License: CC BY 4.0
arXiv:2403.08625v1 [quant-ph] 13 Mar 2024

Variance Minimisation of the Lipkin-Meshkov-Glick Model
on a Quantum Computer

I. Hobday School of Mathematics and Physics, University of Surrey,
Guildford, Surrey, GU2 7XH, United Kingdom
   P. D. Stevenson School of Mathematics and Physics, University of Surrey,
Guildford, Surrey, GU2 7XH, United Kingdom
AWE, Aldermaston, Berkshire, RG7 4PR, United Kingdom
   J. Benstead AWE, Aldermaston, Berkshire, RG7 4PR, United Kingdom School of Mathematics and Physics, University of Surrey,
Guildford, Surrey, GU2 7XH, United Kingdom
(March 13, 2024)
Abstract

Quantum computing can potentially provide advantages for specific computational tasks. The simulation of fermionic systems is one such task that lends itself well to quantum computation, with applications in nuclear physics and electronic systems. Here we present work in which we use a variance minimisation method to find the full spectrum of energy eigenvalues of the Lipkin-Meshkov-Glick model; an exactly-solvable nuclear shell model-type system.

We perform these calculations using both quantum simulators and real quantum hardware accessed via IBM cloud-based quantum computers. Using these IBM quantum computers we are able to obtain all eigenvalues for the cases of three and seven fermions (nucleons) in the Lipkin-Meshkov-Glick model.

I Introduction

The use of quantum algorithms to perform calculations in nuclear physics is a rapidly develo** field [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21]. The exponential scaling of Hilbert space with the number of quantum bits (qubits) and the ability of multiple qubits to exhibit highly entangled wave functions give quantum computers the potential to have a great impact in simulating many-body quantum systems. In particular, one can expect a quantum advantage where quantum computation outperforms classical computation as the size of the system under study becomes sufficiently large.

Richard Feynman first proposed using quantum computers to study quantum systems in the 1980s [22]. Since then, various algorithms have been developed that are able to perform calculations on many-body quantum systems, such as Quantum Phase Estimation (QPE) [23, 24, 25], and Quantum Imaginary Time Evolution [26, 27, 28]. However, these algorithms are often too complex to be fully implemented on current quantum computers.

Current quantum computers are said to be in the Noisy, Intermediate-Scale Quantum (NISQ) era. This is due to the low numbers of qubits, and large amounts of noise and error present within the devices. Current devices also have low coherence times, restricting the amount of time a qubit can maintain its state. This limits the length of the quantum circuit, or circuit depth, that quantum computers can use and still produce meaningful results [29].

I.1 Variational Algorithms

Variational quantum algorithms have established themselves due to the relative ease of running them on NISQ-era quantum computers [30, 31, 32]. These are hybrid algorithms, meaning they use both quantum and classical processes in tandem, reducing the amount of computational work needed to be performed on the quantum computer. Working in conjunction with a classical computer allows for reduced circuit depths and improved error rates by reducing the number of gates in the quantum circuit. This reduction in error allows for meaningful results to be obtained on current quantum hardware.

The Variational Quantum Eigensolver (VQE) algorithm uses the variational principle of quantum mechanics to approximate the ground state of some Hamiltonian. In the work presented here, we apply a version of the VQE which targets any eigenvalue of a quantum system by minimizing the variance of the Hamiltonian rather than the expectation value. Application of the method is made to the simplified nuclear model due to Lipkin, Meshkov, and Glick, which we summarize in the next section. The presentation proceeds with a description of our implementation of the algorithm, followed by results on simulated and real quantum computers.

I.2 Lipkin-Meshkov-Glick Model

The Lipkin-Meshkov-Glick (LMG) model is an exactly solvable nuclear model introduced in the 1960s [33]. It is a simple shell model consisting of two levels, separated by an energy ϵitalic-ϵ\epsilonitalic_ϵ with a permutation-symmetric potential. The two levels are each N𝑁Nitalic_N-fold degenerate and a standard treatment (which we adopt here) considers N𝑁Nitalic_N fermions in the system. The Hamiltonian for the model is given as

H=𝐻absent\displaystyle H=italic_H = 12ϵpσσapσapσ+12Vppσap,σap,σap,σap,σ12italic-ϵsubscript𝑝𝜎𝜎superscriptsubscript𝑎𝑝𝜎subscript𝑎𝑝𝜎12𝑉subscript𝑝superscript𝑝𝜎superscriptsubscript𝑎𝑝𝜎superscriptsubscript𝑎superscript𝑝𝜎subscript𝑎superscript𝑝𝜎subscript𝑎𝑝𝜎\displaystyle\frac{1}{2}\epsilon\sum_{p\sigma}\sigma a_{p\sigma}^{\dagger}a_{p% \sigma}+\frac{1}{2}V\sum_{pp^{\prime}\sigma}a_{p,\sigma}^{\dagger}a_{p^{\prime% },\sigma}^{\dagger}a_{p^{\prime},-\sigma}a_{p,-\sigma}divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ϵ ∑ start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT italic_σ italic_a start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_V ∑ start_POSTSUBSCRIPT italic_p italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_σ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_p , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , - italic_σ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_p , - italic_σ end_POSTSUBSCRIPT (1)
+12Wppσap,σap,σap,σap,σ.12𝑊subscript𝑝superscript𝑝𝜎superscriptsubscript𝑎𝑝𝜎superscriptsubscript𝑎superscript𝑝𝜎subscript𝑎superscript𝑝𝜎subscript𝑎𝑝𝜎\displaystyle+\frac{1}{2}W\sum_{pp^{\prime}\sigma}a_{p,\sigma}^{\dagger}a_{p^{% \prime},-\sigma}^{\dagger}a_{p^{\prime},\sigma}a_{p,-\sigma}.+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_W ∑ start_POSTSUBSCRIPT italic_p italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_σ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_p , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , - italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_σ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_p , - italic_σ end_POSTSUBSCRIPT .

Here the labels p𝑝pitalic_p and psuperscript𝑝p^{\prime}italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT run over the N𝑁Nitalic_N degenerate states in each level, while σ=±1𝜎plus-or-minus1\sigma=\pm 1italic_σ = ± 1 labels the level. The strength V𝑉Vitalic_V controls pair (de-)excitations between the two levels, while the W𝑊Witalic_W term scatters one particle up and another down. The model can be simplified by writing the Hamiltonian in the quasi-spin basis, as

H=ϵJz+12V(J+2+J2)+12W(J+J+JJ+),𝐻italic-ϵsubscript𝐽𝑧12𝑉superscriptsubscript𝐽2superscriptsubscript𝐽212𝑊subscript𝐽subscript𝐽subscript𝐽subscript𝐽H=\epsilon J_{z}+\frac{1}{2}V(J_{+}^{2}+J_{-}^{2})+\frac{1}{2}W(J_{+}J_{-}+J_{% -}J_{+}),italic_H = italic_ϵ italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_V ( italic_J start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_J start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_W ( italic_J start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT - end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) , (2)

where Jzsubscript𝐽𝑧J_{z}italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and J±subscript𝐽plus-or-minusJ_{\pm}italic_J start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT are given by

Jz=12pσσapσapσsubscript𝐽𝑧12subscript𝑝𝜎𝜎superscriptsubscript𝑎𝑝𝜎subscript𝑎𝑝𝜎J_{z}=\frac{1}{2}\sum_{p\sigma}\sigma a_{p\sigma}^{\dagger}a_{p\sigma}italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT italic_σ italic_a start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p italic_σ end_POSTSUBSCRIPT (3)

and

J±=pap±1ap1subscript𝐽plus-or-minussubscript𝑝superscriptsubscript𝑎plus-or-minus𝑝1subscript𝑎minus-or-plus𝑝1J_{\pm}=\sum_{p}a_{p\pm 1}^{\dagger}a_{p\mp 1}italic_J start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_p ± 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_p ∓ 1 end_POSTSUBSCRIPT (4)

respectively.

These quasi-spin operators satisfy angular momentum commutation relations. The total quasi-spin operator J2=12(J+J+JJ+)+Jz2superscript𝐽212subscript𝐽subscript𝐽subscript𝐽subscript𝐽superscriptsubscript𝐽𝑧2J^{2}=\frac{1}{2}(J_{+}J_{-}+J_{-}J_{+})+J_{z}^{2}italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_J start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT - end_POSTSUBSCRIPT + italic_J start_POSTSUBSCRIPT - end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) + italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT commutes with the Hamiltonian, so that each J𝐽Jitalic_J value may be considered separately. This symmetry allows one to reduce the maximum Hamiltonian matrix dimension from 2Nsuperscript2𝑁2^{N}2 start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT to N+1𝑁1N+1italic_N + 1. Further simplification is possible since the Hamiltonian also exhibits good “parity” [34], which splits each matrix of a given J𝐽Jitalic_J into two block-diagonal parts.

The LMG model is applicable as a test-platform for quantum computing algorithms applied to nuclear physics due to its exactly solvable nature, and the ability to scale the model indefinitely by increasing the number of particles in the system. The model has already been studied using real and simulated quantum computers, particularly for calculation of its ground state [13, 16], but also for excited states using a quantum-assisted algorithm to prepare a generalised eigenvalue problem for solution on a classical computer for the excited state spectrum [17]. For our work, we use an LMG model Hamiltonian with the parameters Vϵ=0.5𝑉italic-ϵ0.5\frac{V}{\epsilon}=0.5divide start_ARG italic_V end_ARG start_ARG italic_ϵ end_ARG = 0.5, Wϵ=0𝑊italic-ϵ0\frac{W}{\epsilon}=0divide start_ARG italic_W end_ARG start_ARG italic_ϵ end_ARG = 0, making the common choice to ignore the W𝑊Witalic_W term which does not affect ground state correlations.

II Encoding Methodology

The LMG-model Hamiltonian must be encoded in such a way as to allow it to be processed by a quantum computer. For circuit-based quantum computers, this form is conveniently expressed as linear combinations of Pauli spin matrices. Often a representation of a Hamiltonian in second-quantized form, such as Eq. (1), leads to a map** from creation and annihilation operators to Pauli strings, kee** the second quantized representation where solutions of any particle number can be encoded. Such encodings as Jordan-Wigner [35] or Bravyi-Kitaev [36, 37] fall in this category. Having made use of the quasispin representation, and specialising to the standard Lipkin Model in which the number of particles is fixed at N𝑁Nitalic_N - the same as the degeneracy of each of the two unperturbed levels - we make a more specialised and efficient encoding starting from a representation of the Hamiltonian in the quasispin basis [13] in which only fixed-particle-number states are considered.

For a given specific LMG Hamiltonian, HNsubscript𝐻𝑁H_{N}italic_H start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT with fixed N𝑁Nitalic_N we make a matrix representation in the quasispin basis and consider (as usual) the maximum quasispin value, that being the one containing the unperturbed ground state. The matrix dimension is N+1𝑁1N+1italic_N + 1, but owing to the parity symmetry, it can be written as two block diagonal parts which can be treated independently. When N𝑁Nitalic_N is odd, the two submatrices are of equal dimension, (N+1)/2𝑁12(N+1)/2( italic_N + 1 ) / 2. In the present work we consider N=3,7𝑁37N=3,7italic_N = 3 , 7 so that (N+1)/2𝑁12(N+1)/2( italic_N + 1 ) / 2 is of the form 2msuperscript2𝑚2^{m}2 start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT and each matrix whose eigenvalues are sought has a power-of-two dimension. For a Hamiltonian HNsubscript𝐻𝑁H_{N}italic_H start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT for which 2m=(N+1)/2superscript2𝑚𝑁122^{m}=(N+1)/22 start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT = ( italic_N + 1 ) / 2 we can decompose in terms of Pauli matrices as

HN=i=022m1βiq=1mσi4[q]subscript𝐻𝑁superscriptsubscript𝑖0superscript22𝑚1subscript𝛽𝑖superscriptsubscripttensor-product𝑞1𝑚subscript𝜎subscript𝑖4delimited-[]𝑞H_{N}=\sum_{i=0}^{2^{2m}-1}\beta_{i}\bigotimes_{q=1}^{m}\sigma_{{i_{4}}[q]}italic_H start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 start_POSTSUPERSCRIPT 2 italic_m end_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⨂ start_POSTSUBSCRIPT italic_q = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT [ italic_q ] end_POSTSUBSCRIPT (5)

where βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is a coefficient and the tensor product tensor-product\bigotimes produces all Pauli strings from the set {σ0,σ1,σ2,σ3}={I,X,Y,Z}subscript𝜎0subscript𝜎1subscript𝜎2subscript𝜎3𝐼𝑋𝑌𝑍\{\sigma_{0},\sigma_{1},\sigma_{2},\sigma_{3}\}=\{I,X,Y,Z\}{ italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT } = { italic_I , italic_X , italic_Y , italic_Z } and the notation σi4[q]subscript𝜎subscript𝑖4delimited-[]𝑞\sigma_{{i_{4}}[q]}italic_σ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT [ italic_q ] end_POSTSUBSCRIPT means the qthsuperscript𝑞𝑡q^{th}italic_q start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT digit from the right of i𝑖iitalic_i when represented in base 4.

For the given numerical Hamiltonians with particular values for the interaction strengths V/ϵ𝑉italic-ϵV/\epsilonitalic_V / italic_ϵ and W/ϵ𝑊italic-ϵW/\epsilonitalic_W / italic_ϵ, the coefficients βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT were found using a custom code [38], whose functionality is also found in common quantum computing libraries (e.g. the pauli_decompose method of PennyLane [39]). Specific examples of the encodings are found in subsequent sections.

III Finding Excited States

In a typical Variational Quantum Algorithm (VQA), an ansatz is made for a wave function representing the ground state of the problem at hand. The ansatz contains parameters which can be adjusted to optimise the wave function and bring it as close as possible to the true ground state. Usually the expectation value of the Hamiltonian, expressed in terms of Pauli operators, is evaluated on the quantum computer, while a classical algorithm manages the parameter optimisation.

Here, we adapt the typical VQA to seek the minimum of the variance of the Hamiltonian,

σ2=H2H2.superscript𝜎2delimited-⟨⟩superscript𝐻2superscriptdelimited-⟨⟩𝐻2\sigma^{2}=\langle H^{2}\rangle-\langle H\rangle^{2}.italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ⟨ italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ - ⟨ italic_H ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (6)

The variance is a positive semi-definite function which is zero when the wave function is an eigenfunction of the Hamiltonian. Hence, its set of equally deep global minima will correspond to the set of eigenstates of the Hamiltonian, except possibly for accidental zeros of the variance which can be checked for.

This method allows the use of the same quantum circuit ansatzes as when using the VQE to find the ground state of H𝐻Hitalic_H directly, but requires additional circuit measurements to be performed for the terms in H2delimited-⟨⟩superscript𝐻2\langle H^{2}\rangle⟨ italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩. This increases the time taken to perform calculations but does not introduce any additional circuit depth or variational parameters to the calculations compared with an energy-minimizing ground state VQE. Since no additional circuit complexity is introduced, using the variance to find excited states of the nuclear system is possible on NISQ-era quantum hardware for problems where the standard VQE is applicable.

In our implementation, we make use of the IBM qiskit environment to perform our VQE on real quantum hardware, as noted later where individual results are presented, using qiskit’s classical COBYLA solver for the ansatz parameter optimisation.

IV N=3𝑁3N=3italic_N = 3 Simulation and Measurement

IV.1 N=3𝑁3N=3italic_N = 3 Hamiltonian

The Hamiltonian for the N=3𝑁3N=3italic_N = 3 LMG model is split into its two parity submatrices labelled as A𝐴Aitalic_A and B𝐵Bitalic_B. For our chosen values of V/ϵ=0.5𝑉italic-ϵ0.5V/\epsilon=0.5italic_V / italic_ϵ = 0.5 and W/ϵ=0𝑊italic-ϵ0W/\epsilon=0italic_W / italic_ϵ = 0 the submatrices can be expressed as

HN=3A=[1.50.8660.8660.5]subscript𝐻𝑁3𝐴matrix1.50.8660.8660.5H_{N=3A}=\begin{bmatrix}-1.5&-0.866\\ -0.866&0.5\end{bmatrix}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT = [ start_ARG start_ROW start_CELL - 1.5 end_CELL start_CELL - 0.866 end_CELL end_ROW start_ROW start_CELL - 0.866 end_CELL start_CELL 0.5 end_CELL end_ROW end_ARG ] (7)

and

HN=3B=[0.50.8660.8661.5].subscript𝐻𝑁3𝐵matrix0.50.8660.8661.5H_{N=3B}=\begin{bmatrix}-0.5&-0.866\\ -0.866&1.5\end{bmatrix}.italic_H start_POSTSUBSCRIPT italic_N = 3 italic_B end_POSTSUBSCRIPT = [ start_ARG start_ROW start_CELL - 0.5 end_CELL start_CELL - 0.866 end_CELL end_ROW start_ROW start_CELL - 0.866 end_CELL start_CELL 1.5 end_CELL end_ROW end_ARG ] . (8)

When decomposed into the form of Pauli matrices, the Hamiltonians can be represented as

HN=3A=0.51.0Z00.8660254X0subscript𝐻𝑁3𝐴0.51.0subscript𝑍00.8660254subscript𝑋0H_{N=3A}=-0.5-1.0Z_{0}-0.8660254X_{0}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT = - 0.5 - 1.0 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 0.8660254 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (9)

and

HN=3B=0.51.0Z00.8660254X0subscript𝐻𝑁3𝐵0.51.0subscript𝑍00.8660254subscript𝑋0H_{N=3B}=0.5-1.0Z_{0}-0.8660254X_{0}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_B end_POSTSUBSCRIPT = 0.5 - 1.0 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 0.8660254 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (10)

respectively, where we have kept more significant figures than given in the matrix representation.

We note that the two submatrices differ only by the signs of all the diagonal elements, while the off-diagonal elements are indetical. This means that the eigenspectra of the two submatrices differ only by their sign. This is a general result (for odd N𝑁Nitalic_N) and one needs only compute the spectrum of one of the submatrices in order to know the spectrum for both, but we test the method in the N=3 case for both submatrices.

For the evaluation of the variance, Eq. (6), the square of the Hamiltonian operator is needed. From the square of the matrix representations

HN=3A2=[1.00.8660.8663.0]subscriptsuperscript𝐻2𝑁3𝐴matrix1.00.8660.8663.0H^{2}_{N=3A}=\begin{bmatrix}1.0&0.866\\ 0.866&3.0\end{bmatrix}italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT = [ start_ARG start_ROW start_CELL 1.0 end_CELL start_CELL 0.866 end_CELL end_ROW start_ROW start_CELL 0.866 end_CELL start_CELL 3.0 end_CELL end_ROW end_ARG ] (11)

and

HN=3B2=[1.00.8660.8663.0]superscriptsubscript𝐻𝑁3𝐵2matrix1.00.8660.8663.0H_{N=3B}^{2}=\begin{bmatrix}1.0&-0.866\\ -0.866&3.0\end{bmatrix}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = [ start_ARG start_ROW start_CELL 1.0 end_CELL start_CELL - 0.866 end_CELL end_ROW start_ROW start_CELL - 0.866 end_CELL start_CELL 3.0 end_CELL end_ROW end_ARG ] (12)

their Pauli representations are found as

HN=3A2=2.0+0.8660254X0+1.0Z0subscriptsuperscript𝐻2𝑁3𝐴2.00.8660254subscript𝑋01.0subscript𝑍0H^{2}_{N=3A}=2.0+0.8660254X_{0}+1.0Z_{0}italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT = 2.0 + 0.8660254 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 1.0 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (13)

and

HN=3B2=2.00.8660254X01.0Z0subscriptsuperscript𝐻2𝑁3𝐵2.00.8660254subscript𝑋01.0subscript𝑍0H^{2}_{N=3B}=2.0-0.8660254X_{0}-1.0Z_{0}italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N = 3 italic_B end_POSTSUBSCRIPT = 2.0 - 0.8660254 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1.0 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (14)

IV.2 Circuit Ansatz

From the submatrices presented in Eqs. (7) and (8), or their Pauli form in Eqs. (9) and (10), it is clear that the wavefunctions of these submatrices can be expressed using a single qubit ansatz. Thus we use a one-parameter circuit, shown in Fig. 2, to cover the Hilbert space to find both eigenvalues associated with the respective submatrix being measured.

IV.3 Simulation

We perform variance minimisations, starting from random initial parameters, using noiseless simulations. The results of these simulations are plotted in Fig. 1 for the submatrix HN=3Asubscript𝐻𝑁3𝐴H_{N=3A}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT, which show the energy values converging to their respective exact eigenvalues.

Refer to caption
Figure 1: Iterative variance minimisation to converge on energy eigenstates (dashed lines) of the HN=3Asubscript𝐻𝑁3𝐴H_{N=3A}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT sector of the N=3𝑁3N=3italic_N = 3 LMG model with parameters Vϵ=0.5𝑉italic-ϵ0.5\frac{V}{\epsilon}=0.5divide start_ARG italic_V end_ARG start_ARG italic_ϵ end_ARG = 0.5, Wϵ=0𝑊italic-ϵ0\frac{W}{\epsilon}=0divide start_ARG italic_W end_ARG start_ARG italic_ϵ end_ARG = 0, using a simulator.

IV.4 Quantum Hardware Results

We perform the calculations of the N=3𝑁3N=3italic_N = 3 LMG excited state spectrum using a parameter sweep, since this is feasible for a single parameter and will help visualise the results. For this, we use IBM’s cloud-based quantum computer, IMBQ_manila. This is a 5-qubit quantum computer using superconducting circuits and a “nearest neighbour” entanglement scheme (though for this N=3𝑁3N=3italic_N = 3 case we use only a single qubit). We perform the calculations using 20,000 measurements per circuit (shots), and one circuit for each term in the Hamiltonians. We also introduce additional circuits to mitigate some of the measurement readout bias that is innate to the quantum computer. These additional circuits prepare a state that represents each possible output of the quantum computer, using Pauli X gates. For a single qubit, this method of mitigation adds two additional circuits, one measuring the state |0ket0|0\rangle| 0 ⟩ and the other measuring the state |1ket1|1\rangle| 1 ⟩, per iteration step of the VQE. This allows for up-to-date corrections to be applied to our counts [40]. We also perform VQE variance minimisations, starting from random initial parameters using IBMQ_manila as described in section III.

We present the results of the calculations in Fig. 3, which shows the parameter sweep, in Fig. 4 which shows the variance minimisations, and in Table 1, which shows the resulting states for which the Hamiltonian variance is zero. These results obtained on a quantum computer are close to the exact values for the respective eigenstates, within error.

{quantikz}
\lstick\ket

0 & \qw \gateR_y(θ) \qw \qw

Figure 2: Single-qubit circuit ansatz.
Refer to caption
Figure 3: Parameter sweep of the block-diagonal submatrices, A𝐴Aitalic_A and B𝐵Bitalic_B, for the N=3𝑁3N=3italic_N = 3 LMG model, using 50 iteration steps, using quantum hardware.
Refer to caption
Figure 4: Iterative variance minimisation to converge on energy eigenstates (dashed lines) of the HN=3Asubscript𝐻𝑁3𝐴H_{N=3A}italic_H start_POSTSUBSCRIPT italic_N = 3 italic_A end_POSTSUBSCRIPT sector of the N=3𝑁3N=3italic_N = 3 LMG model with parameters Vϵ=0.5𝑉italic-ϵ0.5\frac{V}{\epsilon}=0.5divide start_ARG italic_V end_ARG start_ARG italic_ϵ end_ARG = 0.5, Wϵ=0𝑊italic-ϵ0\frac{W}{\epsilon}=0divide start_ARG italic_W end_ARG start_ARG italic_ϵ end_ARG = 0, using quantum hardware. Random starting parameters are used, and two illustrative results are shown, resulting in the full spectrum.
Table 1: N=3𝑁3N=3italic_N = 3 LMG model block-diagonal submatrix results calculated using the IBMQ_manila quantum computer using the circuit in Fig. 2 using 20000 shots. The Ground and second excited state results are achieved using submatrix A, with submatrix B used for the first and third excited states.
Eigenstate Exact Value (ϵitalic-ϵ\epsilonitalic_ϵ) Variance QC Result (ϵitalic-ϵ\epsilonitalic_ϵ)
Ground 1.8231.823-1.823- 1.823 0.073 -1.788 ±plus-or-minus\pm± 0.062
2nd 0.8230.8230.8230.823 0.001 0.826 ±plus-or-minus\pm± 0.064
1st 0.8230.823-0.823- 0.823 -0.004 -0.816 ±plus-or-minus\pm± 0.063
3rd 1.8231.8231.8231.823 0.001 1.810 ±plus-or-minus\pm± 0.063

V N=7𝑁7N=7italic_N = 7 Simulation and Measurement

V.1 Ansatz

The N=7𝑁7N=7italic_N = 7 submatrix featuring the lowest unperturbed eigenvalue in the diagonal for our choice of V/ϵ=0.5𝑉italic-ϵ0.5V/\epsilon=0.5italic_V / italic_ϵ = 0.5, W=0𝑊0W=0italic_W = 0 is

HN=7 Submatrix=[3.52.291002.2911.53.873003.8730.53.354003.3542.5].subscript𝐻𝑁7 Submatrixmatrix3.52.291002.2911.53.873003.8730.53.354003.3542.5H_{N=7\text{ Submatrix}}=\begin{bmatrix}-3.5&-2.291&0&0\\ -2.291&-1.5&-3.873&0\\ 0&-3.873&0.5&-3.354\\ 0&0&-3.354&2.5\\ \end{bmatrix}.italic_H start_POSTSUBSCRIPT italic_N = 7 Submatrix end_POSTSUBSCRIPT = [ start_ARG start_ROW start_CELL - 3.5 end_CELL start_CELL - 2.291 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL - 2.291 end_CELL start_CELL - 1.5 end_CELL start_CELL - 3.873 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - 3.873 end_CELL start_CELL 0.5 end_CELL start_CELL - 3.354 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - 3.354 end_CELL start_CELL 2.5 end_CELL end_ROW end_ARG ] . (15)

When encoded into the form of Pauli spin matrices, this becomes

HN=7=subscript𝐻𝑁7absent\displaystyle H_{N=7}=italic_H start_POSTSUBSCRIPT italic_N = 7 end_POSTSUBSCRIPT = 0.52.8225X11.0Z11.9365X0X10.52.8225subscript𝑋11.0subscript𝑍11.9365subscript𝑋0subscript𝑋1\displaystyle-0.5-2.8225X_{1}-1.0Z_{1}-1.9365X_{0}X_{1}- 0.5 - 2.8225 italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 1.0 italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 1.9365 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
1.9365Y0Y12.0Z0+0.5315Z0X1.1.9365subscript𝑌0subscript𝑌12.0subscript𝑍00.5315subscript𝑍0subscript𝑋1\displaystyle-1.9365Y_{0}Y_{1}-2.0Z_{0}+0.5315Z_{0}X_{1}.- 1.9365 italic_Y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 2.0 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 0.5315 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT . (16)

As with the N=3𝑁3N=3italic_N = 3 case the eigenvalues of the opposite parity submatrix can be obtained from those of the above matrix through multiplication by 11-1- 1.

From Eq. (V.1), a two-qubit circuit is required to represent the wave function of the system. A sufficiently general circuit for this case, with three parameters, is shown in Fig. 5.

{quantikz}\lstick\ket

0 & \qw \targ \gateR_y(θ_1) \targ\qw\qw
\lstick\ket0 \gateR_y(θ_0) \ctrl-1\qw \ctrl-1 \gateR_y(θ_2) \qw

Figure 5: Quantum circuit to calculate the eigenvalues of LMG model N=7𝑁7N=7italic_N = 7 Hamiltonian.

For the calculation of the variance, the square of HN=7subscript𝐻𝑁7H_{N=7}italic_H start_POSTSUBSCRIPT italic_N = 7 end_POSTSUBSCRIPT is needed. Matrix multiplication gives

HN=72=[17.49868111.4558.8730430.11.45522.498813.87312.9900428.8730433.87326.49944510.0620.12.99004210.06217.499316],subscriptsuperscript𝐻2𝑁7matrix17.49868111.4558.873043011.45522.498813.87312.9900428.8730433.87326.49944510.062012.99004210.06217.499316H^{2}_{N=7}=\begin{bmatrix}17.498681&11.455&8.873043&0.\\ 11.455&22.49881&3.873&12.990042\\ 8.873043&3.873&26.499445&-10.062\\ 0.&12.990042&-10.062&17.499316\\ \end{bmatrix},italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N = 7 end_POSTSUBSCRIPT = [ start_ARG start_ROW start_CELL 17.498681 end_CELL start_CELL 11.455 end_CELL start_CELL 8.873043 end_CELL start_CELL 0 . end_CELL end_ROW start_ROW start_CELL 11.455 end_CELL start_CELL 22.49881 end_CELL start_CELL 3.873 end_CELL start_CELL 12.990042 end_CELL end_ROW start_ROW start_CELL 8.873043 end_CELL start_CELL 3.873 end_CELL start_CELL 26.499445 end_CELL start_CELL - 10.062 end_CELL end_ROW start_ROW start_CELL 0 . end_CELL start_CELL 12.990042 end_CELL start_CELL - 10.062 end_CELL start_CELL 17.499316 end_CELL end_ROW end_ARG ] , (17)

from which Pauli decomposition produces the form needed for quantum computation as

HN=72=subscriptsuperscript𝐻2𝑁7absent\displaystyle H^{2}_{N=7}=italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N = 7 end_POSTSUBSCRIPT = 20.999063+0.6965X1+1.0Z1+10.9315425X020.9990630.6965subscript𝑋11.0subscript𝑍110.9315425subscript𝑋0\displaystyle 20.999063+0.6965X_{1}+1.0Z_{1}+10.9315425X_{0}20.999063 + 0.6965 italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 1.0 italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 10.9315425 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
+1.9365X0X12.0584995X0Z1+1.93654Y0Y11.9365subscript𝑋0subscript𝑋12.0584995subscript𝑋0subscript𝑍11.93654subscript𝑌0subscript𝑌1\displaystyle+1.9365X_{0}X_{1}-2.0584995X_{0}Z_{1}+1.93654Y_{0}Y_{1}+ 1.9365 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 2.0584995 italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 1.93654 italic_Y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
1.0003175Z0+10.7585Z0X13.5Z0Z1.1.0003175subscript𝑍010.7585subscript𝑍0subscript𝑋13.5subscript𝑍0subscript𝑍1\displaystyle-1.0003175Z_{0}+10.7585Z_{0}X_{1}-3.5Z_{0}Z_{1}.- 1.0003175 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 10.7585 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 3.5 italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT . (18)

V.2 Simulation

Similarly to the N=3𝑁3N=3italic_N = 3 simulation in Section IV.3, we perform noiseless simulations of the variance minimisation to find each eigenvalue. The results of these simulations for the N=7𝑁7N=7italic_N = 7 LMG model are plotted in Fig. 6, which shows convergence across each of the four simulations to their respective exact eigenstate. These simulations are also started from randomised initial parameters.

Refer to caption
Figure 6: Energy values during iteration of the simulated variance minimisation method for the N=7𝑁7N=7italic_N = 7 LMG model with parameters Vϵ=0.5𝑉italic-ϵ0.5\frac{V}{\epsilon}=0.5divide start_ARG italic_V end_ARG start_ARG italic_ϵ end_ARG = 0.5, Wϵ=0𝑊italic-ϵ0\frac{W}{\epsilon}=0divide start_ARG italic_W end_ARG start_ARG italic_ϵ end_ARG = 0. Known exact eigenvalues are shown with dashed lines. Four random starting points are shown corresponding to runs which find the four eigenstates.

V.3 Quantum Hardware Results

Measurements for the N=7𝑁7N=7italic_N = 7 LMG excitation spectra were performed using the IBM_nairobi quantum computer, again using 20,000 shots per circuit. We apply the same method of readout error mitigation as with the N=3𝑁3N=3italic_N = 3 case, described in Section IV.4, as well as applying a method of CNOT mitigation. This mitigation technique adds pairs of CNOT gates to the circuit where there is a CNOT present, allowing for a linear extrapolation to the theoretical case where there are no CNOT gates in the circuit. Adding additional pairs of CNOT gates does not alter the value of the measurement, other than introducing additional error from the added CNOT pairs [41]. The results of the iterative variance minimizations on IBM_nairobi are presented in Fig. 7 and in Table 2. The results are close to the known, exact solutions. However, these results are further from their respective exact solutions than the N=3𝑁3N=3italic_N = 3 results. This is likely due to the increase in circuit depth required to perform the calculations for the N=7𝑁7N=7italic_N = 7 case, as well as the additional qubit, resulting in greater gate error and noise within the circuit.

Refer to caption
Figure 7: Energy values during iteration of the variance minimisation method on quantum hardware for the N=7𝑁7N=7italic_N = 7 LMG model with parameters Vϵ=0.5𝑉italic-ϵ0.5\frac{V}{\epsilon}=0.5divide start_ARG italic_V end_ARG start_ARG italic_ϵ end_ARG = 0.5, Wϵ=0𝑊italic-ϵ0\frac{W}{\epsilon}=0divide start_ARG italic_W end_ARG start_ARG italic_ϵ end_ARG = 0. Known exact eigenvalues are shown with dashed lines. Four random starting points are shown corresponding to runs which find the four eigenstates.
Table 2: N=7𝑁7N=7italic_N = 7 LMG model results calculated using the IBMQ_manila quantum computer using 20000 shots.
Eigenstate Exact Value (ϵitalic-ϵ\epsilonitalic_ϵ) Variance QC Result (ϵitalic-ϵ\epsilonitalic_ϵ)
Ground 6.2086.208-6.208- 6.208 0.139 -6.067 ±plus-or-minus\pm± 0.901
1st 2.9442.944-2.944- 2.944 0.016 -3.151 ±plus-or-minus\pm± 0.503
2nd 1.2081.2081.2081.208 0.010 1.184 ±plus-or-minus\pm± 0.484
3rd 5.9445.9445.9445.944 0.114 5.902 ±plus-or-minus\pm± 0.660

In order to explore further the quality of the solutions obtained we calculate the fidelity, or overlap of the wavefunctions obtained by the quantum computer with the known analytic wavefunctions for each eigenvalue of the N=7𝑁7N=7italic_N = 7 LMG model. These overlaps are given by

ψA|ψQC2superscriptinner-productsubscript𝜓𝐴subscript𝜓𝑄𝐶2\langle\psi_{A}|\psi_{QC}\rangle^{2}⟨ italic_ψ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_Q italic_C end_POSTSUBSCRIPT ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (19)

where ψAsubscript𝜓𝐴\psi_{A}italic_ψ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT is the wavefunction of the analytic solution, and ψQCsubscript𝜓𝑄𝐶\psi_{QC}italic_ψ start_POSTSUBSCRIPT italic_Q italic_C end_POSTSUBSCRIPT is the wavefunction determined by the quantum computer. The results of the overlap calculations are shown in Table 3.

Table 3: Overlaps of the N=7𝑁7N=7italic_N = 7 LMG model result wavefunctions with analytic solution wavefunctions.
Measured Eigenstate 6.26.2-6.2- 6.2 2.942.94-2.94- 2.94 1.21.21.21.2 5.945.945.945.94
Ground 0.9240.9240.9240.924 0.0730.0730.0730.073 0.0020.0020.0020.002 0.0010.0010.0010.001
1st 0.0200.0200.0200.020 0.9470.9470.9470.947 0.0330.0330.0330.033 <0.001absent0.001<0.001< 0.001
2nd 0.0010.0010.0010.001 0.0090.0090.0090.009 0.9910.9910.9910.991 <0.001absent0.001<0.001< 0.001
3rd 0.0010.0010.0010.001 <0.001absent0.001<0.001< 0.001 0.0260.0260.0260.026 0.9720.9720.9720.972

Here we see that all results found using variance minimization on quantum hardware find states that closely overlap with their respective known analytic eigenvector. The ground state introduces slightly more overlap with the first excited state, which could be due to only achieving a partial convergence in the time available for the run on real quantum hardware, as can be seen in Fig. 7.

VI Conclusion

We adapted the Variational Quantum Eigensolver (VQE) to target any eigenstate of a Hamiltonian through minimization of the variance of the Hamiltonian. The cost is in the form of increased circuit evaluations compared with a ground state VQE, but not circuit depth or number of qubits. States close to zero variance could be found on presently-available quantum computing hardware, corresponding to the known eigenstates of a model Hamiltonian. Further exploration of larger problems would be needed to check how reliably one could extract a full spectrum with the necessarily increasingly complex circuits needed to parameterize larger problems.

Acknowledgements.
This work was funded by AWE. We acknowledge the use of IBM Quantum services for this work. The views expressed are those of the authors, and do not reflect the official policy or position of IBM or the IBM Quantum team. In this paper we used IBMQ_manilla and IBMQ_nairobi, which are IBM Quantum Processors. PDS acknowledges support from the UK STFC under grant numbers ST/V001108/1 and ST/W006472/1. UK Ministry of Defence ©Crown owned copyright 2024/AWE.

References

  • Višňák [2015] J. Višňák, Quantum algorithms for computational nuclear physics, EPJ Web of Conferences 100, 01008 (2015).
  • Višňák and Veselý [2017] J. Višňák and P. Veselý, Quantum algorithms for computational nuclear physics revisited, particular case of second quantized formulation, EPJ Web of Conferences 154, 01030 (2017).
  • Dumitrescu et al. [2018] E. F. Dumitrescu, A. J. McCaskey, G. Hagen, G. R. Jansen, T. D. Morris, T. Papenbrock, R. C. Pooser, D. J. Dean, and P. Lougovski, Cloud Quantum Computing of an Atomic Nucleus, Physical Review Letters 120, 210501 (2018).
  • Roggero et al. [2020a] A. Roggero, C. Gu, A. Baroni, and T. Papenbrock, Preparation of excited states for nuclear dynamics on a quantum computer, Physical Review C 102, 1 (2020a).
  • Roggero et al. [2020b] A. Roggero, A. C. Li, J. Carlson, R. Gupta, and G. N. Perdue, Quantum computing for neutrino-nucleus scattering, Physical Review D 101, 74038 (2020b).
  • Lacroix [2020] D. Lacroix, Symmetry-Assisted Preparation of Entangled Many-Body States on a Quantum Computer, Physical Review Letters 125, 230502 (2020).
  • Siwach and Lacroix [2021] P. Siwach and D. Lacroix, Filtering states with total spin on a quantum computer, Physical Review A 104, 062435 (2021).
  • Baroni et al. [2022] A. Baroni, J. Carlson, R. Gupta, A. C. Y. Li, G. N. Perdue, and A. Roggero, Nuclear two point correlation functions on a quantum computer, Phys. Rev. D 105, 074503 (2022).
  • Robin et al. [2021] C. Robin, M. J. Savage, and N. Pillet, Entanglement rearrangement in self-consistent nuclear structure calculations, Physical Review C 103, 034325 (2021).
  • Zhang et al. [2021] D.-B. Zhang, H. Xing, H. Yan, E. Wang, and S.-L. Zhu, Selected topics of quantum computing for nuclear physics, Chinese Physics B 30, 020306 (2021).
  • Robbins and Love [2021] K. Robbins and P. J. Love, Benchmarking near-term quantum devices with the variational quantum eigensolver and the lipkin-meshkov-glick model, Phys. Rev. A 104, 022412 (2021).
  • Pérez-Fernández et al. [2022] P. Pérez-Fernández, J.-M. Arias, J.-E. García-Ramos, and L. Lamata, A digital quantum simulation of the Agassi model, Physics Letters B , 137133 (2022).
  • Cervia et al. [2021] M. J. Cervia, A. B. Balantekin, S. N. Coppersmith, C. W. Johnson, P. J. Love, C. Poole, K. Robbins, and M. Saffman, Lipkin model on a quantum computer, Physical Review C 104, 024305 (2021).
  • Kiss et al. [2022] O. Kiss, M. Grossi, P. Lougovski, F. Sanchez, S. Vallecorsa, and T. Papenbrock, Quantum computing of the 66{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPTLi nucleus via ordered unitary coupled clusters, Physical Review C 106, 034325 (2022).
  • Romero et al. [2022] A. M. Romero, J. Engel, H. L. Tang, and S. E. Economou, Solving nuclear structure problems with the adaptive variational quantum algorithm, Physical Review C 105, 064317 (2022).
  • Chikaoka and Liang [2022] A. Chikaoka and H. Liang, Quantum computing for the Lipkin model with unitary coupled cluster and structure learning ansatz, Chinese Physics C 46, 024106 (2022).
  • Hlatshwayo et al. [2022] M. Q. Hlatshwayo, Y. Zhang, H. Wibowo, R. LaRose, D. Lacroix, and E. Litvinova, Simulating excited states of the Lipkin model on a quantum computer, Physical Review C 106, 024319 (2022).
  • Ruiz Guzman and Lacroix [2022] E. A. Ruiz Guzman and D. Lacroix, Accessing ground-state and excited-state energies in a many-body system after symmetry restoration using quantum computers, Physical Review C 105, 024324 (2022).
  • Hobday et al. [2022] I. Hobday, P. D. Stevenson, and J. Benstead, Quantum computing calculations for nuclear structure and nuclear data, in Quantum Technologies 2022, Vol. 12133, International Society for Optics and Photonics (SPIE, 2022) p. 109.
  • Stetcu et al. [2022] I. Stetcu, A. Baroni, and J. Carlson, Variational approaches to constructing the many-body nuclear ground state for quantum computing, Phys. Rev. C 105, 064308 (2022).
  • Li et al. [2023] Y. H. Li, J. Al-Khalili, and P. Stevenson, A quantum simulation approach to implementing nuclear density functional theory via imaginary time evolution (2023), arXiv:2308.15425 .
  • Feynman [1982] R. P. Feynman, Simulating physics with computers, International Journal of Theoretical Physics 21, 467 (1982).
  • O’Brien et al. [2019] T. E. O’Brien, B. Tarasinski, and B. M. Terhal, Quantum phase estimation of multiple eigenvalues for small-scale (noisy) experiments, New Journal of Physics 21, 023022 (2019).
  • Zhou et al. [2013] X.-Q. Zhou, P. Kalasuwan, T. C. Ralph, and J. L. O’brien, Calculating unknown eigenvalues with a quantum algorithm, Nature photonics 7, 223 (2013).
  • Du et al. [2021] W. Du, J. P. Vary, X. Zhao, and W. Zuo, Ab initio nuclear structure via quantum adiabatic algorithm (2021), arXiv:2105.08910 .
  • Motta et al. [2020] M. Motta, C. Sun, A. T. K. Tan, M. J. O’Rourke, E. Ye, A. J. Minnich, F. G. S. L. Brandão, and G. K.-L. Chan, Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution, Nature Physics 16, 205 (2020).
  • Turro et al. [2022] F. Turro, A. Roggero, V. Amitrano, P. Luchi, K. A. Wendt, J. L. Dubois, S. Quaglioni, and F. Pederiva, Imaginary-time propagation on a quantum chip, Phys. Rev. A 105, 022440 (2022).
  • Jouzdani et al. [2022] P. Jouzdani, C. W. Johnson, E. R. Mucciolo, and I. Stetcu, Alternative approach to quantum imaginary time evolution, Phys. Rev. A 106, 062435 (2022).
  • Preskill [2018] J. Preskill, Quantum computing in the NISQ era and beyond, Quantum 2, 79 (2018).
  • 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, A variational eigenvalue solver on a photonic quantum processor, Nature Communications 5, 4213 (2014).
  • McClean et al. [2016] J. R. McClean, J. Romero, R. Babbush, and A. Aspuru-Guzik, The theory of variational hybrid quantum-classical algorithms, New Journal of Physics 18, 023023 (2016).
  • Tilly et al. [2022] J. Tilly, H. Chen, S. Cao, D. Picozzi, K. Setia, Y. Li, E. Grant, L. Wossnig, I. Rungger, G. H. Booth, and J. Tennyson, The variational quantum eigensolver: a review of methods and best practices (2022).
  • Lipkin et al. [1965] H. Lipkin, N. Meshkov, and A. Glick, Validity of many-body approximation methods for a solvable model: (i). exact solutions and perturbation theory, Nuclear Physics 62, 188 (1965).
  • Agassi et al. [1966] D. Agassi, H. J. Lipkin, and N. Meshkov, Validity of many-body approximation methods for a solvable model: (IV). The deformed Hartree-Fock solution, Nuclear Physics 86, 321 (1966).
  • Jordan and Wigner [1928] P. Jordan and E. Wigner, Über das Paulische Äquivalenzverbot, Zeitschrift für Physik 47, 631 (1928).
  • Bravyi and Kitaev [2002] S. B. Bravyi and A. Y. Kitaev, Fermionic quantum computation, Annals of Physics 298, 210 (2002).
  • Seeley et al. [2012] J. T. Seeley, M. J. Richard, and P. J. Love, The Bravyi-Kitaev transformation for quantum computation of electronic structure, The Journal of Chemical Physics 137, 224109 (2012).
  • Pesce and Stevenson [2021] R. M. N. Pesce and P. D. Stevenson, H2ZIXY: Pauli spin matrix decomposition of real symmetric matrices (2021), arXiv:2111.00627 .
  • Bergholm et al. [2022] V. Bergholm, J. Izaac, M. Schuld, C. Gogolin, S. Ahmed, V. Ajith, M. S. Alam, G. Alonso-Linaje, B. AkashNarayanan, A. Asadi, J. M. Arrazola, U. Azad, S. Banning, C. Blank, T. R. Bromley, B. A. Cordier, J. Ceroni, A. Delgado, O. D. Matteo, A. Dusko, T. Garg, D. Guala, A. Hayes, R. Hill, A. Ijaz, T. Isacsson, D. Ittah, S. Jahangiri, P. Jain, E. Jiang, A. Khandelwal, K. Kottmann, R. A. Lang, C. Lee, T. Loke, A. Lowe, K. McKiernan, J. J. Meyer, J. A. Montañez-Barrera, R. Moyard, Z. Niu, L. J. O’Riordan, S. Oud, A. Panigrahi, C.-Y. Park, D. Polatajko, N. Quesada, C. Roberts, N. Sá, I. Schoch, B. Shi, S. Shu, S. Sim, A. Singh, I. Strandberg, J. Soni, A. Száva, S. Thabet, R. A. Vargas-Hernández, T. Vincent, N. Vitucci, M. Weber, D. Wierichs, R. Wiersema, M. Willmann, V. Wong, S. Zhang, and N. Killoran, Pennylane: Automatic differentiation of hybrid quantum-classical computations (2022), arXiv:1811.04968 .
  • Kandala et al. [2017] A. Kandala, A. Mezzacapo, K. Temme, M. Takita, M. Brink, J. M. Chow, and J. M. Gambetta, Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets, Nature 549, 242 (2017).
  • Li and Benjamin [2017] Y. Li and S. C. Benjamin, Efficient variational quantum simulator incorporating active error minimization, Phys. Rev. X 7, 021050 (2017).