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

The Levy-Lieb embedding of density functional theory and its Quantum Kernel: Illustration for the Hubbard Dimer using near-term quantum algorithms

C. D. Pemmaraju dasc@ibm.com IBM, San Jose, Califonia, USA    Amol Deshmukh IBM Quantum, IBM T.J. Watson Research Center, Yorktown Heights, NY, USA
Abstract

The constrained-search formulation of Levy and Lieb provides a concrete mapping from NN-representable densities to the space of NN-particle wavefunctions and explicitly defines the universal functional of density functional theory. We numerically implement the Levy-Lieb procedure for a paradigmatic lattice system, the Hubbard dimer, using a modified variational quantum eigensolver (VQE) approach. We demonstrate density variational minimization using the resulting hybrid quantum-classical scheme featuring real-time computation of the Levy-Lieb functional along the search trajectory. We further illustrate a fidelity based quantum kernel associated with the density to pure-state embedding implied by the Levy-Lieb procedure and employ the kernel for learning observable functionals of the density. We study the kernel’s ability to generalize with high accuracy through numerical experiments on the Hubbard dimer.

Quantum Computing, Quantum Kernel, Machine Learning, Variational Quantum Eigensolver (VQE), Density Functional Theory (DFT)
preprint: APS/123-QED

I Introduction:

Density Functional Theory (DFT)  [1, 2] is presently the dominant paradigm for material-specific electronic structure simulations in computational materials science [3]. Originally proposed by Hohenberg and Kohn (HK) [1], DFT establishes the one-body density as the basic variable in interacting many-body problems with fixed interaction strength subject to time-independent spatially local external potentials [4]. The electronic Hamiltonian in the Born-Oppenheimer approximation is the most prominent example of such a setup. Given a problem in this class, all ground state observables can in principle be represented as explicit functionals of the ground state one-body density. However, while the density is informationally adequate, deriving explicit density functionals that are universally accurate is a hard problem in the same complexity class as computing generic ground states [5]. The practical success of DFT instead stems from the development of extremely low cost yet usefully accurate [6] approximate models for the quantum mechanical correlation energy in the form of highly compact density functionals [7] circumventing the need to represent many-electron wavefunctions explicitly on classical computers at simulation time. Benefits of the compactness of a density functional representation can also extend to quantum systems where efficient classical heuristics for the wavefunction exist [8, 9]. In the modern context the encodings implied by DFT may be viewed as simply the most condensed description within a hierarchy of functional theories based on reduced density matrices [5, 10].

With the advent of quantum computers the prospect of representing many-electron states on quantum hardware has important implications for DFT and other functional theories [5, 11, 12, 13, 14]. The classical intractability of interacting many-electron wavefunctions partly provided the motivation for functional theories [15] but at the same time also constrained the development of approximate functionals [9]. For instance the study of exact forms of DFT has to date been largely restricted to small model systems where systematic comparisons with wavefunction calculations are possible [9] and pure density functionals for a wider variety of observables are known in model systems than are available for deployment in realistic simulations of matter [9, 8, 16, 17]. Furthermore, the relationship between approximate functionals and underlying many-electron states is often obscured in practice even though the mapping can be recovered ex post facto at significant computational cost [18]. Finally and most relevant to this work is the fact that outside of a few pioneering efforts [19, 20, 21, 22, 23, 24] the explicit density to wavefunction mappings implied by DFT [25, 26, 27, 28, 29] are not widely pursued in numerical studies on classical computers and one typically encounters density functionals only after they are approximated either using formal, empirical or machine learning methods [7, 30]. The availability of quantum devices may alleviate this constraint in future. Therefore in this work we calculate the density to wavefunction mapping [28, 29, 31, 19, 20, 21, 24], as formulated by Levy [25, 26] and Lieb [27] using a density-constrained variational quantum eigensolver (VQE) [32, 33] scheme and demonstrate density variational minimization to identify the ground state of a paradigmatic fermion lattice problem, the Hubbard dimer [16, 17]. Furthermore, we make use of the density-wavefunction map to define a quantum kernel [34, 35, 36, 37] which we call the Levy-Lieb quantum kernel and use it to learn observable functionals in such a way that the relationship to the underlying state space is readily apparent. Our work contrasts with previous research efforts which explored proposals for the popular Kohn-Sham DFT [2] in both near-term and fault-tolerant settings [11, 12, 13, 14]. The rest of the article is organized as follows. In section II, we introduce the Levy-Lieb procedure which defines the density to wavefunction map and in section III we outline a concrete VQE based implementation of the same for lattice systems. Section  IV discusses results of an explicit density variational search for the ground state of the Hubbard dimer. In section V we introduce a fidelity based Levy-Lieb quantum kernel (LLQK) and illustrate its use in machine learning density functionals. We summarize our conclusions in section VI.

Refer to caption
Figure 1: (a) Qiskit [38] circuit diagram for the generalized unitary coupled cluster ansatz used to prepare parameterized quantum states for the asymmetric Hubbard dimer studied in this work. (b) For different values of the UU parameter, a comparison of VQE (filled circles) and exact diagonalization (solid line) results for the ground state energy of the Hubbard dimer as a function of the potential asymmetry Δv\Delta v. The inset shows the absolute error in the VQE energy relative to the exact diagonalization reference on a logarithmic scale.

II The Levy-Lieb mapping

Consider an interacting NN-particle system described by a Hamiltonian operator of the form

H^=T^+W^+v^\hat{H}=\hat{T}+\hat{W}+\hat{v} (1)

where T^,W^\hat{T},\hat{W} are the kinetic and inter-particle interaction operators and the external potential v^\hat{v} is a local one-body operator. Following Eschrig [31], we define the set of NN-representable densities as

𝒥N{n|n0,n1/2𝑳2,𝑑xn=N}\mathcal{J}_{N}\equiv\{n|n\geq 0,\nabla n^{1/2}\in\bm{L}^{2},\int dx~{}n=N\} (2)

with 𝑳2\bm{L}^{2} denoting the space of square-integrable functions, and the set of NN-particle wavefunctions as

𝒲N{Ψ|\displaystyle\mathcal{W}_{N}\equiv\{\Psi| Ψ(x1,..,xN)(anti)symmetric,Ψ|Ψ=1,\displaystyle\Psi(x_{1},..,x_{N})\mathrm{(anti)symmetric},\langle\Psi|\Psi\rangle=1, (3)
iΨ|iΨ<fori=1..N}\displaystyle\langle\nabla_{i}\Psi|\nabla_{i}\Psi\rangle<\infty~{}\mathrm{for}~{}i=1..N\}

The Levy-Lieb density functional FLL[n]F_{\mathrm{LL}}[n] is then defined by

FLL[n]inf{Ψ|T^+W^|Ψ|Ψn,Ψ𝒲N},n𝒥NF_{\mathrm{LL}}[n]\equiv\mathrm{inf}\{\langle\Psi|\hat{T}+\hat{W}|\Psi\rangle|\Psi\rightsquigarrow~{}n,\Psi\in\mathcal{W}_{N}\},n\in\mathcal{J}_{N} (4)

i.e., given a density n𝒥Nn\in\mathcal{J}_{N}, the Levy-Lieb procedure involves searching for the infimum of Ψ|T^+W^|Ψ\langle\Psi|\hat{T}+\hat{W}|\Psi\rangle over the restricted subset of wavefunctions Ψ𝒲N\Psi\in\mathcal{W}_{N} constrained to yield the chosen density nn. We use the notation Ψn\Psi\rightsquigarrow n to indicate Ψ\Psi yields nn and define 𝒲Nn{Ψ𝒲N|Ψn}\mathcal{W}^{n}_{N}\equiv\{\Psi\in\mathcal{W}_{N}|\Psi\rightsquigarrow n\}. Note that the above procedure does not involve the external potential v^\hat{v} and at the end of the search process we have access to both the value of FLL[n]F_{\mathrm{LL}}[n] and a Ψn\Psi\rightsquigarrow n. For general n𝒥Nn\in\mathcal{J}_{N} the obtained Ψ\Psi belongs to a sub-manifold of states on 𝒲Nn\mathcal{W}^{n}_{N} with the same expectation Ψ|T^+W^|Ψ\langle\Psi|\hat{T}+\hat{W}|\Psi\rangle [31] but in the absence of such degeneracies on 𝒲Nn\mathcal{W}^{n}_{N}, nn uniquely determines Ψ\Psi [39]. In particular, on the subset of densities that correspond to non-degenerate ground states, the mapping is bijective due to the Hohenberg-Kohn theorem [1, 31, 39]. In all cases, FLL[n]F_{\mathrm{LL}}[n] is well-defined. The density variational principle follows directly from the Levy-Lieb procedure. Consider the set 𝒲N\mathcal{W}_{N} and define the equivalence relation

ΨiΨjΨinandΨjn\Psi_{i}\sim\Psi_{j}\iff\Psi_{i}\rightsquigarrow n~{}\mathrm{and}~{}\Psi_{j}\rightsquigarrow n (5)

This leads to a partitioning of 𝒲N\mathcal{W}_{N} into disjoint subsets 𝒲Nn\mathcal{W}^{n}_{N} labelled by the density nn as Ψi𝒲NniffΨin\Psi_{i}\in\mathcal{W}^{n}_{N}~{}\mathrm{iff}~{}\Psi_{i}\rightsquigarrow n. On a particular 𝒲Nn\mathcal{W}^{n}_{N}, consider minimizing the expectation of the Hamiltonian from equation 1 with external potential v^\hat{v}

E[v^,n]\displaystyle E[\hat{v},n] =inf{Ψ|T^+W^+v^|Ψ,Ψ𝒲Nn}\displaystyle=\mathrm{inf}\{\langle\Psi|\hat{T}+\hat{W}+\hat{v}|\Psi\rangle,\Psi\in\mathcal{W}^{n}_{N}\}
=inf{Ψ|T^+W^|Ψ+𝑑xv(x)n(x),Ψ𝒲Nn}\displaystyle=\mathrm{inf}\{\langle\Psi|\hat{T}+\hat{W}|\Psi\rangle+\int dx~{}v(x)n(x),\Psi\in\mathcal{W}^{n}_{N}\}
=inf{Ψ|T^+W^|Ψ,Ψ𝒲Nn}+𝑑xv(x)n(x)\displaystyle=\mathrm{inf}\{\langle\Psi|\hat{T}+\hat{W}|\Psi\rangle,\Psi\in\mathcal{W}^{n}_{N}\}+\int dx~{}v(x)n(x)
=FLL[n]+𝑑xv(x)n(x)\displaystyle=F_{\mathrm{LL}}[n]+\int dx~{}v(x)n(x)

Since 𝒲N=n𝒥N𝒲Nn\mathcal{W}_{N}=\bigcup\limits_{n\in\mathcal{J}_{N}}\mathcal{W}^{n}_{N}, for the ground state energy we have

E[v^]\displaystyle E[\hat{v}] min{Ψ|T^+W^+v^|Ψ,Ψ𝒲N}\displaystyle\equiv\mathrm{min}\{\langle\Psi|\hat{T}+\hat{W}+\hat{v}|\Psi\rangle,\Psi\in\mathcal{W}_{N}\} (7)
=min{inf{Ψ|T^+W^+v^|Ψ,Ψ𝒲Nn},n𝒥N}\displaystyle=\mathrm{min}\left\{\mathrm{inf}\{\langle\Psi|\hat{T}+\hat{W}+\hat{v}|\Psi\rangle,\Psi\in\mathcal{W}^{n}_{N}\},n\in\mathcal{J}_{N}\right\}
=min{FLL[n]+𝑑xv(x)n(x),n𝒥N}\displaystyle=\mathrm{min}\{F_{\mathrm{LL}}[n]+\int dx~{}v(x)n(x),n\in\mathcal{J}_{N}\}

Thus the ground state energy E[v^]E[\hat{v}] can be obtained by minimizing E[v^,n]=FLL[n]+𝑑xv(x)n(x)E[\hat{v},n]=F_{\mathrm{LL}}[n]+\int dx~{}v(x)n(x) over n𝒥Nn\in\mathcal{J}_{N}. It follows further that if nGSn_{\mathrm{GS}} is the ground-state density then through equation 4 on 𝒲NnGS\mathcal{W}^{n_{\mathrm{GS}}}_{N}, it determines the ground-state manifold ΨGS\Psi_{\mathrm{GS}}. We note that the Hohenberg-Kohn (HK) functional is a restriction of FLL[n]F_{\mathrm{LL}}[n] to the space of ground state densities [31].

As formulated by Cioslowski [19, 20, 21] and previously illustrated by Mori-Sanchez et al [24] on classical computers, the minimization procedure of equation 7 defines an exact numerical density functional calculation for the ground state provided FLL[n]F_{\mathrm{LL}}[n] and it’s derivative δFLL[n]/δn{\delta}F_{\mathrm{LL}}[n]/{\delta}n are calculated explicitly through constrained search. In this work we calculate these quantities for discrete lattices by employing parameterized quantum circuits executed on a statevector simulator. In the following sections when discussing formal statements unrelated to specific implementation details we will indicate functionals as A[n]A[n] using square brackets and italicized nn as above but when referring to discretized implementations where parametric differentiation is substituted for formal functional differentiation [19, 20, 21, 40, 24], we will indicate functional dependence via A(𝐧)A(\mathbf{n}) using parenthesis and bold 𝐧\mathbf{n} notation for the density.

III Density-constrained VQE

Since first being proposed in 2014, the variational quantum eigensolver (VQE) [32] has been widely adopted on near-term quantum computers as a versatile hybrid classical-quantum algorithm for optimization tasks employing parameterized ansatze. The incorporation of constraints into VQE variational optimization has also been discussed previously [41, 42]. For a recent comprehensive review of VQE we refer readers to reference [33]. For our purpose, since equation 4 implies a search over wavefunction space restricted to a specified density sector, we augment the usual VQE with a density-constraint. If the Hamiltonian takes a form where the potential term appears as a local one body operator v^=iviσc^iσc^iσ\hat{v}=\sum_{i}v_{i}\sum_{\sigma}\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i\sigma} in second-quantization, then the one-body density operator of interest needed to specify the constraint is n^iσ=c^iσc^iσ\hat{n}_{i\sigma}=\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i\sigma} which is easily computed on any discrete index set {iσ}\{i\sigma\} enumerating basis modes ii with spin index σ\sigma. In the literature, this choice of relevant potential and density on a lattice goes by the name site-occupation functional theory [9] but the same form also appears in connection with discretized real space grids [24]. Then for NN electrons on MM sites and parameterized ansatz states |ψ(θ)|\psi(\theta)\rangle with real parameters θ\theta, the Levy-Lieb procedure consists of minimizing the expectation

FLL(θ)=ψ(θ)|T^+W^|ψ(θ)F_{\mathrm{LL}}(\theta)=\langle\psi(\theta)|\hat{T}+\hat{W}|\psi(\theta)\rangle (8)

subject to the constraints

ψ(θ)|σc^iσc^iσ|ψ(θ)=ni,i=1..M\langle\psi(\theta)|\sum_{\sigma}\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i\sigma}|\psi(\theta)\rangle=n_{i},i=1..M (9)

where {ni,i=1..M}\{n_{i},i=1..M\} are prescribed site occupations and additionally we have

0<ni2,i=1Mni=N\begin{array}[]{l}0<n_{i}\leq 2,\\ \sum\limits_{i=1}^{M}n_{i}=N\end{array} (10)

The steps involved in density-constrained VQE (DC-VQE) are outlined in table 1. We expect based on the work of D’Amico et al [23] that for vv-representable densites, i.e., densities that correspond to NN-particle ground states in some potential vv, small errors in computing densities do not lead to large errors in the wavefunction as nearby densities get mapped to nearby wavefunctions in metric space and densities on lattices are vv-representable under very reasonable assumptions [43, 44]. Thus the constrained minimization process should be robust to small numerical errors in computing densities.

Table 1: Density-constrained VQE (DC-VQE)
1. Accept a vector of site occupations
𝐧[ni,i=1..M]~{}\mathbf{n}\equiv[n_{i},i=1..M] such that {0ni2ini=N}\left\{\begin{array}[]{c}0\leq n_{i}\leq 2\\ \sum_{i}n_{i}=N\end{array}\right\}
2. Construct the vector of site density difference operators
𝐃^[σc^iσc^iσni,i=1..M]\mathbf{\hat{D}}\equiv[\sum_{\sigma}\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i\sigma}-n_{i},i=1..M]
3. Set circuit parameters θθguess\theta\leftarrow\theta_{guess} to prepare
initial state Ψ(θ)=R(θ)|0\Psi(\theta)=R(\theta)|0\rangle
4. For 𝐝Ψ(θ)|𝐃^|Ψ(θ)\mathbf{d}\equiv\langle\Psi(\theta)|\mathbf{\hat{D}}|\Psi(\theta)\rangle measured on a quantum device,
update θ\theta to minimize 𝐝||\mathbf{d}|| using a classical optimizer.
For 𝐧𝒥N\mathbf{n}\in\mathcal{J}_{N}, 𝐝<ϵ||\mathbf{d}||<\epsilon at the minimum and yields a state
Ψ(θ)𝒲N𝐧,𝐧Bϵ(𝐧)\Psi(\theta)\in\mathcal{W}^{\mathbf{n^{\prime}}}_{N},\mathbf{n^{\prime}}\in B_{\epsilon}(\mathbf{n})
5. With FLL(θ)Ψ(θ)|T^+W^|Ψ(θ)F_{\mathrm{LL}}(\theta)\equiv\langle\Psi(\theta)|\hat{T}+\hat{W}|\Psi(\theta)\rangle and 𝐝=Ψ(θ)|𝐃^|Ψ(θ)\mathbf{d}=\langle\Psi(\theta)|\mathbf{\hat{D}}|\Psi(\theta)\rangle
both measured on a quantum device, update θ\theta to minimize
FLL(θ)F_{\mathrm{LL}}(\theta) under the constraint 𝐝=0||\mathbf{d}||=0 using a constrained
classical optimizer.
6. At the optimum θ\theta^{*} set FLL(𝐧)=FLL(θ)F_{\mathrm{LL}}(\mathbf{n})=F_{\mathrm{LL}}(\theta^{*}) and optionally
save θ\theta^{*}. Since 𝐝<ϵ||\mathbf{d}||<\epsilon we have Ψ(θ)𝐧,𝐧Bϵ(𝐧)\Psi(\theta^{*})\rightsquigarrow\mathbf{n^{\prime}},\mathbf{n^{\prime}}\in B_{\epsilon}(\mathbf{n}).

Once a subroutine for calculating FLL(𝐧)F_{\mathrm{LL}}(\mathbf{n}) is available, density variational minimization for a given external potential can be implemented as outlined in table 2 wherein, following earlier implementations for classical computers [19, 20, 21, 24], we compute the derivative δFLL(𝐧)/δ𝐧{\delta}F_{\mathrm{LL}}(\mathbf{n})/{\delta}\mathbf{n} using finite differences while maintaining normalization of the density. The next section outlines a specific implementation of DC-VQE and density variational search for the case of the asymmetric Hubbard dimer.

Table 2: Density variational minimization
1. Accept a vector of on-site potentials 𝐯[vi,i=1..M]~{}\mathbf{v}\equiv[v_{i},i=1..M]
to define the full Hamiltonian
2. Use either a non-interacting or mean-field solution to setup
guess site occupations 𝐧\mathbf{n} and ansatz parameters θguess\theta_{guess}
3. Update 𝐧\mathbf{n} to minimize FLL(𝐧)+𝐯.𝐧F_{\mathrm{LL}}(\mathbf{n})+\mathbf{v.n} subject to the
constraints {0<ni2ini=N}\left\{\begin{array}[]{c}0<n_{i}\leq 2\\ \sum_{i}n_{i}=N\end{array}\right\} using a classical constrained
optimizer that calls DC-VQE to evaluate FLL(𝐧)F_{\mathrm{LL}}(\mathbf{n})
4. At the optimum 𝐧\mathbf{n^{*}} return the ground state energy
E(𝐯)=FLL(𝐧)+𝐯.𝐧E(\mathbf{v})=F_{\mathrm{LL}}(\mathbf{n^{*}})+\mathbf{v.{n^{*}}}
Refer to caption
Figure 2: The Levy-Lieb functional for the asymmetric Hubbard dimer is shown for different values of the UU parameter.

IV Asymmetric Hubbard dimer

The asymmetric Hubbard dimer has been explored extensively as a paradigmatic model system within lattice density functional theory [16] and time-dependent density functional theory [17] and serves as an ideal test case to explore near-term variational quantum algorithms and quantum machine learning in the context of DFT. The relevant fermion Hamiltonian

H^=\displaystyle\hat{H}= tσ=,(c1σc2σ+h.c.)\displaystyle-t\sum_{\sigma=\uparrow,\downarrow}(c^{\dagger}_{1\sigma}c_{2\sigma}+h.c.) (11)
+Uin^in^i+ivi(n^i+n^i)\displaystyle+U\sum_{i}\hat{n}_{i\uparrow}\hat{n}_{i\downarrow}+\sum_{i}v_{i}(\hat{n}_{i\uparrow}+\hat{n}_{i\downarrow})

features a local multiplicative external potential term as the source of inhomogeneity in the model. It is further characterized by two dimensionless parameters U/tU/t and Δv/t\Delta v/t where Δv=v1v2\Delta v=v_{1}-v_{2} is the potential asymmetry between the two dimer sites. To enable direct comparison with previous benchmark results [16, 17] we fix 2t=12t=1 and vary UU, Δv\Delta v across different calculations. Additionally, since for the on-site occupations 𝐧=[n1,n2]\mathbf{n}=[n_{1},n_{2}] we have n1+n2=Nn_{1}+n_{2}=N, a single parameter n1n_{1} is chosen to specify the density distribution. We consider the case of N=2N=2 particles in four site-localized spin orbitals and restrict our attention to the lowest energy spin-singlet (S2=0,Sz=0S^{2}=0,S_{z}=0) which is sufficient for analyzing ground state DFT of the dimer [16]. The above setup leads via the Jordan-Wigner mapping [45] to a 4-qubit problem on a quantum computer. For this simple model we find that a generalized unitary coupled cluster [46] (UCC) wavefunction ansatz featuring single and double excitations along with two Trotter steps is sufficient to reproduce the exact-diagonalization result for the ground state energies over the parameter range of interest in conjunction with VQE [32, 33] (See fig 1). For a general discussion of specialized ansatze for lattice models in different dimensions we refer the readers to recent literature [33]. In our study the UCC circuit ansatz which features ten parameters as shown in figure 1(a) is set up using the Qiskit [38] framework and executed on a statevector simulator backend. VQE classical parameter optimization is performed using the L-BFGS-B [47] optimizer as implemented in the SciPy [48] package. As shown in fig. 1(b) for a range of (U,Δv)\left(U,\Delta v\right) parameters, energies of the dimer spin-singlet ground-state as obtained from VQE are within 10610^{-6} of corresponding exact-diagonalization values and in good correspondence with figure 33 of reference [16]. We then proceed to employ the same circuit ansatz to prepare states for evaluation within density-constrained VQE simulations of the dimer.

Refer to caption
Figure 3: (a) Evolution of the error in satisfying the density constraint of equation 9 during the course of computing the Levy-Lieb functional via constrained optimization. (b) Absolute error relative to a VQE reference in the total energy during a density variational minimization for the ground state. (c) Absolute error relative to a VQE reference in the site occupation n1n_{1} during a density variational minimization for the ground state.

As outlined in section III and table 1, density-constrained VQE (DC-VQE) is intended as a subroutine to calculate the Levy-Lieb functional FLL(𝐧)F_{\mathrm{LL}}(\mathbf{n}) for a specified density 𝐧\mathbf{n} and makes no reference to the external potentials 𝐯\mathbf{v}. Thus with the operator sum T^+W^\hat{T}+\hat{W} and a set of site-occupation difference operators 𝐃\mathbf{D} as defined in table 1 specified, the UCC ansatz parameters are optimized with VQE under an occupation-preserving constraint (equation 9) to yield FLL(𝐧)F_{\mathrm{LL}}(\mathbf{n}). We implement the constraint in practice by setting up the SLSQP [49] constrained optimizer from SciPy [48] to request site-occupation measurements from the quantum simulator. For the dimer case, the DC-VQE simulated result for FLLF_{\mathrm{LL}} which is just a function of n1n_{1} is plotted in figure 2 for different values of UU. As noted previously by Carrascal et al [16], there is no known analytical expression for FLL(n1)F_{\mathrm{LL}}(n_{1}) but it is a monotonic function on the interval n1(0,1]n_{1}\in(0,1] and approaches a bound UU as n10n_{1}\rightarrow 0. Furthermore since dFLL(n1)dn1=Δv2\frac{dF_{\mathrm{LL}}(n_{1})}{dn_{1}}=\frac{\Delta v}{2} [16] its slope increases sharply as n10n_{1}\rightarrow 0 but remains finite for physical potentials. It is apparent that the DC-VQE result for FLL(n1)F_{\mathrm{LL}}(n_{1}) exhibits the expected features.

In connection with DC-VQE it is instructive to analyze the degree to which the occupation-preserving constraint of equation 9 is respected during the optimization process for FLL(n1)F_{\mathrm{LL}}(n_{1}). This is shown in figure 3(a) where we plot the constraint vector magnitude 𝐝||\mathbf{d}|| as defined in table 1 along a typical optimization trajectory. We see that starting from a small value of 𝐝||\mathbf{d}|| as the initial state is prepared to lie near 𝒲N𝐧\mathcal{W}^{\mathbf{n}}_{N} the SLSQP optimizer maintains a low baseline value of 107\sim 10^{-7} for 𝐝||\mathbf{d}|| but with isolated spikes in between where it assumes larger values signifying departures of the trial wavefunction from 𝒲N𝐧\mathcal{W}^{\mathbf{n}}_{N} at some instances. Therefore our approach for finding FLL(n1)F_{\mathrm{LL}}(n_{1}) is not a strict constrained-search in the sense of Levy [25, 24] but since 𝐝||\mathbf{d}|| is always small at convergence we ensure that the optimal state lies numerically close to 𝒲N𝐧\mathcal{W}^{\mathbf{n}}_{N} as required.

With a means of calculating FLL(n1)F_{\mathrm{LL}}(n_{1}) at hand, we then proceed to perform density functional simulations for a given external potential asymmetry Δv\Delta v to identify the dimer ground state. This involves a variational search over occupation-numbers as outlined in table 2 combined with on the fly calculation of FLL(n1)F_{\mathrm{LL}}(n_{1}) using DC-VQE and it’s derivative FLL(n1)F_{\mathrm{LL}}^{\prime}(n_{1}) using finite differences. Thus the density variational minimizer repeatedly calls DC-VQE at occupations 𝐧\mathbf{n} it encounters along the optimization trajectory. To initialize the DFT simulation we solve the non-interacting problem with U=0U=0 and use the resulting occupation numbers as the starting guess. One could also for instance employ an approximate exchange-correlation potential vxcv_{xc} [9, 16] in the dimer Hamiltonian to provide an initial density guess that will in most cases be closer to the true ground state density than the non-interacting guess. For the given Δv\Delta v we also compute the regular VQE ground state energy and density as the benchmark result. In figure 3(b,c) we show convergence with respect to VQE of the DFT energy and site occupation n1n_{1} as a function of the number of FLL(n1)F_{\mathrm{LL}}(n_{1}) evaluations which represent the most intensive part of the simulation. We set Δv=1\Delta v=1 and show trajectories for U=1U=1 and U=5U=5. In this parameter regime the correlation energy is typically sizable (see figure 6 of reference [16]) while at the same time the non-interacting and exact densities are expected to differ significantly. We find that around 20-30 evaluations of FLL(n1)F_{\mathrm{LL}}(n_{1}) are needed to converge the DFT energy and occupations to within 10410^{-4} and 10310^{-3} of the respective VQE reference values.

To conclude this section we note that while Levy-Lieb constrained-search [25, 26, 27, 28, 29] clarifies the coarse-graining step involved in switching our description of quantum systems from wavefunctions to densities, DFT based on repeated real-time evaluations of the exact FLL[𝐧]F_{\mathrm{LL}}[\mathbf{n}] using explicit constraints would be inefficient relative to unconstrained minimization in wavefunction space. Furthermore constrained optimization of parameterized quantum circuits is generally non-convex [50, 33, 36, 37] and gradient based optimizers may get stuck in one of many local minima and therefore one needs to employ additional heuristics to locate the global minimum of the quantity being optimized. In the dimer case investigated here, similar to the case of variational quantum deflation on small molecules as noted previously [51], randomly initializing different starting guesses for the UCC ansatz parameters proved sufficient to identify the correct optimum for FLL(𝐧)F_{\mathrm{LL}}(\mathbf{n}). For larger problem sizes more sophisticated global optimization schemes will be needed. Ideally specialized density-preserving parameterized ansatze that efficiently explore a specified density sector of Hilbert space would be desirable.

Refer to caption
Figure 4: (a) For the asymmetric Hubbard dimer, training, predicted and actual values of FLL(n1)F_{\mathrm{LL}}(n_{1}) when learning with the Levy-Lieb quantum kernel in an interpolaive setting. (b) Evolution of the RMS test error as a function of the number of training points for interpolative learning of FLL(n1)F_{\mathrm{LL}}(n_{1}) with different values of UU. Results are shown both for the Levy-Lieb quantum kernel (filled-circles connect by solid lines) and the classical Gaussian kernel (dots connected by dashed lines). (c) Training, predicted and actual values of FLL(n1)F_{\mathrm{LL}}(n_{1}) when learning with the Levy-Lieb quantum kernel in an extrapolative setting. (d) Evolution of the RMS test error as a function of the number of training points for extrapolative learning of FLL(n1)F_{\mathrm{LL}}(n_{1}) with different values of UU. Results are shown both for the Levy-Lieb quantum kernel (filled-circles connected by solid lines) and the classical Gaussian kernel (dots connected by dashed lines). (e) Training, predicted and actual values of FLL(n1)F_{\mathrm{LL}}(n_{1}) when learning with the classical Gaussian kernel in an interpolative setting. (f) Distribution of test errors along the domain of n1n_{1} when learning FLL(n1)F_{\mathrm{LL}}(n_{1}) in an interpolaive task with the classical Gaussian kernel.

V The Levy-Lieb Quantum Kernel

Restricting ourselves to non-degenerate settings where nn uniquely determines Ψ\Psi [39] as outlined in section II, we can think of the Levy-Lieb mapping as implementing a particular feature map [35, 34, 37, 36] encoding densities into the feature space LL\mathcal{F}_{\mathrm{LL}} of pure-state density matrices ρLL[n]|ΨLL[n]ΨLL[n]|LL\rho_{\mathrm{LL}}[n]\equiv|\Psi_{\mathrm{LL}}[n]\rangle\langle\Psi_{\mathrm{LL}}[n]|\in\mathcal{F}_{\mathrm{LL}}. Additionally for the same density nn, different embeddings can be realized through different choices for the operator sum T^+W^\hat{T}+\hat{W}. So for a specified T^+W^\hat{T}+\hat{W}, this allows us to define a fidelity based Levy-Lieb quantum kernel (LLQK)

κLL(n,n)Tr{ρLL[n]ρLL[n]}=|ΨLL[n]|ΨLL[n]|2\kappa_{\mathrm{LL}}(n,n^{\prime})\equiv\mathrm{Tr}\{\rho_{\mathrm{LL}}[n^{\prime}]\rho_{\mathrm{LL}}[n]\}=|\langle\Psi_{\mathrm{LL}}[n^{\prime}]|\Psi_{\mathrm{LL}}[n]\rangle|^{2} (12)

for densities n,n𝒳N𝒥Nn,n^{\prime}\in\mathcal{X}_{N}\subset\mathcal{J}_{N} where 𝒳N{n𝒥N|nΨLL𝒲Nn}\mathcal{X}_{N}\equiv\{n\in\mathcal{J}_{N}|n\iff\Psi_{\mathrm{LL}}\in\mathcal{W}^{n}_{N}\} is the set on which the Levy-Lieb map is one-to-one. Note that this restriction is not needed if one only wishes to learn the functional FLL[𝐧]F_{\mathrm{LL}}[\mathbf{n}] alone and furthermore in situations with degeneracies one may optionally provide additional quantum numbers to pick a particular state on the degenerate manifold to uniquely specify the LLQK. We can use the LLQK in a machine learning model of the form

f𝒪[n]=Tr{ρLL[n]𝒪}\displaystyle f_{\mathcal{O}}[n]=\mathrm{Tr}\{\rho_{\mathrm{LL}}[n]\mathcal{O}\} =\displaystyle= i=1MαmTr{ρLL[n]ρLL[nm]}\displaystyle\sum\limits_{i=1}^{M}\alpha_{m}\mathrm{Tr}\{\rho_{\mathrm{LL}}[n]\rho_{\mathrm{LL}}[n_{m}]\} (13)
=\displaystyle= i=1MαmκLL(nm,n)\displaystyle\sum\limits_{i=1}^{M}\alpha_{m}\kappa_{\mathrm{LL}}(n_{m},n)

where the goal is to learn the optimal operator measurement 𝒪\mathcal{O} or equivalently the coefficients αm\alpha_{m}\in\mathbb{R} given a set of MM labelled training data {nm,f𝒪[nm]}\{n_{m},f_{\mathcal{O}}[n_{m}]\}. In particular since ground state density functionals are generated as expectation values of the form

O[n]=ΨLL[n]|O^|ΨLL[n]=Tr{ρLL[n]O^},O[n]=\langle\Psi_{\mathrm{LL}}[n]|\hat{O}|\Psi_{\mathrm{LL}}[n]\rangle=\mathrm{Tr}\{\rho_{\mathrm{LL}}[n]\hat{O}\}, (14)

where O^=O^\hat{O}=\hat{O}^{\dagger} is Hermitian, then given data (nm,O[nm])(n_{m},O[n_{m}]), m=1Mm=1...M the optimal measurement 𝒪\mathcal{O} is just the projection of O^\hat{O} in 𝒮LLM=span{ρLL[n1],ρLL[n2]ρLL[nM]}\mathcal{S}^{M}_{\mathrm{LL}}=\mathrm{span}\{\rho_{\mathrm{LL}}[n_{1}],\rho_{\mathrm{LL}}[n_{2}]...\rho_{\mathrm{LL}}[n_{M}]\}. In other words, the reproducing kernel hilbert space (RKHS) of the LLQK [36, 37] consists only of density functionals of the form O[n]=Tr{ρLL[n]O^}O[n]=\mathrm{Tr}\{\rho_{\mathrm{LL}}[n]\hat{O}\}.

Here we explore the learning abilities of the LLQK in the context of the asymmetric Hubbard dimer. We calculate the LLQK for the dimer using the optimal UCC ansatz states obtained from DC-VQE for specified occupations n1n_{1} in the relevant interval (0,1]\left(0,1\right]. Since we consider pure state embeddings the fidelities are straightforwardly computed on a statevector simulator using unitary adjoints as |Ψ(θ)|Ψ(θ)|2=|0|𝒰(θ)𝒰(θ)|0|2|\langle\Psi(\theta^{\prime})|\Psi(\theta)\rangle|^{2}=|\langle 0|\mathcal{U}^{\dagger}(\theta^{\prime})\mathcal{U}(\theta)|0\rangle|^{2}. On an actual quantum computer, the kernel entries can be estimated by evolving the initial state |0|0\rangle with 𝒰(θ)𝒰(θ)\mathcal{U}^{\dagger}(\theta^{\prime})\mathcal{U}(\theta) and calculating the ratio of all-zero outcomes to the total number of shots. We expect state fidelities to be robust to potential non-uniqueness of the mapping from densities n1n_{1} to raw UCC parameters θ\theta as under the Levy-Lieb procedure distances in density space should get mapped to physically meaningful distances in state space. Furthermore, since in this study we fix 2t=12t=1, we only vary the interaction term W^\hat{W} by varying UU in the dimer Hamiltonian. UU can therefore be thought of as a hyperparamter that fixes the kernel and we study kernels κLL(U)\kappa^{(U)}_{\mathrm{LL}} labeled by the UU value employed to generate the state embeddings leading to the kernel.

Refer to caption
Figure 5: (a) Decay of the singular values of the Levy-Lieb quantum kernel matrices for different UU. (b) Relative kernel alignment between Levy-Lieb quantum kernel matrices for different values of UU. (c) RMS test errors while learning the interacting FLL(5)(n1)F_{\mathrm{LL}}^{(5)}(n_{1}) with U=5U=5 using the non-interacting quantum kernel κLL(0)\kappa^{(0)}_{\mathrm{LL}} and vice versa.

For a first set of learning tasks we consider FLL(n1)F_{\mathrm{LL}}(n_{1}) as the target function and construct a data set by sampling n1(0,1]n_{1}\in(0,1] on a uniform grid with spacing 0.01 and at each n1n_{1} calculate FLL(U)(n1)F_{\mathrm{LL}}^{(U)}(n_{1}) and Ψ(U)(n1)\Psi^{(U)}(n_{1}) for different UU. We then consider two kinds of train-test data splits: (L1) an in-distribution or interpolative learning task and (L2) an out-of-distribution or extrapolative learning task. For L1 we select MM training samples distributed as uniformly as possible over the interval (0,1](0,1] while including the end points n1=0+n_{1}=0^{+} and n1=1n_{1}=1 in the training set. All remaining points are used as test data points. This ensures that the sample mean for n1n_{1} over the training and test sets is very close and in this 1D example every test data point has at at least one training point on either side. For L2 we select MM training samples distributed uniformly over the last 40%40\% of the interval (0,1](0,1] and include n1=1n_{1}=1 in the training set, while all other points are used as test data. In this instance the sample mean of n1n_{1} over the training and test data is far apart and over 50% of the test data can be reached only through 1D extrapolation. For a given MM and specified UU, we precompute the kernel matrices κLL(U)\kappa^{(U)}_{\mathrm{LL}} for the training and test data sets by evaluating fidelities on a noiseless quantum simulator. For each training set size MM considered, we verify the M×MM\times M κLL(U)\kappa^{(U)}_{\mathrm{LL}} matrices are positive semidefinite and then use classical kernel ridge regression (KRR) as implemented in the scikit-learn library [52] to perform fitting and prediction of FLL(U)(n1)F_{\mathrm{LL}}^{(U)}(n_{1}). The regularization hyperparameter of KRR is set to a small value of 4×10154\times 10^{-15} only to ensure numerical stability of the KRR fit process as MM approaches the rank of the kernel matrices. Thus we work in a setting where low training fit error is favored. For tasks L1, L2 we also use a classical Gaussian kernel κG=exp(𝐧𝐧2/2σ2)\kappa_{G}=\mathrm{exp}(-||\mathbf{n}-\mathbf{n}^{\prime}||^{2}/2\sigma^{2}) to fit and predict FLL(U)(n1)F_{\mathrm{LL}}^{(U)}(n_{1}) for each U. In order to choose the Gaussian kernel hyperparameter σ\sigma we use kernal alignment [53, 54, 55] defined for two given kernel matrices K,KK,K^{\prime} as:

A(K,K)=K,KFK,KFK,KFA(K,K^{\prime})=\frac{\langle K,K^{\prime}\rangle_{F}}{\sqrt{\langle K,K\rangle_{F}\langle K^{\prime},K^{\prime}\rangle_{F}}} (15)

Accordingly for each (U,M)(U,M), we separately optimize σ\sigma so as to maximize A(κG,κLL(U))A(\kappa_{G},\kappa^{(U)}_{\mathrm{LL}}) on the training data set.

Results for learning tasks L1, L2 as a function of the number of training points MM are shown in fig. 4. We use the root mean square (RMS) error on the test data set as our performance metric. Figure 4(a), shows plots for FLL(n1)F_{\mathrm{LL}}(n_{1}) with MM=11 training points and corresponding predicted values from task L1 overlaid. The training data points occur at roughly regular intervals on (0,1](0,1] and the predictions are indistinguishable from reference data to the naked eye. In fact at M=11M=11 the RMS test error for all values of UU is under 10310^{-3} as seen from figure 4(b) where the evolution of the test error is plotted against MM. We see that with the LLQK the test error quickly falls to under 10210^{-2} at M=4M=4 for all of the UU values considered and then continues to fall gradually as more training points are added eventually reaching lower than 10610^{-6}. The RMS test errors obtained by using the Gaussian kernel on task L1 are also plotted in figure 4(b) and show a saturation of the prediction error at around 10210^{-2} over the range of MM investigated. We plot predicted values for FLL(n1)F_{\mathrm{LL}}(n_{1}) obtained from the Gaussian kernel within task L1 for M=19M=19 in figure 4(e). Since the test error is around 10210^{-2} the fit looks good to the naked eye. In order to understand why the error does not seem to improve beyond 10210^{-2} we plot the error distribution over the entire test data set on the n1n_{1} interval (0,1)(0,1) in figure 4(f). We see that the prediction errors are primarily concentrated near n10n_{1}\rightarrow 0 where FLL(n1)F_{\mathrm{LL}}(n_{1}) exhibits a sharp increase in its slope [16] and to a lesser extent near n11n_{1}\rightarrow 1. Thus within the chosen setup for task L1, the Gaussian kernel does not generalize as well near n10n_{1}\rightarrow 0 as it is able to elsewhere. Further, on the interpolative learning task L1, we do not find significant improvements in the prediction error of the Gaussian kernel either by increasing the KRR regularization parameter and allowing for higher training loss or by manually tuning σ\sigma. However, since the Gaussian kernel is a universal kernel [56], we expect to be able to reduce the prediction errors by providing more training data near n10n_{1}\rightarrow 0.

In fig 4(c) we plot training and predicted data points as obtained with the LLQK for FLL(n1)F_{\mathrm{LL}}(n_{1}) within task L2 where the training set is drawn only from the last 40%40\% of the (0,1](0,1] interval. We plot the fit obtained for M=13M=13 as at this training set size the LLQK RMS test error is under 10310^{-3} for all U values considered. Figure 4(d) shows the evolution of the LLQK prediction error as a function of MM for task L2. We see that in extrapolating FLL(n1)F_{\mathrm{LL}}(n_{1}) out to densities far away from the training interval the error falls more gradually compared to task L1 but once a sufficient number of training points are provided it declines further to around 10610^{-6}. Thus in this simple model of the asymmetric Hubbard dimer, the LLQK is able to generalize with high accuracy regardless of how the training and test data are distributed over the domain of n1n_{1}. In contrast, the Gaussian kernel is not expected to generalize with naive KRR to extrapolative test data and especially with the small regularization hyperparameter originally chosen, we see from Figure 4(d) that prediction errors are unacceptably high. We find (not shown) that increasing the KRR regularization parameter on task L2 for the Gaussian kernel from 4×10154\times 10^{-15} to around 10510^{-5} improves prediction errors somewhat from 10010^{0} to around 10110^{-1} at which level they remain as a function of MM.

The observed behavior of the LLQK for the dimer suggests it leads to a restricted model with a low effective dimension in feature space. For M=16M=16 we show the decay of the singular values of the κLL(U)\kappa^{(U)}_{\mathrm{LL}} matrices for different UU in figure 5(a). It is apparent that the magnitudes of the singular values initially decay rapidly suggesting a low effective dimension. Additionally we find that as the size of the training set MM increases, the numerically computed rank of the κLL(U)\kappa^{(U)}_{\mathrm{LL}} matrices stops growing around 16 - 21 irrespective of how the training data is distributed over the domain of n1n_{1}. So even as nominally the 4-qubit system is associated with a 256 dimensional feature space, the Levy-Lieb embedding for the dimer is itself restricted to a small subspace and once enough training points are provided to effectively span important features within this subspace, operator measurements fit to the training data are predictive for test data if the function being predicted is of the form O(U)[nm]=Tr{ρLL(U)[nm]O^}O^{(U)}[n_{m}]=\mathrm{Tr}\{\rho_{\mathrm{LL}}^{(U)}[n_{m}]\hat{O}\}. To emphasise this latter point we consider the transferability of the κLL(U)\kappa^{(U)}_{\mathrm{LL}} kernels for a specified UU in terms of learning density functionals associated with a different interaction strength UU^{\prime}. To this end, firstly in figure 5(b), we show the relative kernel alignment computed using equation 15, between κLL(U)\kappa^{(U)}_{\mathrm{LL}} matrices associated with different UU. Not surprisingly the alignment A(κLL(U),κLL(U))A(\kappa^{(U^{\prime})}_{\mathrm{LL}},\kappa^{(U)}_{\mathrm{LL}}) decreases as |UU||U-U^{\prime}| increases. In figure 5(c) we show the prediction errors associated with an interpolative learning task where we attempt to learn the interacting Levy-Lieb functional FLL(5)(n1)F_{\mathrm{LL}}^{(5)}(n_{1}) using the non-interacting LLQK κLL(0)\kappa^{(0)}_{\mathrm{LL}} and vice versa. We find in this case that the prediction errors level out just above 10210^{-2} as the number of training data points approaches the embedding rank, which is qualitatively different behavior than what we observed for each kernel κLL(U)\kappa^{(U)}_{\mathrm{LL}} when it was used to learn functionals for the same UU.

Refer to caption
Figure 6: (a) Training, predicted and actual values of TLL(n1)T_{\mathrm{LL}}(n_{1}) when learning with the Levy-Lieb quantum kernel in an extrapolative setting. (b) Evolution of the RMS test error as a function of the number of training points for extrapolative learning of TLL(n1)T_{\mathrm{LL}}(n_{1}) using the Levy-Lieb quantum kernel (filled-circles connect by solid lines) for different values of UU (c) Training, predicted and actual values of HXCLL(n1)HXC_{\mathrm{LL}}(n_{1}) when learning with the Levy-Lieb quantum kernel in an interpolative setting. (d) Evolution of the RMS prediction error as a function of the number of training points for interpolative learning of HXCLL(n1)HXC_{\mathrm{LL}}(n_{1}) for different values of UU.

Next, we show that results obtained in terms of learning FLL(n1)F_{\mathrm{LL}}(n_{1}) are not specific to its particular form and other density functionals are expected to behave similarly. Accordingly we consider two additional functionals the first being the kinetic energy functional

TLL(U)[n]=ΨLL(U)[n]|T^|ΨLL(U)[n]T^{(U)}_{\mathrm{LL}}[n]=\langle\Psi_{\mathrm{LL}}^{(U)}[n]|\hat{T}|\Psi_{\mathrm{LL}}^{(U)}[n]\rangle (16)

and the second being the so-called Hartree-Exchange-Correlation (HXCHXC) functional defined as

HXCLL(U)[n]\displaystyle HXC^{(U)}_{\mathrm{LL}}[n] =ΨLL(U)[n]|T^+W^|ΨLL(U)[n]\displaystyle=\langle\Psi_{\mathrm{LL}}^{(U)}[n]|\hat{T}+\hat{W}|\Psi_{\mathrm{LL}}^{(U)}[n]\rangle (17)
ΨLL(0)[n]|T^|ΨLL(0)[n]\displaystyle-\langle\Psi_{\mathrm{LL}}^{(0)}[n]|\hat{T}|\Psi_{\mathrm{LL}}^{(0)}[n]\rangle
=FLL(U)[n]FLL(0)[n]\displaystyle=F_{\mathrm{LL}}^{(U)}[n]-F_{\mathrm{LL}}^{(0)}[n]

The kinetic energy functional TLL(U)[n]T^{(U)}_{\mathrm{LL}}[n] is interesting because even for the dimer, it is a non-monotonic function and its form evolves non-trivially with UU (see Fig. 6(a)). The HXCLL(U)[n]HXC^{(U)}_{\mathrm{LL}}[n] functional is instructive because for the dimer as seen graphically in figure 6(c) it has a relatively simple monotonic behavior as a function of n1n_{1} but in its definition it involves both the interacting ΨLL(U)[n]\Psi_{\mathrm{LL}}^{(U)}[n] and non-interacting ΨLL(0)[n]\Psi_{\mathrm{LL}}^{(0)}[n] states. We consider learning TLL(U)(n1)T^{(U)}_{\mathrm{LL}}(n_{1}) for the dimer in an extrapolative setting similar to task L2 described previously and find as shown in figure 6(a,b) that in spite of its non-monotonic behavior, TLL(U)(n1)T^{(U)}_{\mathrm{LL}}(n_{1}) is easily generalized by κLL(U)\kappa^{(U)}_{\mathrm{LL}} to far away test data with high accuracy as the number of training data points approaches the kernel rank. For HXCLL(U)(n1)HXC^{(U)}_{\mathrm{LL}}(n_{1}) we consider an interpolative setting similar to task L1 above and attempt to learn HXCLL(U)(n1)HXC^{(U)}_{\mathrm{LL}}(n_{1}) using κLL(U)\kappa^{(U)}_{\mathrm{LL}}. We also increase the KRR regularization parameter in this instance to 10610^{-6} to improve generalization behavior. Note that HXCLL(0)(n1)=0HXC^{(0)}_{\mathrm{LL}}(n_{1})=0 for the dimer by definition and is ignored. We find once again that because of the involvement of states with U=0U=0, kernels κLL(U)\kappa^{(U)}_{\mathrm{LL}} with U>0U>0 show a prediction error profile for HXCLL(U)(n1)HXC^{(U)}_{\mathrm{LL}}(n_{1}) that flattens out as a function of MM (see Fig. 6(d)) in sharp contrast to the generalization ability apparent for TLL(U)(n1)T^{(U)}_{\mathrm{LL}}(n_{1}). Furthermore we see that the level at which the error stops improving depends on UU. This illustrates the selective nature of the RKHS associated with the Levy-Lieb embedding of densities.

Our results for the Hubbard dimer suggest that if somehow we only had access to the Levy-Lieb quantum kernel but not the states themselves, we could still calculate density functionals of observables to high accuracy given sufficient training data. Since the dimer is a small system, assuming we have the correct kernel at hand say in the form of the LLQK, it is easy to provide enough training data to exceed the effective dimension of the embedding and achieve very low prediction error. As we do not conduct system size dependent studies in this work, we do not draw conclusions about the scaling of the effective dimension associated with the Levy-Lieb embedding and thus data requirements for learning density functionals. In general such an analysis would have to be conducted in a context specific manner as based on complexity theoretic arguments [12, 5], we do not expect generic quantum advantage for machine learning the universal functional of DFT. The learning abilities of general quantum models of the form f(x)=Tr(O^ρ(x))f(x)=\mathrm{Tr}(\hat{O}\rho(x)) have been analyzed recently by several authors [5, 35, 36, 57, 37]. Huang et al showed that while quantum kernels of the type κQ(x,x)=|x|x|2=Tr(ρ(x)ρ(x))\kappa_{Q}(x,x^{\prime})=|\langle x|x^{\prime}\rangle|^{2}=\mathrm{Tr}(\rho(x)\rho(x^{\prime})) can be expected to learn a very general class of functions, in the worst case, the training data requirements to achieve small prediction errors could be very large. Furthermore recent theoretical works [57, 37] have also discussed the intrinsic weakness of fidelity based kernels with regards to exponentially decaying off-diagonal kernel matrix elements in high dimensions and suggested projected kernels based on reduced density matrices (RDMs). Fortunately for most observables relevant to DFT and materials science mapping densities to feature spaces based on RDMs would also suffice [10]. Additionally with regards to classical machine learning, it should be noted that techniques related to machine learning DFT [8, 58, 59, 30, 60] employing both implicit kernel methods and deep learning are now very mature and results from the Gaussian kernel based KRR used in this work for illustration purposes on the asymmetric Hubbard dimer are not meant to suggest quantum advantage for DFT.

VI Conclusions

We discussed the constrained-search formulation of density functional theory (DFT) within the context of near-term quantum algorithms highlighting the relationship between exact density functionals and underlying wavefunctions. We used parameterized quantum circuits to implement a form of density variational minimization involving run-time calculations of the exact Levy-Lieb functional and illustrated its equivalence to unconstrained wavefunction optimization for obtaining ground states. Interpreting the Levy-Lieb mapping from densities to wavefunctions as a feature space embedding into pure states we demonstrated a quantum kernel that allows us to learn density functionals of observables without obscuring the underlying state space. We hope that our work contributes to improving the explainability of DFT and other reduced variable theories to a broader audience interested in quantum theory.

References

  • Hohenberg [1964] P. Hohenberg, Physical Review 136, B864 (1964).
  • Kohn and Sham [1965] W. Kohn and L. J. Sham, Physical Review 140, A1133 (1965).
  • Noorden et al. [2014] R. V. Noorden, B. Maher,  and R. Nuzzo, Nature News 514, 550 (2014).
  • Parr and Yang [1989] R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules, Vol. 16 (1989) p. 352.
  • Schuch and Verstraete [2009] N. Schuch and F. Verstraete, Nature Physics 5, 732 (2009).
  • Zhang et al. [2018] Y. Zhang, D. A. Kitchaev, J. Yang, T. Chen, S. T. Dacek, R. A. Sarmiento-Pérez, M. A. Marques, H. Peng, G. Ceder, J. P. Perdew,  and J. Sun, npj Computational Materials 4, 1 (2018).
  • Mardirossian and Head-Gordon [2017] N. Mardirossian and M. Head-Gordon, https://doi.org/10.1080/00268976.2017.1333644 115, 2315 (2017).
  • Li et al. [2016] L. Li, T. E. Baker, S. R. White,  and K. Burke, Physical Review B 94 (2016), 10.1103/PhysRevB.94.245129.
  • Capelle and Campo [2013] K. Capelle and V. L. Campo, Physics Reports 528 (2013), 10.1016/j.physrep.2013.03.002.
  • Ludeña et al. [2013] E. V. Ludeña, F. J. Torres,  and C. Costa, Journal of Modern Physics 04, 391 (2013).
  • Gaitan and Nori [2009] F. Gaitan and F. Nori, Physical Review B - Condensed Matter and Materials Physics 79 (2009), 10.1103/PhysRevB.79.205117.
  • Baker and Poulin [2020] T. E. Baker and D. Poulin, Physical Review Research 2 (2020), 10.1103/PhysRevResearch.2.043238.
  • Hatcher et al. [2019] R. Hatcher, J. A. Kittl,  and C. Bowen,   (2019).
  • Senjean et al. [2022] B. Senjean, S. Yalouz,  and M. Saubanère,   (2022).
  • Kohn [1999] W. Kohn, Reviews of Modern Physics 71 (1999), 10.1103/revmodphys.71.1253.
  • Carrascal et al. [2015] D. Carrascal, J. Ferrer, J. C. Smith,  and K. Burke, Journal of Physics: Condensed Matter  (2015), 10.1088/0953-8984/27/39/393001.
  • Carrascal et al. [2018] D. J. Carrascal, J. Ferrer, N. Maitra,  and K. Burke, European Physical Journal B 91 (2018), 10.1140/epjb/e2018-90114-9.
  • Coe and D’Amico [2010] J. P. Coe and I. D’Amico (2010).
  • Cioslowski [1988] J. Cioslowski, Physical Review Letters 60 (1988), 10.1103/PhysRevLett.60.2141.
  • Cioslowski [1989] J. Cioslowski, International Journal of Quantum Chemistry 36 (1989), 10.1002/qua.560360829.
  • Cioslowski [1991] J. Cioslowski, Physical Review A 43 (1991), 10.1103/PhysRevA.43.1223.
  • Capelle [2003] K. Capelle, Journal of Chemical Physics 119 (2003), 10.1063/1.1593014.
  • D’Amico et al. [2011] I. D’Amico, J. P. Coe, V. V. França,  and K. Capelle, Physical Review Letters 106 (2011), 10.1103/PhysRevLett.106.050401.
  • Mori-Sánchez and Cohen [2018] P. Mori-Sánchez and A. J. Cohen, Journal of Physical Chemistry Letters 9, 4910 (2018).
  • Levy [1979] M. Levy, Physics 76, 6062 (1979).
  • Levy [1982] M. Levy, Physical Review A 26 (1982), 10.1103/PhysRevA.26.1200.
  • Lieb [1983] E. H. Lieb, International Journal of Quantum Chemistry 24 (1983), 10.1002/qua.560240302.
  • Levy and Perdew [1985] M. Levy and J. P. Perdew, The Constrained Search Formulation of Density Functional Theory (1985).
  • Zhao and Parr [1993] Q. Zhao and R. G. Parr, The Journal of Chemical Physics 98 (1993), 10.1063/1.465093.
  • von Lilienfeld and Burke [2020] O. A. von Lilienfeld and K. Burke, Nature Communications 11, 10 (2020).
  • Eschrig [2003] H. H. Eschrig,  , 226 (2003).
  • Peruzzo et al. [2014] A. Peruzzo, J. McClean, P. Shadbolt, M. H. Yung, X. Q. Zhou, P. J. Love, A. Aspuru-Guzik,  and J. L. O’Brien, Nature Communications 5 (2014), 10.1038/ncomms5213.
  • Tilly et al. [2021] J. Tilly, H. Chen, S. Cao, D. Picozzi, K. Setia, Y. Li, E. Grant, L. Wossnig, I. Rungger, G. H. Booth,  and J. Tennyson,   (2021).
  • Havlíček et al. [2019] V. Havlíček, A. D. Córcoles, K. Temme, A. W. Harrow, A. Kandala, J. M. Chow,  and J. M. Gambetta, Nature 567 (2019), 10.1038/s41586-019-0980-2.
  • Schuld and Killoran [2019] M. Schuld and N. Killoran, Physical Review Letters 122 (2019), 10.1103/PhysRevLett.122.040504.
  • Schuld and Petruccione [2021] M. Schuld and F. Petruccione,   (2021).
  • Kübler et al. [2021] J. M. Kübler, S. Buchholz,  and B. Schölkopf,   (2021).
  • ANIS et al. [2021] M. S. ANIS, Abby-Mitchell, H. Abraham, AduOffei, R. Agarwal, G. Agliardi, M. Aharoni, V. Ajith, I. Y. Akhalwaya, G. Aleksandrowicz, T. Alexander, M. Amy, S. Anagolum, Anthony-Gandon, E. Arbel, A. Asfaw, A. Athalye, A. Avkhadiev, C. Azaustre, P. BHOLE, A. Banerjee, S. Banerjee, W. Bang, A. Bansal, P. Barkoutsos, A. Barnawal, G. Barron, G. S. Barron, L. Bello, Y. Ben-Haim, M. C. Bennett, D. Bevenius, D. Bhatnagar, P. Bhatnagar, A. Bhobe, P. Bianchini, L. S. Bishop, C. Blank, S. Bolos, S. Bopardikar, S. Bosch, S. Brandhofer, Brandon, S. Bravyi, N. Bronn, Bryce-Fuller, D. Bucher, A. Burov, F. Cabrera, P. Calpin, L. Capelluto, J. Carballo, G. Carrascal, A. Carriker, I. Carvalho, A. Chen, C.-F. Chen, E. Chen, J. C. Chen, R. Chen, F. Chevallier, K. Chinda, R. Cholarajan, J. M. Chow, S. Churchill, CisterMoke, C. Claus, C. Clauss, C. Clothier, R. Cocking, R. Cocuzzo, J. Connor, F. Correa, Z. Crockett, A. J. Cross, A. W. Cross, S. Cross, J. Cruz-Benito, C. Culver, A. D. Córcoles-Gonzales, N. D, S. Dague, T. E. Dandachi, A. N. Dangwal, J. Daniel, M. Daniels, M. Dartiailh, A. R. Davila, F. Debouni, A. Dekusar, A. Deshmukh, M. Deshpande, D. Ding, J. Doi, E. M. Dow, P. Downing, E. Drechsler, E. Dumitrescu, K. Dumon, I. Duran, K. EL-Safty, E. Eastman, G. Eberle, A. Ebrahimi, P. Eendebak, D. Egger, ElePT, Emilio, A. Espiricueta, M. Everitt, D. Facoetti, Farida, P. M. Fernández, S. Ferracin, D. Ferrari, A. H. Ferrera, R. Fouilland, A. Frisch, A. Fuhrer, B. Fuller, M. GEORGE, J. Gacon, B. G. Gago, C. Gambella, J. M. Gambetta, A. Gammanpila, L. Garcia, T. Garg, S. Garion, J. R. Garrison, J. Garrison, T. Gates, H. Georgiev, L. Gil, A. Gilliam, A. Giridharan, Glen, J. Gomez-Mosquera, Gonzalo, S. de la Puente González, J. Gorzinski, I. Gould, D. Greenberg, D. Grinko, W. Guan, D. Guijo, J. A. Gunnels, H. Gupta, N. Gupta, J. M. Günther, M. Haglund, I. Haide, I. Hamamura, O. C. Hamido, F. Harkins, K. Hartman, A. Hasan, V. Havlicek, J. Hellmers, Ł. Herok, S. Hillmich, H. Horii, C. Howington, S. Hu, W. Hu, C.-H. Huang, J. Huang, R. Huisman, H. Imai, T. Imamichi, K. Ishizaki, Ishwor, R. Iten, T. Itoko, A. Ivrii, A. Javadi, A. Javadi-Abhari, W. Javed, Q. Jianhua, M. Jivrajani, K. Johns, S. Johnstun, Jonathan-Shoemaker, JosDenmark, JoshDumo, J. Judge, T. Kachmann, A. Kale, N. Kanazawa, J. Kane, Kang-Bae, A. Kapila, A. Karazeev, P. Kassebaum, T. Kehrer, J. Kelso, S. Kelso, H. van Kemenade, V. Khanderao, S. King, Y. Kobayashi, Kovi11Day, A. Kovyrshin, R. Krishnakumar, P. Krishnamurthy, V. Krishnan, K. Krsulich, P. Kumkar, G. Kus, R. LaRose, E. Lacal, R. Lambert, H. Landa, J. Lapeyre, J. Latone, S. Lawrence, C. Lee, G. Li, T. J. Liang, J. Lishman, D. Liu, P. Liu, Lolcroc, A. K. M, L. Madden, Y. Maeng, S. Maheshkar, K. Majmudar, A. Malyshev, M. E. Mandouh, J. Manela, Manjula, J. Marecek, M. Marques, K. Marwaha, D. Maslov, P. Maszota, D. Mathews, A. Matsuo, F. Mazhandu, D. McClure, M. McElaney, C. McGarry, D. McKay, D. McPherson, S. Meesala, D. Meirom, C. Mendell, T. Metcalfe, M. Mevissen, A. Meyer, A. Mezzacapo, R. Midha, D. Miller, H. Miller, Z. Minev, A. Mitchell, N. Moll, A. Montanez, G. Monteiro, M. D. Mooring, R. Morales, N. Moran, D. Morcuende, S. Mostafa, M. Motta, R. Moyard, P. Murali, D. Murata, J. Müggenburg, T. NEMOZ, D. Nadlinger, K. Nakanishi, G. Nannicini, P. Nation, E. Navarro, Y. Naveh, S. W. Neagle, P. Neuweiler, A. Ngoueya, T. Nguyen, J. Nicander, Nick-Singstock, P. Niroula, H. Norlen, NuoWenLei, L. J. O’Riordan, O. Ogunbayo, P. Ollitrault, T. Onodera, R. Otaolea, S. Oud, D. Padilha, H. Paik, S. Pal, Y. Pang, A. Panigrahi, V. R. Pascuzzi, S. Perriello, E. Peterson, A. Phan, K. Pilch, F. Piro, M. Pistoia, C. Piveteau, J. Plewa, P. Pocreau, A. Pozas-Kerstjens, R. Pracht, M. Prokop, V. Prutyanov, S. Puri, D. Puzzuoli, Pythonix, J. Pérez, Quant02, Quintiii, R. I. Rahman, A. Raja, R. Rajeev, I. Rajput, N. Ramagiri, A. Rao, R. Raymond, O. Reardon-Smith, R. M.-C. Redondo, M. Reuter, J. Rice, M. Riedemann, Rietesh, D. Risinger, P. Rivero, M. L. Rocca, D. M. Rodríguez, RohithKarur, B. Rosand, M. Rossmannek, M. Ryu, T. SAPV, N. R. C. Sa, A. Saha, A. Ash-Saki, S. Sanand, M. Sandberg, H. Sandesara, R. Sapra, H. Sargsyan, A. Sarkar, N. Sathaye, N. Savola, B. Schmitt, C. Schnabel, Z. Schoenfeld, T. L. Scholten, E. Schoute, M. Schulterbrandt, J. Schwarm, J. Seaward, Sergi, I. F. Sertage, K. Setia, F. Shah, N. Shammah, W. Shanks, R. Sharma, Y. Shi, J. Shoemaker, A. Silva, A. Simonetto, D. Singh, D. Singh, P. Singh, P. Singkanipa, Y. Siraichi, Siri, J. Sistos, I. Sitdikov, S. Sivarajah, Slavikmew, M. B. Sletfjerding, J. A. Smolin, M. Soeken, I. O. Sokolov, I. Sokolov, V. P. Soloviev, SooluThomas, Starfish, D. Steenken, M. Stypulkoski, A. Suau, S. Sun, K. J. Sung, M. Suwama, O. Słowik, H. Takahashi, T. Takawale, I. Tavernelli, C. Taylor, P. Taylour, S. Thomas, K. Tian, M. Tillet, M. Tod, M. Tomasik, C. Tornow, E. de la Torre, J. L. S. Toural, K. Trabing, M. Treinish, D. Trenev, TrishaPe, F. Truger, G. Tsilimigkounakis, D. Tulsi, D. Tuna, W. Turner, Y. Vaknin, C. R. Valcarce, F. Varchon, A. Vartak, A. C. Vazquez, P. Vijaywargiya, V. Villar, B. Vishnu, D. Vogt-Lee, C. Vuillot, J. Weaver, J. Weidenfeller, R. Wieczorek, J. A. Wildstrom, J. Wilson, E. Winston, WinterSoldier, J. J. Woehr, S. Woerner, R. Woo, C. J. Wood, R. Wood, S. Wood, J. Wootton, M. Wright, L. Xing, J. YU, B. Yang, U. Yang, J. Yao, D. Yeralin, R. Yonekura, D. Yonge-Mallo, R. Yoshida, R. Young, J. Yu, L. Yu, Yuma-Nakamura, C. Zachow, L. Zdanski, H. Zhang, I. Zidaru, B. Zimmermann, C. Zoufal, aeddins ibm, alexzhang13, b63, bartek bartlomiej, bcamorrison, brandhsn, chetmurthy, deeplokhande, dekel.meirom, dime10, dlasecki, ehchen, ewinston, fanizzamarco, fs1132429, gadial, galeinston, georgezhou20, georgios ts, gruu, hhorii, hhyap, hykavitha, itoko, jeppevinkel, jessica angel7, jezerjojo14, jliu45, johannesgreiner, jscott2, klinvill, krutik2966, ma5x, michelle4654, msuwama, nico lgrs, nrhawkins, ntgiwsvp, ordmoj, sagar pahwa, pritamsinha2304, rithikaadiga, ryancocuzzo, saktar unr, saswati qiskit, septembrr, sethmerkel, sg495, shaashwat, smturro2, sternparky, strickroman, tigerjack, tsura crisaldo, upsideon, vadebayo49, welien, willhbang, wmurphy collabstar, yang.luh,  and M. Čepulkovskis, “Qiskit: An open-source framework for quantum computing,”  (2021).
  • Capelle et al. [2007] K. Capelle, C. A. Ullrich,  and G. Vignale, Physical Review A - Atomic, Molecular, and Optical Physics 76 (2007), 10.1103/PhysRevA.76.012508.
  • Gonis et al. [2016] A. Gonis, X.-G. Zhang, M. Däne, G. M. Stocks,  and D. M. Nicholson, Journal of Physics and Chemistry of Solids 89, 23 (2016).
  • Ryabinkin et al. [2019] I. G. Ryabinkin, S. N. Genin,  and A. F. Izmaylov, Journal of Chemical Theory and Computation 15 (2019), 10.1021/acs.jctc.8b00943.
  • Kuroiwa and Nakagawa [2021] K. Kuroiwa and Y. O. Nakagawa, Physical Review Research 3 (2021), 10.1103/PhysRevResearch.3.013197.
  • Kohn [1983] W. Kohn, Physical Review Letters 51, 17 (1983).
  • Chayes et al. [1985] J. T. Chayes, L. Chayes,  and M. B. Ruskai, Journal of Statistical Physics 38 (1985).
  • Jordan and Wigner [1928] P. Jordan and E. Wigner, Zeitschrift für Physik 47 (1928), 10.1007/BF01331938.
  • Lee et al. [2019] J. Lee, W. J. Huggins, M. Head-Gordon,  and K. B. Whaley, Journal of Chemical Theory and Computation 15, 311 (2019).
  • Zhu et al. [1997] C. Zhu, R. H. Byrd, P. Lu,  and J. Nocedal, ACM Transactions on Mathematical Software 23 (1997), 10.1145/279232.279236.
  • Virtanen et al. [2020] P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wilson, K. J. Millman, N. Mayorov, A. R. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, İlhan Polat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen, E. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, P. van Mulbregt, A. Vijaykumar, A. P. Bardelli, A. Rothberg, A. Hilboll, A. Kloeckner, A. Scopatz, A. Lee, A. Rokem, C. N. Woods, C. Fulton, C. Masson, C. Häggström, C. Fitzgerald, D. A. Nicholson, D. R. Hagen, D. V. Pasechnik, E. Olivetti, E. Martin, E. Wieser, F. Silva, F. Lenders, F. Wilhelm, G. Young, G. A. Price, G. L. Ingold, G. E. Allen, G. R. Lee, H. Audren, I. Probst, J. P. Dietrich, J. Silterra, J. T. Webber, J. Slavič, J. Nothman, J. Buchner, J. Kulick, J. L. Schönberger, J. V. de Miranda Cardoso, J. Reimer, J. Harrington, J. L. C. Rodríguez, J. Nunez-Iglesias, J. Kuczynski, K. Tritz, M. Thoma, M. Newville, M. Kümmerer, M. Bolingbroke, M. Tartre, M. Pak, N. J. Smith, N. Nowaczyk, N. Shebanov, O. Pavlyk, P. A. Brodtkorb, P. Lee, R. T. McGibbon, R. Feldbauer, S. Lewis, S. Tygier, S. Sievert, S. Vigna, S. Peterson, S. More, T. Pudlik, T. Oshima, T. J. Pingel, T. P. Robitaille, T. Spura, T. R. Jones, T. Cera, T. Leslie, T. Zito, T. Krauss, U. Upadhyay, Y. O. Halchenko,  and Y. Vázquez-Baeza, Nature Methods 17 (2020), 10.1038/s41592-019-0686-2.
  • Kraft [1988] D. Kraft, Technical Report DFVLR-FB 88 (1988).
  • Bittel and Kliesch [2021] L. Bittel and M. Kliesch,   (2021).
  • Higgott et al. [2019] O. Higgott, D. Wang,  and S. Brierley, Quantum 3 (2019), 10.22331/q-2019-07-01-156.
  • Pedregosa et al. [2011] F. Pedregosa, G. Varoquaux, A. Gramfort, V. Michel, B. Thirion, O. Grisel, M. Blondel, P. Prettenhofer, R. Weiss, V. Dubourg, J. Vanderplas, A. Passos, D. Cournapeau, M. Brucher, M. Perrot,  and E. Duchesnay, Journal of Machine Learning Research 12, 2825 (2011).
  • Hubregtsen et al. [2021] T. Hubregtsen, D. Wierichs, E. Gil-Fuster, P.-J. H. S. Derks, P. K. Faehrmann,  and J. J. Meyer,   (2021).
  • Cristianini et al. [2001] N. Cristianini, J. Shawe-Taylor, A. Elisseeff,  and J. Kandola (MIT Press, 2001).
  • Kandola et al. [2002] J. Kandola, J. Shawe-Taylor,  and N. Cristianini, Technical Report  (2002).
  • Micchelli et al. [2006] C. A. Micchelli, Y. Xu,  and H. Zhang, Journal of Machine Learning Research 7 (2006).
  • Huang et al. [2021] H. Y. Huang, M. Broughton, M. Mohseni, R. Babbush, S. Boixo, H. Neven,  and J. R. McClean, Nature Communications 12 (2021), 10.1038/s41467-021-22539-9.
  • Nelson et al. [2019] J. Nelson, R. Tiwari,  and S. Sanvito, Physical Review B 99, 075132 (2019).
  • Moreno et al. [2020] J. R. Moreno, G. Carleo,  and A. Georges, Physical Review Letters 125 (2020), 10.1103/PhysRevLett.125.076402.
  • Kirkpatrick et al. [2021] J. Kirkpatrick, B. McMorrow, D. H. Turban, A. L. Gaunt, J. S. Spencer, A. G. Matthews, A. Obika, L. Thiry, M. Fortunato, D. Pfau, L. R. Castellanos, S. Petersen, A. W. Nelson, P. Kohli, P. Mori-Sánchez, D. Hassabis,  and A. J. Cohen, Science 374 (2021), 10.1126/science.abj6511.