] Matlab optimization toolbox, 2020. Fmincon optimizer, with maximum function tolerance and itera-

20 tions, The MathWorks, Natick, MA, USA.

21